DOI:
10.1039/D4TA04550E
(Paper)
J. Mater. Chem. A, 2024, Advance Article
Frustrated Lewis pairs on metal-cation vacancy catalysts enhanced the electroreduction of NO to NH3†
Received
2nd July 2024
, Accepted 1st August 2024
First published on 5th August 2024
Abstract
Electrocatalytic reduction of nitrogen oxide (NOx) in the industrial flue gas to synthesize NH3 can not only greatly improve the electrochemical efficiency of NH3 production, but also realize the emission reduction and resource utilization of air pollutants. Purposeful structural optimization of catalysts is a key research challenge in this field. Density functional theory (DFT) calculations revealed that the formation of surface oxygen vacancies could be promoted by constructing metal cation vacancies on the catalyst surface, which in turn optimized the NO adsorption configuration and improved the catalytic activity. By this guidance, CoAl layered double hydroxide (CoAl-LDH) nanosheets containing Al atomic vacancies were prepared by alkali etching. Experimental results show that a suitable alkali etching time with aqueous NaOH can significantly enhance the electrocatalytic activity of the CoAl-LDH nanosheets for NO reduction to NH3. Compared with pristine CoAl-LDH (0.097 mg cm−2 h−1), the NH3 yield rate reached 0.310 mg cm−2 h−1 at −1.1 V vs. the reversible hydrogen electrode after alkali etching for 4 h, which increased by 2.2 times. Catalyst characterization and theoretical calculation results indicate that the abundant frustrated Lewis pairs (FLPs) on the surface of catalysts containing cation vacancies are the origin of the enhanced activity.
1. Introduction
NOx emissions resulting from fossil fuel combustion are a significant contributor to industrial emissions and remain a primary cause of environmental issues such as acid rain, photochemical smog, ozone depletion, global warming, and a serious threat to human health.1–3 NO, which accounts for up to 95% of NOx, is the dominant component of industrial flue gas.4 Converting NO from flue gas into harmless or valuable forms of N is crucial.2,5 Selective catalytic reduction (SCR) is the most widely used and established technology worldwide for flue gas denitrification. However, the catalytic process occurs at high temperatures and employs precious NH3 as a catalyst, which does not meet environmental friendliness requirements.6,7 Consequently, developing alternative denitrification processes that are sustainable, energy-efficient, and environmentally friendly is crucial.
NH3 is a zero-carbon fuel, a possible future hydrogen energy source, and a crucial chemical for manufacturing nitrogen fertilizers, paints, plastics, and more.8,9 Industrial NH3 is primarily produced by the highly energy-intensive Haber–Bosch process, which is responsible for approximately 1.5% of total global carbon dioxide (CO2) emissions and consumes 2% of the worldwide annual energy supply.10–12 The electrochemical nitrogen reduction reaction (NRR) is considered a promising alternative to the Haber–Bosch process, but its development has been impeded by the challenges associated with activating N2 and its limited solubility in the electrolyte.13,14 Nitrate possesses higher water solubility and lower NO bond dissociation energy, and electrocatalytic nitrate reduction to synthesize NH3 for simultaneous nitrate wastewater treatment and NH3 production has also been widely studied.15–21 Some researchers have ingeniously employed the NO reduction reaction (NORR) for electrosynthesis of NH3, which not only addresses the removal of NO but also generates NH3 as a value-added chemical.16,22 This highlights the significant potential of the NORR for practical applications.
Layered double hydroxides (LDHs) have garnered considerable attention as an essential catalyst in the areas of adsorption,23 catalysis,24 energy storage,25 and electrochemistry26 owing to their low raw material cost, versatility, customizable chemical composition, and unique layered structure. Nevertheless, the absence of accessible active sites and the low conductivity of the native LDH itself impede further enhancement of the catalytic performance. Fortunately, these deficiencies can be rectified through the precise adjustment of the electronic structure.27 Generating cationic vacancies has recently been identified as an effective method for modulating the electronic configuration of the host material and the coordination environment of the metal sites, thereby tuning the electrocatalytic properties. These defects facilitate charge transfer and redox reaction kinetics and can also serve as additional host sites for the insertion of protons or alkali metal cations, thereby promoting ion diffusion during electrochemical cycling.28,29 Cationic vacancies are regarded as a promising avenue for regulating catalytic selectivity.30
In recent years, frustrated Lewis pairs (FLPs) constructed on catalyst surfaces have demonstrated significant potential for adsorption and activation of small molecules such as H2,31 CO2,32–34 N2,35,36 and others. Lewis acids (LAs) with empty orbitals and Lewis bases (LBs) with lone-pair electrons in FLPs provide opportunities for dissociation and activation of gas molecules due to large steric hindrance formed by active regions instead of coordination bonds.37 Research has demonstrated that FLPs may comprise oxygen vacancies (VO) and neighboring hydroxyl (–OH) or lattice oxygen atoms, creating distinct reaction sites for catalytic activation.38,39 Defect engineering represents a viable strategy in the design of FLPs for metal compounds that contain both a metal cation as the LA and an oxygen anion as the LB.38,40
Inspired by this, metal cation vacancies were constructed on the catalyst surface by a simple alkali etching strategy to promote the generation of surface VO and protons, and FLPs were successfully constructed in which the unsaturated Al sites on the surface of LDH and the neighboring –OH species acted as the LA and LB sites, respectively. The LA provided the empty orbitals and the LB provided the protons for the adsorption and activation of NO molecules, which enhanced the electrocatalytic NORR synthesis of NH3 activity. As a result, the etched nanosheets exhibited better performance in NH3 synthesis compared to pristine CoAl-LDH. Specifically, after alkali etching for 4 h, the NH3 yield increased to 0.310 mg cm−2 h−1, which was 2.2 times higher than that of the pristine CoAl-LDH.
2. Experiments and calculations
2.1. Materials and chemicals
Cobaltous nitrate hexahydrate (Co(NO3)2·6H2O), aluminum nitrate nonahydrate (Al(NO3)3·9H2O) and sodium nitroferricyanide dihydrate (C5FeN6Na2O·2H2O) were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. Sodium hydroxide (NaOH) and urea (CH4N2O) were obtained from Sinopharm Chemical Reagent Co., Ltd (China). Salicylic acid (C7H6O3), ammonium chloride (NH4Cl), sodium citrate dihydrate (C6H5Na3O7·2H2O), and sodium hypochlorite solution (NaClO) were purchased from Shanghai Macklin Biochemical Co., Ltd (China). N2, Ar, and mixed gas of NO/Ar (1 vol% NO) were bought from Jing Hui Gas (China). Deionized water was used in all the experimental processes. All the chemicals were of analytical grade and used without further purification.
2.2. Catalyst preparation
2.2.1. Synthesis of CoAl-LDH nanosheets. CoAl-LDH nanosheets were prepared by the coprecipitation method. 9 mmol Co(NO3)2·6H2O, 3 mmol Al(NO3)3·9H2O and 60 mmol urea were dissolved in 75 mL of deionized water. After stirring for 0.5 h at room temperature, a clear solution was obtained. The solution was transferred to a 100 mL Teflon-lined stainless steel autoclave, sealed and heated at 120 °C for 14 h, and then naturally cooled to room temperature. The products were filtered and washed repeatedly with deionized water until the supernatant was neutral. Finally, the products were collected by freeze-drying.
2.2.2. Synthesis of etched CoAl -LDH nanosheets. CoAl-LDH nanosheets (50 mg) were dispersed in 0.5 M NaOH (75 mL). The solution was transferred to a 100 mL Teflon-lined stainless steel autoclave, sealed, and heated at 120 °C for 2 h, 4 h, 6 h, 8 h and 10 h, respectively, and then naturally cooled to room temperature. The products were centrifuged, washed thoroughly with deionized water, and dried at 80 °C.
2.3. Materials characterization
The crystalline phases were identified using a Rigaku Smartlab 3KW X-ray diffraction (XRD) instrument with Cu Kα radiation of 0.15418 nm wavelength (40 kV voltage, 100 mA current) at a scan rate of 10° min−1 within the 2θ range of 10–80°. The morphology of the prepared samples was characterized using a Zeiss Gemini 300 scanning electron microscope (SEM). Fourier transform infrared spectroscopy (FTIR) was carried out on the iS10 FTIR spectrometer of Negoli, with a wave number range of 400–4000 cm−1, a resolution of 4 cm−1, and 32 scans. X-ray photoelectron spectroscopy (XPS) (Thermo Scientific K-Alpha) was employed for elemental mapping using a monochromatic Al Kα radiation source. Binding energy calibration was performed by referencing the C 1s main peak at 284.8 eV. The elemental ratio was investigated by inductively coupled plasma optical emission spectrometry (ICP-OES). N2 adsorption–desorption isotherms were measured on an Autosorb iQ, Quantachrome Instruments, US analyzer. Pore volumes, sizes, and specific surface areas were determined by the Brunauer–Emmett–Teller (BET) method. Temperature-programmed desorption of NH3/CO2 (NH3/CO2-TPD) was performed using an Extrel MAX300 chemisorption analysis tester. K-edge X-ray absorption fine structure (XAFS) analysis was performed at the Beijing Synchrotron Radiation Facility 4B9A beamline.
2.4. Electrochemical measurements
Electrochemical measurements were performed at ambient temperature in an H-type cell with a three-electrode configuration separated by a Nafion 117 membrane (Electrochemical Workstation, CHI 660E, CH Instruments Inc., Shanghai). The electrolyte solution was 0.1 M K2SO4. The membrane was treated first in H2O2 (5 wt%) aqueous solution at 80 °C for 1 h. Then, it was treated in 0.5 M H2SO4 for 1 h at 80 °C and finally in water for 4 h. The carbon cloth utilized in this experiment was purchased from Canrd (WOS 1011 type). The prepared catalyst loaded on the carbon cloth (1 × 1 cm2) was used as the working electrode, and a Pt plate and Ag/AgCl (saturated KCl electrolyte) were used as the counter and reference electrodes, respectively. The potential without iR compensation was converted to the RHE scale using the following equation (E(V vs. RHE) = E(V vs. Ag/AgCl) + 0.21 + 0.0591 × pH). The electrochemical activity of catalysts was measured for the NORR. Generally, 20 mg of catalyst was ultrasonically dispersed in 750 μL of isopropanol, 100 μL of Nafion solution (5 wt%, Du Pont), and 2 mL of water to form a homogeneous ink. Then 1000 μL of the ink was dropped on the surface of carbon cloth (1 × 3 cm2) and dried at room temperature. To eliminate possible alkaline contaminants, the feed gas was passed through a saturator filled with a 0.05 M H2SO4 solution to remove impurities from the inlet gas. Before the electrochemical characterization, the electrolyte was purged with 1% NO/Ar at a rate of 30 sccm for 30 min and kept constant during electrochemical testing. The cyclic voltammetry (CV) test was conducted with a scan rate of 50 mV s−1 ranging from 0 to −1.8 V vs. RHE. The linear sweep voltammetry (LSV) test was carried out with a scan rate of 5 mV s−1. Subsequently, the chronoamperometric test was conducted at various potentials, and the stability experiment was executed at a potential of 1.0 V vs. RHE.
2.5. Determination of NH3 concentration
The amount of NH3 in the solution was determined by colorimetry using the indophenol blue method.4 A certain amount of electrolyte was taken out from the electrolytic cell and diluted to 2 mL to the detection range. Then, 2 mL of a 1 M NaOH solution that contained salicylic acid and sodium citrate was added. Then, 1 mL of 0.05 M NaClO containing 1 M NaOH and 0.2 mL of 1 wt% C5FeN6Na2O·2H2O were added to the above solution. After standing at room temperature for 2 h, the UV-Vis absorption spectrum was measured. The concentration of NH3 was determined using the absorbance at a wavelength of 665 nm. The concentration–absorbance curve was calibrated using a series of standard NH4Cl solutions. The fitting curve (y = 0.4453x − 0.0083, R2 = 0.9998) showed a good linear relation of the absorbance value with NH3 concentrations.
2.6. Calculations of NH3 yield and Faraday efficiency
The average NH3 yield (mg cm−2 h−1) was calculated using the following equation:41
NH3 yield = 10−3 × C × V/(S × t) |
where C (mg L−1) is the measured mass concentration of produced NH3; V (100 mL) is the total volume of the cathodic electrolyte, S (1 cm2) is the geometric area of the working electrode, and t (h) is time for electrocatalysis.
The Faraday efficiency (FE) is the amount of charge used to produce NH3 during electrolysis divided by the total amount of charge passing through the electrode. The FE was calculated according to the following equation:2
FE (%) = (10−6 × n × F × C × V)/(17 × Q) × 100 |
where
n is the number of electrons required per mole of NH
3 produced;
F (96
485 C mol
−1) is the Faraday constant;
C (mg L
−1) is the measured mass concentration of NH
3;
V (100 mL) is the total volume of the cathodic electrolyte;
Q is the total charge across the working electrode.
2.7. Computational details
The atomic models of the functionalized LDHs were designed assuming their complete surface termination by OH atoms. The periodic system had a vacuum thickness of 15 Å, which was used to eliminate spurious interactions between the adsorbate and the periodic image of the bottom layer of the surface. All the computations were performed with spin-polarized DFT implemented in the DMol3 code of the Materials Studio. A generalized gradient approximation (GGA) with the Perdew, Burke, and Ernzerhof (PBE) exchange–correlation functionals was employed. The double numerical plus polarization (DNP) was used as the basis set. Effective core potentials were adopted as the core treatment to conduct a metal relativistic effect. The geometry convergence tolerance for energy change was 2 × 10−5 Ha, the max force was 0.004 Ha Å−1, and the max displacement was 0.005 Å. We chose 4.5 Å to be the real-space global orbital cutoff radius. Structural optimization was performed without any constraints. The Gibbs free energy change (ΔG) of each elemental step was calculated according to the CHE model, which used one-half of the chemical potential of hydrogen as the chemical potential of the proton–electron pairs. ΔG was calculated from the equation: ΔG = ΔE + ΔEZPE + ΔGU − TΔS, where ΔE is the electronic energy difference directly obtained from DFT calculations, ΔEZPE is the zero-point energy, T is the temperature (T = 298.15 K), ΔS is the change in entropy, and ΔGU represents the free energy contributions related to the applied electrode potential U.
3. Results and discussion
3.1. Effects of metal-cation vacancies on NO adsorption properties
Initially, DFT calculations were conducted to determine the activation of NO on the LDH surface and to investigate the crucial role of metal cation vacancies in facilitating NO adsorption. The active sites for NO adsorption on LDH are believed to be Al vacancies, and the adsorption energy of NO plays a vital role in the electrocatalytic synthesis of the NH3 reaction.42 The calculated effect of LDH surface defects on NO adsorption performance is shown in Fig. 1. LDH containing VO is more likely to attract NO molecules, providing active sites for molecular adsorption (Fig. 1a). The adsorption energy confirmed that LDH with VO has more negative adsorption energy than that free of VO, suggesting that NO adsorption is more favorable on LDH with VO (Fig. 1b). This is consistent with literature studies reporting that unsaturated metal sites are the ones that are more likely to adsorb small molecules.43–45 The higher adsorption energy could be attributed to the partial charge transfer of LDH and NO, thus weakening the N–O bond. Moreover, the effect of metal vacancies on the VO formation energy was also calculated, and it was found that the formation energy of the VO active site in LDH containing Al vacancies was lower (0.76 eV) than that of the defect-free LDH (1.59 eV) (Fig. 1c). Apparently, the Al vacancies in LDH are more favorable for the synthesis of VO active sites, which promotes the adsorption and activation of NO, thereby improving the reaction kinetics of electrocatalytic NO reduction to NH3.
|
| Fig. 1 Effect of material surface defects on NO adsorption properties. (a) Electrostatic potential of NO and LDH surfaces. (b) Adsorption energy of NO on LDH and LDH surfaces containing VO. (c) Effect of metal vacancies on the formation energy of VO. | |
3.2. Preparation and characterization of catalysts
DFT calculations reflected the advantage of Al vacancies in LDH electrocatalysts for NO adsorption reduction. The pristine CoAl-LDH sample was synthesized through a hydrothermal method. Subsequently, CoAl-LDH samples with Al vacancy defects were prepared by etching with NaOH aqueous solution for different durations using a facile alkali etching method (Fig. 2a). This paper presents the samples of pristine CoAl and etched CoAl (NaOH X h). The study investigated the changes in Co and Al atomic contents during the etching process using ICP-OES. The results showed that successful removal of partial Al sites and creation of metal cation defect sites were achieved after alkali etching of CoAl-LDH. This was evidenced by the significant decrease in Al content and increase in Co content compared to pristine CoAl (Fig. S1†).
|
| Fig. 2 Synthesis and characterization of CoAl-LDH nanosheets. (a) Schematic diagram of the synthetic procedure. (b) SEM patterns of the CoAl sample. (c) XRD and (d) FTIR patterns of CoAl, CoAl (NaOH 4 h) and CoAl (NaOH 10 h) samples. | |
The morphology of the CoAl-LDH samples was observed using SEM. The pristine CoAl exhibited a typical two-dimensional lamellar structure (Fig. 2b). It is noteworthy that the lamellar structure was maintained after 4 hours of alkali etching (CoAl (NaOH 4 h)) (Fig. S2a†). However, after 10 hours of alkali etching (CoAl (NaOH 10 h)), the catalyst structure was disrupted and etched into a flower-cluster-like shape (Fig. S2b†). The N2 adsorption–desorption experiments indicate that all samples exhibit type IV adsorption isotherms according to the IUPAC classification, displaying H3 hysteresis loops, which correspond to mesoporous materials (Fig. S3a†).46 The pore size distribution of most samples is around 3.82 nm (Fig. S3b†). The pore volume and surface area of the samples increased with the increase of alkali etching time (Table S1†). Will the high specific surface area of CoAl (NaOH 10 h) provide more active sites? This will be verified later in the electrochemical activity tests.
XRD was used to investigate the crystal structure of CoAl-LDH nanosheets (Fig. 2c). For pristine CoAl, typical diffraction peaks at 2θ = 11.6, 23.27, 34.42, 39.05, 46.48, 60.18, and 61.57 can be indexed to (003), (006), (012), (015), (018), (110), and (113), which correspond to the standard map JCPDS: 51-0045. These diffraction peaks and the absence of other peaks indicate the presence of a pure LDH phase.47–49 The XRD pattern exhibits good crystallinity of the prepared samples with a low and stable baseline and high-intensity diffraction peaks. The distinct (110) and (113) peaks are evidence of well-dispersed Co2+ and Al3+ metal cations on the pristine CoAl layer.49,50 Impressively, the (003) peaks weakened or even disappeared with increased etching time, indicating that the alkali etching treatment resulted in a smaller crystal size and a higher degree of disorder. The decrease in the intensity of the diffraction peaks of the etched CoAl-LDH samples is accompanied by the appearance of some stray peaks, which may be due to the diversification of some crystal surfaces by the lattice distortion caused by etching, resulting in the appearance of new diffraction peaks.
The functional groups present on the surface of CoAl-LDH nanosheets were identified using FTIR (Fig. 2d). The observed peaks at 3503 and 3382 cm−1 indicate the stretching and deformation vibrations of O–H, respectively.51 Similarly, the absorption peak at 3628 cm−1 is associated with the stretching vibration of H–O–H, and the peak at 2190 cm−1 is the bending vibration of NHO−.52 The bands at 1358 cm−1 in CoAl, CoAl (NaOH 4 h), and 1365 cm−1 in CoAl (NaOH 10 h) are attributed to the CO antisymmetric stretching of CO32−.53 The H2O bending vibration is observed at 1545 cm−1, while the remaining absorption bands below 800 cm−1 can be attributed to metal–oxygen (M–O) and metal–hydroxyl (M–OH) vibrations in the LDH.54 The peaks at 580 cm−1 and 668 cm−1 correspond to stretching vibrations of the Co–O and Al–O bonds respectively.55 The M–O peaks in CoAl (NaOH 4 h) and CoAl (NaOH 10 h) are significantly enhanced, which may be due to the presence of more VO resulting from cation vacancies after etching. This can be further confirmed by XPS.
The valence states of the elements in the CoAl-LDH nanosheets synthesized were investigated by using XPS. The corresponding results can be seen in Fig. 3. All elements have been normalized to carbon at 284.8 eV. Fig. S4† displays the composition of the CoAl-LDH nanosheets, mainly consisting of four elements, Co, Al, C, and O. Furthermore, Fig. 3 presents the fine XPS spectra of Co, O, and Al elements in the mentioned nanosheets. Fig. 3a illustrates that the Co 2p spectrum can be deconvoluted into two peaks at 780.78 eV and 797.24 eV, corresponding to Co 2p3/2 and Co 2p1/2, respectively. In addition, the presence of satellite peaks at 787.15 eV and 802.15 eV suggests the presence of a highly spinning divalent state of Co2+ in the sample.56 The binding energies of CoAl (NaOH 4 h) and CoAl (NaOH 10 h) nanosheets are negatively shifted compared to those of pristine CoAl nanosheets, indicating an increase in the electron density caused by alkali etching. The Al 2p spectrum indicates the presence of Al3+ with a peak at 74.09 eV binding energy (Fig. 3b).57 Moreover, the peak signals of CoAl (NaOH 4 h) samples display the highest binding energies, signifying the presence of numerous unsaturated Al sites generated as a result of alkali etching treatment compared to pristine CoAl.36 The appearance of Co 2p and Al 2p peaks confirms the formation of CoAl-LDH. In CoAl (NaOH 4 h), the binding energy of Co 2p is negatively shifted while that of Al 2p is positively shifted in comparison to pristine CoAl. The electron density of Co 2p increases while that of Al 2p decreases, indicating a significant transfer of electrons from Co to the Al cation hole. Fig. 3c shows that the C–C peak is positioned at 284.8 eV based on the bonding between sp2 hybridized carbon atoms, and using it as a reference peak, the peak with a bonding energy of 289.1 eV corresponds to the oxygen-containing functional group CO.48 The VO level can be determined from the O 1s nuclear energy spectrum obtained through XPS.58 The O 1s spectrum can be deconvoluted into three peaks, where the peak at 530.12 eV is from oxygen atoms bound to the metal (M–OH), the peak at 531.36 eV is attributed to O atoms near the VO, and the higher energy peak at 532.54 eV is chemisorbed or dissociated oxygen species (Fig. 3d).59,60 Consistent with the FTIR results, the VO peak of CoAl (NaOH 4 h) shifted to a higher binding energy compared to CoAl samples, which promoted the formation of new neighboring VO with high electron attraction effects.61 Meanwhile, the M–OH peak of the CoAl (NaOH 4 h) sample shows the highest binding energy, indicating that the appropriate alkaline etching time is more favorable for the formation of –OH species. Based on the above results, it can be inferred that alkali etching creates metal cation defect sites by altering the Co and Al contents in CoAl-LDH nanosheets, which promotes the formation of nearby VO.
|
| Fig. 3 Elemental valence characterization of CoAl-LDH nanosheets. High-resolution XPS spectrum of (a) Co 2p, (b) Al 2p, (c) C 1s and (d) O 1s. | |
It is interesting to investigate the alterations in the chemical state and coordination surrounding Co at an atomic level during alkali etching, utilizing X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS). In the normalized XANES spectra, CoAl-LDH exhibits a characteristic Co K-edge spectrum with the pre-edge peak corresponding to the 1s → 3d transition and the white line peak resulting from the 1s → 4p transition (Fig. 4a).62–64 The distinct absorption edge positions of the Co K-edge XANES spectra imply that the Co valence of CoAl-LDH nanosheets increases with the duration of alkali etching, and the oxidation state is between Co2+ and Co3+ (Fig. 4a and b, Table S2†). Alkali etching resulted in a significant change in the atomic valence state of Co in CoAl-LDH, consistent with the XPS results (Fig. 3a). The k3-weighted EXAFS spectra of the Fourier transform (FT) reveal notable dissimilarities in the local atomic structures around the Co atoms among the diverse specimens (Fig. 4c, Table S3†). Within the CoAl-LDH samples, the Co K-edge EXAFS spectra display two distinctive shells in the R-space spectrum, corresponding to the Co–O and Co–Co/Co–Al coordination shells. The Co–O peak, which underwent alkali etching treatment, exhibits a considerable reduction in intensity compared to the pristine CoAl nanosheet. The wavelet transform (WT) indicates significant WT peaks at about 6.0 and 10.0 Å associated with Co–O and Co–Co/Co–Al scattering, respectively (Fig. 4d). There is a continuous weakening of the Co–Al bond after alkali etching, which corresponds with the ICP-OES results (Fig. S1†). The first shell Co–O results in q-space were accurately fitted (Fig. 4e, Table S3†). The alkali etched CoAl (NaOH X h) exhibits a substantially lower Co–O coordination number N and bond length R than pristine CoAl, indicating the presence of hydroxyl vacancies (VOH) in the layer after alkali etching (Fig. 4f, Table S3†).65 The bond lengths of the Co–O in the samples showed no significant changes despite increasing the alkali etching time, while the coordination number experienced a slight decrease. The amount of disorder in the Debye–Waller factor σ2 follows the order CoAl (NaOH 2 h) > CoAl (NaOH 6 h) > CoAl (NaOH 4 h) > CoAl > CoAl (NaOH 8 h) > CoAl (NaOH 10 h). The abundance of VOH theoretically could supply further coordination of unsaturated sites for the NORR. Nevertheless, the low disorder amount of the Debye–Waller factor in CoAl (NaOH 10 h) may restrict its electrocatalytic activity. These XAFS data provide solid evidence for the existence of coordination unsaturation sites and significant structural distortions of etched CoAl-LDH.66 Combined, these structural alterations would decrease the surface energy of LDH nanosheets and augment the abundance of NH3-synthesizing active sites.
|
| Fig. 4 XAFS measurement results of different CoAl samples. (a) XANES spectra. The standard Co foil and Co2O3 are also presented for better comparison. (b) The dependence of Co valence states on the white line positions. (c) Fourier transform (FT) spectra of Co K-edge EXAFS oscillations k3χ(k). (d) Wavelet transform (WT) contour maps. (e) The first shell (Co–O) fitting results in q-space. (f) Co K-edge EXAFS fitting parameters (coordination number, N; bond length, R and Debye-Waller factor, σ2) for the first-shell around the Co atom. | |
3.3. Performance of electrocatalytic NO reduction to NH3
The electrocatalytic activity of CoAl-LDH nanosheets was investigated in an H-type electrolytic cell separated by a Nafion 117 membrane and coupled to a three-electrode configuration. Before the electrochemical experiments, the feed of 1% NO/Ar underwent strict experimental control to eliminate the interference of other impurity gases. The electrocatalytic performance of CoAl-LDH nanosheets was evaluated using linear scanning voltammetry (LSV). As shown in Fig. 5a and S5,† higher current densities were achieved with a NO-saturated electrolyte compared to an Ar-saturated solution. Additionally, Fig. 5b demonstrates that the reduction current densities of the etched CoAl-LDH were all higher than that of the pristine CoAl-LDH, especially for CoAl (NaOH 4 h), confirming the superior electrochemical performance of the etched CoAl-LDH for the NORR.67 After 20 min of electrolysis, the cathodic electrolyte for the NORR was collected and stained with the indophenol blue indicator to quantify the amount of NH3 produced (Fig. S6†). The results showed that CoAl (NaOH 4 h) had the highest NH3 yield of 0.310 mg cm−2 h−1 at −1.1 V vs. RHE, with the corresponding highest FE (Fig. 5c and d). The NH3 yield was 2.2 times increased compared to the pristine CoAl (0.097 mg cm−2 h−1). Meanwhile, Ar was passed to the cathode electrolyte for comparison purposes. It was observed that only a small amount of NH3 was produced in the product, reaching 0.043 mg cm−2 h−1 at −1.1 V vs. RHE (Fig. 5d). This amount had a negligible effect on the experimental results. Additional constant potential electrolysis tests were performed at −1.0 V vs. RHE, and the obtained time-dependent current density curves indicated that the CoAl-LDH nanosheets exhibited excellent stability for 24 h (Fig. 5e). The results of the electrocatalytic activity of CoAl-LDH nanosheets provide compelling evidence that metal cation vacancies generated by alkali etching play a crucial role in enhancing the activity and selectivity of electrocatalytic NORR synthesis of NH3.68,69 It is proposed that the vacancies on the etched CoAl-LDH surface function as active sites, thereby facilitating the adsorption and activation of NO molecules. It is anticipated that this enhancement will elevate the current density, which will in turn result in an increased NH3 yield. The observed increase in both current density and NH3 yield in the etched CoAl-LDH is indicative of its enhanced electrochemical performance, thereby validating the hypothesis that these vacancies significantly contribute to the electrocatalytic activity.
|
| Fig. 5 Electrochemical NORR performance of CoAl-LDH nanosheets. (a) The LSV curves of CoAl (NaOH 4 h) in Ar and NO saturated electrolytes. (b) The LSV curves of CoAl-LDH nanosheets. (c) NH3 yield by electrocatalytic reduction of CoAl-LDH nanosheets. (d) NH3 yield and FE of in CoAl (NaOH 4 h) during aeration of Ar and NO. (e) Long-term stability test at −1.0 V vs. RHE. of CoAl-LDH nanosheets. (f) Tafel slope plots of CoAl-LDH nanosheets. | |
To further evaluate the surface properties, and electrochemical kinetics and to determine the reason for the high NORR activity of CoAl (NaOH 4 h), electrochemical active surface area (ECSA) and Tafel tests were performed. ECSA was determined by calculating the double-layer capacitance (Cdl) using CV data obtained within the non-faradaic region. Fig. S7† displays the CV curves of CoAl-LDH nanosheets at different sweep speeds ranging from 20–120 mV s−1 and the Cdl obtained from the fitting. The Cdl values for CoAl, CoAl (NaOH 4 h), and CoAl (NaOH 10 h) are 0.55, 0.41, and 0.51 mF cm−2, respectively, with no significant difference. The Tafel test results indicate that CoAl (NaOH 4 h) exhibits a significantly lower Tafel slope (307.8 mV dec−1) compared to CoAl (337.7 mV dec−1) and CoAl (NaOH 10 h) (503.6 mV dec−1) (Fig. 5f). This suggests that CoAl (NaOH 4 h) exhibits faster electrochemical reaction kinetics, resulting in a lower kinetic energy barrier to drive the NORR electrocatalytic reaction.70
3.4. Exploration of the origin of the high-performance electrocatalyst
Metal cation vacancy defects were successfully created through the exploitation of the alkali solubility of Al, promoting the formation of VO and enhancing the performance of the electrochemical NORR for NH3 synthesis, as theoretically calculated. Both pristine CoAl and CoAl (NaOH 4 h) exhibit similar lamellar structures. In contrast, CoAl (NaOH 10 h) exhibits flower clusters and shows a greater surface area in N2 adsorption–desorption measurements, which could potentially provide more active sites for reactants. However, it is noteworthy that the catalysts in the electrochemical performance tests exhibited similar ECSA. Nevertheless, CoAl (NaOH 4 h) was able to produce more NH3, suggesting that the catalyst's performance enhancement is not solely due to morphology and surface area effects, but also due to its ability to activate the active sites of NO molecules massively. The lower energy barrier may be attributed to the fact that FLPs promote the adsorption and activation of NO molecules.34 The structural characterization of the catalysts reveals unsaturated metal sites and neighboring –OH groups, which offer more opportunities for the formation of FLPs.
To further confirm the presence of acid–base sites of FLPs in CoAl-LDH nanosheets, NH3/CO2-TPD tests were conducted (Fig. S8†). The presence of two distinct peaks of NH3 (around 320 °C and 419 °C) and CO2 (around 324 °C and 432 °C) detachment indicate the occurrence of medium-strength and strong acidic and basic centers in the catalyst. The strong acidic and basic centers present in CoAl (NaOH 4 h) and CoAl (NaOH 10 h) disappeared following alkali etching. Moderately acidic and basic sites of medium-strength were detected in CoAl (NaOH 4 h) samples, and the optimal amount of LA and LB sites was more conducive to the generation of steric hindrance to promote the formation of FLPs.36
DFT calculations further reveal the origin of the efficient NORR performance of the cation-defect-rich alkali-etched LDH nanosheet (D-LDH) electrocatalysts. Fig. 6a shows the Gibbs free energy distribution plots of NORR processes of LDH and D-LDH catalysts. The first protonation process of adsorbed NO molecules (*NO) is identified as the potential-determining step (PDS) of the NORR for both LDH and D-LDH catalysts (*NO + H+ + e− → *NO). The corresponding Gibbs free energy change values of PDS are 0.71 eV (LDHs) and 0.32 eV (D-LDHs), as calculated. The previously mentioned DFT calculations indicate that D-LDH catalysts have lower PDS values. Better NORR electrocatalytic activity corresponds to lower PDS values, which is concordant with the results of catalyst characterization and electrochemical NORR performance tests. The LDH surface containing metal cation vacancies is more favorable for the formation of VO, and the abundant LA (VO) and LB (–OH) sites provide empty orbitals and protons respectively for NO molecules, lowering the reaction energy barriers and facilitating the synthesis of NH3.71
|
| Fig. 6 (a) NORR reaction energy barriers on LDH and D-LDH surfaces. (b) Mechanistic pathway diagram for NH3 electrosynthesis at FLP sites. | |
Based on experimental results and theoretical calculations, FLPs can enhance the NH3 synthesis performance of the electrochemical NORR in the following ways: alkali etching creates metal cation vacancies, promoting the formation of VO on the surface. Metal cation vacancies (or VO) and neighboring –OH form FLPs due to spatial site–barrier interaction. LA sites can be used as empty orbitals to adsorb activated NO molecules, and LB sites provide protons to generate the *NOH intermediates required for the reaction, facilitating the formation of NH3. It is important to note that the FLP sites are regenerated when the product NH3 is released after synthesis (Fig. 6b).
4. Conclusions
In summary, this work presents a novel strategy for targeted structural optimization of catalysts to boost the performance of the electrocatalytic NO synthesis of NH3. CoAl-LDH nanosheets containing cationic vacancies were fabricated utilizing a straightforward alkali etching technique. After 4 h of alkali etching, the NH3 yield of CoAl-LDH reached 0.310 mg cm−2 h−1 at −1.1 V vs. RHE, which is a 2.2-fold improvement compared with that of the pristine CoAl-LDH (0.097 mg cm−2 h−1). Experimental and theoretical results indicate that the VO and neighboring –OH species in the artificial FLPs formed by LDH rich in cationic vacancies act synergistically to promote NO adsorption and activation, opening a new avenue for the use of LDH electrocatalysts to construct FLPs for electrocatalytic NORR synthesis of NH3.
Data availability
Data will be made available on request.
Author contributions
Honghong Yi: conceptualization; formal analysis; funding acquisition; methodology; project administration; resources; writing – original draft. Ruzhu Jia: conceptualization; data curation; formal analysis; validation; writing – original draft. Xiaolong Tang: conceptualization; data curation; resources; supervision. Dongjuan Kang: methodology; resources; supervision. Qingjun Yu: investigation; resources; supervision. Fengyu Gao: conceptualization; methodology; resources; supervision. Shunzheng Zhao: conceptualization; formal analysis; funding acquisition; methodology; project administration; supervision; writing – review & editing. Yunpeng Liu: data curation; funding acquisition; project administration; supervision; visualization; writing – review & editing.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the National Key R&D Program of China (2021YFB3500702), National Natural Science Foundation of China (Project No. 12305372, 21677010, and 51808037), Fundamental Research Funds for the Central Universities (No. FRF-IDRY-20-018), and Special fund of Beijing Key Laboratory of Indoor Air Quality Evaluation and Control (No. BZ0344KF21-04). We also thank the 4B9A beamline (Beijing Synchrotron Radiation Facility, BSRF) for supporting the beam time of XAFS measurements.
References
- H. Li, L. Schill, R. Fehrmann and A. Riisager, Inorg. Chem. Front., 2023, 10, 727–755 RSC .
- Q. Liu, Y. Lin, L. Yue, J. Liang, L. Zhang, T. Li, Y. Luo, M. Liu, J. You, A. A. Alshehri, Q. Kong and X. Sun, Nano Res., 2022, 15, 5032–5037 CrossRef CAS .
- M. Fu, C. Li, P. Lu, L. Qu, M. Zhang, Y. Zhou, M. Yu and Y. Fang, Catal. Sci. Technol., 2014, 4, 14–25 RSC .
- X. Peng, Y. Mi, H. Bao, Y. Liu, D. Qi, Y. Qiu, L. Zhuo, S. Zhao, J. Sun, X. Tang, J. Luo and X. Liu, Nano Energy, 2020, 78, 105321 CrossRef CAS .
- W. Zhang, X. Qin, T. Wei, Q. Liu, J. Luo and X. Liu, J. Colloid Interface Sci., 2023, 638, 650–657 CrossRef CAS PubMed .
- J. Cheng, W. Li, X. Wang, H. Liu and Y. Chen, Chem. Eng. J., 2022, 431, 133694 CrossRef CAS .
- H. Niu, Z. Zhang, X. Wang, X. Wan, C. Kuai and Y. Guo, Small, 2021, 17, 2102396 CrossRef CAS PubMed .
- K. H. R. Rouwenhorst, A. G. J. Van der Ham and L. Lefferts, Int. J. Hydrogen Energy, 2021, 46, 21566–21579 CrossRef CAS .
- M. Yang, T. Wei, J. He, Q. Liu, L. Feng, H. Li, J. Luo and X. Liu, Nano Res., 2023, 17, 1209–1216 CrossRef .
- G. Hochman, A. S. Goldman, F. A. Felder, J. M. Mayer, A. J. M. Miller, P. L. Holland, L. A. Goldman, P. Manocha, Z. Song and S. Aleti, ACS Sustainable Chem. Eng., 2020, 8, 8938–8948 CrossRef CAS .
- J. Zheng, L. Jiang, Y. Lyu, S. P. Jiang and S. Wang, Energy Environ. Mater., 2022, 5, 452–457 CrossRef CAS .
- H. Yin, Z. Chen, S. Xiong, J. Chen, C. Wang, R. Wang, Y. Kuwahara, J. Luo, H. Yamashita, Y. Peng and J. Li, Chem Catal., 2021, 1, 1088–1103 CrossRef CAS .
- K. Liu, J. Fu, L. Zhu, X. Zhang, H. Li, H. Liu, J. Hu and M. Liu, Nanoscale, 2020, 12, 4903–4908 RSC .
- K. Chen, G. Zhang, X. Li, X. Zhao and K. Chu, Nano Res., 2023, 16, 5857–5863 CrossRef CAS .
- K. Chen, D. Ma, Y. Zhang, F. Wang, X. Yang, X. Wang, H. Zhang, X. Liu, R. Bao and K. Chu, Adv. Mater., 2024, 2402160 CrossRef CAS PubMed .
- K. Chen, J. Xiang, Y. Guo, X. Liu, X. Li and K. Chu, Nano Lett., 2024, 24, 541–548 CrossRef CAS PubMed .
- S. Chen, G. Qi, R. Yin, Q. Liu, L. Feng, X. Feng, G. Hu, J. Luo, X. Liu and W. Liu, Nanoscale, 2023, 15, 19577–19585 RSC .
- M. Yang, T. Wei, J. He, Q. Liu, L. Feng, H. Li, J. Luo and X. Liu, Nano Res., 2024, 17, 1209–1216 CrossRef CAS .
- Z.-X. Ge, T.-J. Wang, Y. Ding, S.-B. Yin, F.-M. Li, P. Chen and Y. Chen, Adv. Energy Mater., 2022, 12, 2103916 CrossRef CAS .
- W. Zhong, Q.-L. Hong, X. Ai, C. Zhang, F.-M. Li, X.-F. Li and Y. Chen, Adv. Mater., 2024, 36, 2314351 CrossRef CAS PubMed .
- X. Zhu, X. Fan, H. Lin, S. Li, Q. Zhai, Y. Jiang and Y. Chen, Adv. Energy Mater., 2023, 13, 2300669 CrossRef CAS .
- R. Zhao, Z. Ma, Y. Yu, X. Xia, B. Song, T. Zhou and Y. Huang, Nano Res., 2023, 16, 8533–8541 CrossRef CAS .
- X. Wang, Y. Cai, T. Han, M. Fang, K. Chen and X. Tan, J. Hazard. Mater., 2020, 399, 123081 CrossRef CAS PubMed .
- Z.-Q. Liu, X. Liang, F.-X. Ma, Y.-X. Xiong, G. Zhang, G. Chen, L. Zhen and C.-Y. Xu, Adv. Energy Mater., 2023, 13, 2203609 CrossRef CAS .
- J. Lin, H. Jia, H. Liang, S. Chen, Y. Cai, J. Qi, C. Qu, J. Cao, W. Fei and J. Feng, Chem. Eng. J., 2018, 336, 562–569 CrossRef CAS .
- Y. Yang, L. Dang, M. J. Shearer, H. Sheng, W. Li, J. Chen, P. Xiao, Y. Zhang, R. J. Hamers and S. Jin, Adv. Energy Mater., 2018, 8, 1703189 CrossRef .
- Y.-F. Qi, K.-Y. Wang, Y. Sun, J. Wang and C. Wang, ACS Sustainable Chem. Eng., 2022, 10, 645–654 CrossRef CAS .
- P. Gao, Z. Chen, Y. Gong, R. Zhang, H. Liu, P. Tang, X. Chen, S. Passerini and J. Liu, Adv. Energy Mater., 2020, 10, 1903780 CrossRef CAS .
- P. Liang, D. Meng, Y. Liang, Z. Wang, C. Zhang, S. Wang and Z. Zhang, Chem. Eng. J., 2021, 409, 128196 CrossRef CAS .
- F. Bai, X. Qu, C. Li, S. Liu, J. Sun, X. Chen and W. Yang, ACS Appl. Mater. Interfaces, 2022, 14, 28033–28043 CrossRef CAS PubMed .
- E. J. Lawrence, V. S. Oganesyan, D. L. Hughes, A. E. Ashley and G. G. Wildgoose, J. Am. Chem. Soc., 2014, 136, 6031–6036 CrossRef CAS PubMed .
- J. Sheng, Y. He, M. Huang, C. Yuan, S. Wang and F. Dong, ACS Catal., 2022, 12, 2915–2926 CrossRef CAS .
- S. Zhang, Z. Xia, Y. Zou, F. Cao, Y. Liu, Y. Ma and Y. Qu, J. Am. Chem. Soc., 2019, 141, 11353–11357 CrossRef CAS PubMed .
- Y. Yang, Y.-X. Pan, X. Tu and C.-j. Liu, Nano Energy, 2022, 101, 107613 CrossRef CAS .
- Z. Chen, J. Zhao, Y. Jiao, T. Wang and L. Yin, J. Energy Chem., 2022, 66, 628–634 CrossRef CAS .
- M. Yuan, H. Zhang, Y. Xu, R. Liu, R. Wang, T. Zhao, J. Zhang, Z. Liu, H. He, C. Yang, S. Zhang and G. Zhang, Chem Catal., 2022, 2, 309–320 CrossRef CAS .
- C. Jia, X. Kan, X. Zhang, G. Lin, W. Liu, Z. Wang, S. Zhu, D. Ju and J. Liu, Chem. Eng. J., 2022, 427, 131554 CrossRef CAS .
- M. Ma, L. Li, G. Tian, Z. Geng, X. Zhang, X. Zhao and G. Li, Chem. Eng. J., 2023, 470, 144291 CrossRef CAS .
- Q. Wang, Z. Miao, Y. Zhang, T. Yan, L. Meng and X. Wang, ACS Catal., 2022, 12, 4016–4025 CrossRef CAS .
- W. Uhl and E.-U. Würthwein, in Frustrated Lewis Pairs II: Expanding the Scope, ed. G. Erker and D. W. Stephan, Springer, Berlin, Heidelberg, Topics in Current Chemistry, 2012, vol. 334, pp. 101–119 Search PubMed .
- Y.-i. Kwon, S. K. Kim, Y. B. Kim, S. J. Son, G. D. Nam, H. J. Park, W.-C. Cho, H. C. Yoon and J. H. Joo, ACS Energy Lett., 2021, 6, 4165–4172 CrossRef CAS .
- Y. Zhao, L. Zheng, R. Shi, S. Zhang, X. Bian, F. Wu, X. Cao, G. I. N. Waterhouse and T. Zhang, Adv. Energy Mater., 2020, 10, 2002199 CrossRef CAS .
- D. Yan, C. Xia, W. Zhang, Q. Hu, C. He, B. Y. Xia and S. Wang, Adv. Energy Mater., 2022, 12, 2202317 CrossRef CAS .
- Z. Li, X. Wang, M. Zeng, K. Chen, D. Cao, Y. Huang, Y. Zhu, W. Zhang, N. Wang and Y. A. Wu, J. Colloid Interface Sci., 2022, 625, 363–372 CrossRef CAS PubMed .
- J. Di, C. Chen, C. Zhu, P. Song, J. Xiong, M. Ji, J. Zhou, Q. Fu, M. Xu, W. Hao, J. Xia, S. Li, H. Li and Z. Liu, ACS Appl. Mater. Interfaces, 2019, 11, 30786–30792 CrossRef CAS PubMed .
- Q. Li, H. Yi, X. Tang, S. Zhao, B. Zhao, D. Liu and F. Gao, Chem. Eng. J., 2016, 284, 103–111 CrossRef CAS .
- Y.-J. Wang, J.-Y. Zhang, S.-S. Hou, J.-X. Wu, C. Wang, Y.-M. Li, G.-Y. Jiang and G.-Q. Cui, Pet. Sci., 2022, 19, 3080–3087 CrossRef CAS .
- Y. Zhang, S.-c. Xue, X.-h. Yan, H.-L. Gao and K.-z. Gao, Electrochim. Acta, 2023, 442, 141822 CrossRef CAS .
- Y. Wang, L. Dou and H. Zhang, ACS Appl. Mater. Interfaces, 2017, 9, 38784–38795 CrossRef CAS PubMed .
- L. Dou and H. Zhang, J. Mater. Chem. A, 2016, 4, 18990–19002 RSC .
- J. Fang, M. Li, Q. Li, W. Zhang, Q. Shou, F. Liu, X. Zhang and J. Cheng, Electrochim. Acta, 2012, 85, 248–255 CrossRef CAS .
- L. Wu, X.-X. Ding, X.-F. Zhao, Y.-H. Liu, X.-L. Hao, A.-T. Tang, G. Zhang and F.-S. Pan, Rare Met., 2023, 42, 697–704 CrossRef .
- R.-r. Shan, L.-g. Yan, K. Yang, S.-j. Yu, Y.-f. Hao, H.-q. Yu and B. Du, Chem. Eng. J., 2014, 252, 38–46 CrossRef CAS .
- Z. Tian, Q. Li, J. Hou, L. Pei, Y. Li and S. Ai, J. Catal., 2015, 331, 193–202 CrossRef CAS .
- C. Jing, Y. Huang, L. Xia, Y. Chen, X. Wang, X. Liu, B. Dong, F. Dong, S. Li and Y. Zhang, Appl. Surf. Sci., 2019, 496, 143700 CrossRef CAS .
- Y. Dou, S. Zhang, T. Pan, S. Xu, A. Zhou, M. Pu, H. Yan, J. Han, M. Wei, D. G. Evans and X. Duan, Adv. Funct. Mater., 2015, 25, 2243–2249 CrossRef CAS .
- P. Yue, H. She, L. Zhang, B. Niu, R. Lian, J. Huang, L. Wang and Q. Wang, Appl. Catal., B, 2021, 286, 119875 CrossRef CAS .
- J. Di, J. Xia, M. F. Chisholm, J. Zhong, C. Chen, X. Cao, F. Dong, Z. Chi, H. Chen, Y.-X. Weng, J. Xiong, S.-Z. Yang, H. Li, Z. Liu and S. Dai, Adv. Mater., 2019, 31, 1807576 CrossRef PubMed .
- Y. Zhong, C. Wu, X. Jia, S. Sun, D. Chen, W. Yao, H. Ding, J. Zhang and T. Ma, Chem. Eng. J., 2023, 465, 142893 CrossRef CAS .
- M. Xing, J. Zhang, F. Chen and B. Tian, Chem. Commun., 2011, 47, 4947–4949 RSC .
- Y. Zhong, C. Wu, Y. Feng, D. Chen, Y. Wang, D. Hao and H. Ding, Appl. Surf. Sci., 2022, 585, 152656 CrossRef CAS .
- N. H. M. D. Dostagir, R. Rattanawan, M. Gao, J. Ota, J.-y. Hasegawa, K. Asakura, A. Fukouka and A. Shrotri, ACS Catal., 2021, 11, 9450–9461 CrossRef CAS .
- W. Yang, X. Liu, L. Chen, L. Liang and J. Jia, Chem. Commun., 2017, 53, 4034–4037 RSC .
- M. Shen, L.-R. Zheng, W. He, C. Ruan, C. Jiang, K. Ai and L. Lu, Nano Energy, 2015, 17, 120–130 CrossRef CAS .
- C. Ning, Z. Wang, S. Bai, L. Tan, H. Dong, Y. Xu, X. Hao, T. Shen, J. Zhao, P. Zhao, Z. Li, Y. Zhao and Y.-F. Song, Chem. Eng. J., 2021, 412, 128362 CrossRef CAS .
- R. Zhang, Y.-C. Zhang, L. Pan, G.-Q. Shen, N. Mahmood, Y.-H. Ma, Y. Shi, W. Jia, L. Wang, X. Zhang, W. Xu and J.-J. Zou, ACS Catal., 2018, 8, 3803–3811 CrossRef CAS .
- Y. Xu, Y. Wen, T. Ren, H. Yu, K. Deng, Z. Wang, X. Li, L. Wang and H. Wang, Appl. Catal., B, 2023, 320, 121981 CrossRef CAS .
- B. Zhang, S. Qiu, Y. Xing, G. Zhao, W. Liao, L. Mu and N. Zhao, Langmuir, 2023, 39, 18152–18160 CrossRef CAS PubMed .
- Y.-j. Wu, J. Yang, T.-x. Tu, W.-q. Li, P.-f. Zhang, Y. Zhou, J.-f. Li, J.-t. Li and S.-G. Sun, Angew. Chem., Int. Ed., 2021, 60, 26829–26836 CrossRef CAS PubMed .
- X. Min and B. Liu, Small, 2023, 19, 2300794 CrossRef CAS PubMed .
- M. Yuan, J. Chen, Y. Xu, R. Liu, T. Zhao, J. Zhang, Z. Ren, Z. Liu, C. Streb, H. He, C. Yang, S. Zhang and G. Zhang, Energy Environ. Sci., 2021, 14, 6605–6615 RSC .
|
This journal is © The Royal Society of Chemistry 2024 |