Single-Crystalline Mesoporous Palladium and Palladium-Copper Nanocubes for Highly Efficient Electrochemical CO 2 Reduction

电化学 介孔材料 材料科学 还原(数学) 纳米技术 无机化学 冶金 化学 催化作用 电极 物理化学 有机化学 几何学 数学
作者
Hao Lv,Fang Lv,Huaiyu Qin,Xiaowen Min,Lizhi Sun,Ning Han,Dongdong Xu,Yanguang Li,Ben Liu
出处
期刊:CCS Chemistry [Chinese Chemical Society]
卷期号:4 (4): 1376-1385 被引量:32
标识
DOI:10.31635/ccschem.021.202100958
摘要

Open AccessCCS ChemistryRESEARCH ARTICLE1 Apr 2022Single-Crystalline Mesoporous Palladium and Palladium-Copper Nanocubes for Highly Efficient Electrochemical CO2 Reduction Hao Lv†, Fang Lv†, Huaiyu Qin, Xiaowen Min, Lizhi Sun, Na Han, Dongdong Xu, Yanguang Li and Ben Liu Hao Lv† College of Chemistry, Sichuan University, Chengdu 610064 †H. Lv and F. Lv contributed equally to this work.Google Scholar More articles by this author , Fang Lv† Institute of Functional Nano and Soft Materials (FUNSOM), Jiangsu Key Laboratory for Carbon-Based Functional Materials and Devices, Soochow University, Suzhou 215123 †H. Lv and F. Lv contributed equally to this work.Google Scholar More articles by this author , Huaiyu Qin College of Chemistry, Sichuan University, Chengdu 610064 Google Scholar More articles by this author , Xiaowen Min College of Chemistry, Sichuan University, Chengdu 610064 Google Scholar More articles by this author , Lizhi Sun Jiangsu Key Laboratory of New Power Batteries, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023 Google Scholar More articles by this author , Na Han Institute of Functional Nano and Soft Materials (FUNSOM), Jiangsu Key Laboratory for Carbon-Based Functional Materials and Devices, Soochow University, Suzhou 215123 Google Scholar More articles by this author , Dongdong Xu Jiangsu Key Laboratory of New Power Batteries, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023 Google Scholar More articles by this author , Yanguang Li *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] Institute of Functional Nano and Soft Materials (FUNSOM), Jiangsu Key Laboratory for Carbon-Based Functional Materials and Devices, Soochow University, Suzhou 215123 Google Scholar More articles by this author and Ben Liu *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected]da.edu.cn College of Chemistry, Sichuan University, Chengdu 610064 Jiangsu Key Laboratory of New Power Batteries, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023 Google Scholar More articles by this author https://doi.org/10.31635/ccschem.021.202100958 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail Mesoporous single crystals have unique potential in catalysis, but remain unexplored owing to the enormous synthetic challenge that they pose. Herein, we report a facile soft-template method to prepare palladium (Pd) and Pd alloy nanocubes with single-crystallinity and abundant mesoporosity. The successful formation of these exotic nanostructures essentially relies on the cointroduction of cetyltrimethylammonium chloride as the surfactant template and extra Cl− ions as the facet-selective capping agent under well controlled experimental conditions. Thanks to their large surface areas and penetrating mesoporous channels, our products exhibit a great performance for electrochemical CO2 reduction. The best sample from alloying palladium with copper enables the efficient formate production with high selectivity (90∼100%) over a broad potential range, and great stability even under the working potential as cathodic as −0.5 V versus a reversible hydrogen electrode. These performance metrics are far superior to previous Pd-based materials, and underscore the structural advantages of our products. Download figure Download PowerPoint Introduction Single-crystalline mesoporous metals constitute a unique class of materials for catalytic applications. Their ordered mesopores expose large accessible surface areas and active sites, and facilitate the rapid diffusion of reactants and products,1–6 while their single-crystallinity ensures long-range structural coherence and enhances electron transport.7–10 Unfortunately, the combination of ordered mesoporosity and single-crystallinity appears inherently unfavorable. Complete crystallization of mesoporous walls significantly compromises structural stability and integrity, leading to the distortion or even collapse of ordered mesopores and consequently the loss of accessible surface areas and active sites.11–14 Indeed, most mesoporous materials (such as silica, carbon, and metal oxides) are amorphous or poorly crystalline in nature. There are only scarce reports available about single-crystalline mesoporous oxides or nitrides in the literature.14–23 The preparation of single-crystalline mesoporous metals is even more challenging since metals generally have larger surface energy than oxides, and have a greater propensity to minimize their surface areas and form disordered structures.20,24,25 This synthetic obstacle largely precludes their potential application at the current stage. One of the emerging catalytic processes is the electrochemical CO2 reduction reaction (CO2RR). It converts CO2 to value-added industrial chemicals or fuels, and is an essential step to close the artificial carbon cycle. Most CO2RR electrocatalysts are composed of metals.26–29 Depending on the selection of catalysts and experimental conditions, different reduction products can be attained from CO2 reduction, of which formic acid or formate is recommended as the most economically viable.30,31 Pd is so far the only known candidate capable of selectively reducing CO2 to formate at close-to-zero overpotential in aqueous solution.32–35 It unfortunately suffers from severe CO poisoning and is subject to rapid selectivity and stability loss with increasing overpotential.36–38 To this end, efforts have been made through nanostructural engineering and/or compositional regulation to enhance the operational stability of Pd-based materials. The performance gain, however, remains limited. In this contribution, we develop a facile aqueous method to prepare single-crystalline mesoporous Pd (s-mesoPd) and Pd alloy nanocubes in the copresence of a suitable quaternary ammonium surfactant and extra Cl− ions. The product exhibits uniform nanocubic morphology with the {100} enclosure, and features abundant mesoporous channels and single-crystallinity. When investigated for CO2 reduction in an aqueous solution, the best catalyst achieved unprecedented selectivity and stability for the formate production at the cathodic potential that was previously believed to cause quick catalyst deactivation. Experimental Methods Synthesis of s-mesoPd nanocubes and other nanocrystals Synthesis of s-mesoPd nanocubes In a typical synthesis of s-mesoPd nanocubes, 0.325 mg of cetyltrimethylammonium chloride (CTAC) was first dissolved in 5.0 mL of deionized H2O, followed by the addition of 0.24 mL of 80 mM KCl and 0.25 mL of 10 mM H2PdCl4. After incubation at 50 °C for 30 min, 0.50 mL of 0.30 M freshly prepared l-ascorbic acid (AA) was rapidly injected into the above solution with gentle shaking. The color of the solution gradually changed from orange to dark brown. After another 30 min, the product was collected by centrifugation and washed several times with ethanol and H2O. The surfactant type and/or KCl concentration were varied under otherwise identical conditions to investigate their effects on the product morphology and crystallinity. The size of s-mesoPd nanocubes was adjusted by varying the added amount of H2PdCl4 under otherwise identical conditions. Synthesis of p-mesoPd nanoparticles In a typical synthesis of p-mesoPd nanoparticles, 0.325 mg of CTAC was first dissolved in 5.0 mL of deionized H2O, followed by the addition of 0.25 mL of 10 mM H2PdCl4. After incubation at 50 °C for 30 min, 0.50 mL of 0.30 M freshly prepared AA was rapidly injected into the above solution with gentle shaking. After another 30 min, the product was collected by centrifugation and washed several times with ethanol and H2O. Synthesis of s-mesoPdCu nanocubes In a typical synthesis, 0.325 mg of CTAC was first dissolved in 5.0 mL of deionized H2O, followed by the addition of 0.24 mL of 80 mM KCl, 0.25 mL of 10 mM H2PdCl4, and 0.125 mL of 10 mM Cu(NO3)2. After incubation at 50 °C for 30 min, 0.50 mL of 0.30 M freshly prepared AA was rapidly injected into the above solution with gentle shaking. After another 30 min, the product was collected by centrifugation and washed several times with ethanol and H2O. Synthesis of s-mesoPdRh nanocubes In a typical synthesis, 0.325 mg of CTAC was first dissolved in 5.0 mL of deionized H2O, followed by the addition of 0.24 mL of 80 mM KCl, 0.25 mL of 10 mM H2PdCl4, and 0.125 mL of 10 mM (NH4)3RhCl6. After incubation at 50 °C for 30 min, 0.50 mL of 0.30 M freshly prepared AA was rapidly injected into the above solution with gentle shaking. After another 30 min, the product was collected by centrifugation and washed several times with ethanol and H2O. Synthesis of Pd nanoparticles In a typical synthesis, 13.0 mg of cetylpyridinium bromide was dissolved in 5.0 mL of H2O, followed by the addition of 0.25 mL of 10 mM H2PdCl4. After incubation at 50 °C for 30 min, 0.50 mL of 0.30 M freshly prepared AA was rapidly injected into the above solution with gentle shaking. After another 30 min, the product was collected by centrifugation and washed several times with ethanol/H2O. Electrochemical measurements Electrochemical CO2RR was carried out in a gas-tight H-cell controlled by the standard three-electrode system as reported in our previous work.38 To prepare the working electrodes, 1.00 mg of the catalyst powders under investigation and 0.50 mg of Ketjenblack carbon were added to 6.0 μL of 5 wt % Nafion solution and 250 μL of ethanol, and sonicated for >30 min to form a uniform dispersion. This catalyst ink was then dropped onto a 1 × 1 cm2 glassy carbon plate to achieve a catalyst loading of 1 mg cm−2 and dried under ambient conditions. For CO2RR measurements, the working electrode and a saturated calomel reference electrode (SCE) were placed in the cathodic compartment. A graphitic rod counter electrode was placed in the anodic compartment. These two compartments were separated by a Nafion-117 membrane, and each filled with 30 mL 0.10 M KHCO3 electrolyte presaturated with CO2 (pH = 6.8). All the potential readings in our study were measured against SCE and converted with respect to a reversible hydrogen electrode (RHE) with ∼90% iR compensation. Only geometric current densities were reported. Polarization curves were recorded from the cathodic sweeping of the working electrode at the scan rate of 10 mV s−1. Chronoamperometric analysis was carried out at a few selected potential points for the selectivity and stability assessment. At the end of the chronoamperometric study, formate accumulated in the catholyte was analyzed using ion chromatography (Dionex ICS-600; Thermo Scientific, USA) by comparing with the calibration curve from a series of standard formate solutions. Characterizations Scanning electron microscopy (SEM) images were collected using a JSM-7600F field emission scanning electron microscope (JEOL, Japan). SEM samples were prepared by dropcasting a suspension of the sample powder onto a silicon wafer. Transmission electron microscopy (TEM) studies were carried out using a JEM-F200 field emission transmission electron microscope (JEOL, Japan) with an accelerating voltage of 200 kV. TEM samples were prepared by dropcasting a diluted suspension of the sample powder onto a carbon coated copper grid (300 mesh). Scanning TEM (STEM) images were collected on a Talos F200X scanning/transmission electron microscope (Thermo Scientific, USA) operating at an accelerating voltage of 200 kV and equipped with an energy-dispersive X-ray spectroscopy (EDS) detector for elemental mapping analysis. Small-angle X-ray scattering (SAXS) patterns were measured using an Anton Paar SAXSess mc2 instrument (Austria). Powder X-ray diffraction (XRD) patterns were recorded on powder samples using a D/max 2500 VL/PC diffractometer (Japan) equipped with graphite-monochromatized Cu Kα radiation in 2θ ranging from 30° to 90°. The working voltage and current were 40 kV and 100 mA, respectively. X-ray photoelectron spectroscopy (XPS) was performed on a Thermo ESCALAB 250Xi X-ray photoelectron spectrometer (Thermo Scientific, USA) using Al Kα radiation. The binding energy of the C 1s peak (284.8 eV) was employed as a standard to calibrate the binding energies of other elements (Pd and Cu). Results and Discussion s-mesoPd nanocubes were prepared via a facile one-step soft-template method by reducing PdCl42− with AA in the presence of CTAC as the surfactant template and KCl as the facet-selective capping agent (Figure 1a). The cointroduction of CTAC and extra Cl− ions holds the decisive key to the successful formation of s-mesoPd nanocubes. The SAXS pattern of the final product shows a well-defined peak at 1.04 nm−1, suggesting mesoporous structure with an average periodicity of 6.04 nm (Figure 1b). The powder XRD pattern of s-mesoPd displays a set of diffraction peaks in the 2θ range of 30–90° assignable to face-centered cubic (fcc) Pd (JCPDS: 05-0681), evidencing that it is crystalline at the atomic scale (Figure 1c). Under SEM, the product is observed to have a cubic morphology with rounded edges and corners and an average size of ∼100 nm (Figure 1d). Figure 1e depicts an individual nanocube viewed perpendicular to its face, edge, or corner. Careful examination reveals that the nanocube surface is riddled with mesopores, which become more evident under STEM and TEM imaging (Figures 1f and 1g and Supporting Information Figure S1). They extend from the center and form cylindrical mesoporous channels radially penetrating the entire nanocube. The meopore size and framework thickness are measured to be 2.7 and 3.2 nm, respectively, consistent with the average periodicity derived from the above SAXS analysis. Surprisingly, we find that each nanocube is a single crystal despite its abundant mesoporosity. The selected area electron diffraction (SAED) pattern of the nanocube shown in Figure 1h only exhibits a single set of bright spots along the [100] zone direction of fcc Pd (Figure 1i). High-resolution TEM imaging of four randomly selected regions of the nanocube discloses the same lattice orientation and an identical d-spacing of 0.196 nm from the (200) plane (Figure 1j). These results unambiguously confirm that s-mesoPd nanocubes are single-crystalline and enclosed with six {100} facets. Figure 1 | Synthesis and characterizations of s-mesoPd nanocubes. (a) Schematic synthetic procedure. (b) SAXS analysis. (c) XRD pattern. (d and e) SEM images, inset in (d) illustrates the nanocube size distribution. (f) STEM image, inset shows the {100} enclosure of a nanocube. (g) TEM image. (h) TEM image of a nanocube and corresponding (i) SAED pattern and (j) high-resolution TEM images from different regions in (h). Download figure Download PowerPoint Here, the combination of the overall nanocubic morphology, abundant mesoporosity and single-crystallinity is highly unusual, and not available in previous literature to the best of our knowledge. We believe that CTAC and Cl− ions play important roles in precisely regulating the nanocrystal growth. In what follows, a series of control experiments are carried out to elucidate their effects. We first explored the effect of the surfactant template on the product structure and morphology (Figure 2a). In the absence of any surfactant, the product consists of concave nanocubes free of any mesoporous channels ( Supporting Information Figures S2a–S2c). When shorter-chained octyltrimethylammonium chloride (C8TAC) or dodecyltrimethylammonium chloride (C12TAC) are used instead of CTAC under otherwise identical conditions, resultant nanocubes become smaller in size and have less ordered mesoporous channels ( Supporting Information Figures S2d–S2i). When longer-chained behenyltrimethylammonium chloride (C22TAC) is used, two-dimensional (2D) Pd nanosheets are yielded, presumably from templating the lamellar mesophase of C22TAC in solution ( Supporting Information Figures S2j–S2l and S3).39 Figure 2 | Exploration of different experimental parameters. (a) TEM images of products prepared with no surfactant or in the presence of quaternary ammonium surfactants with different chain lengths under otherwise identical conditions. (b) TEM images of products prepared with no extra Cl− or different Cl− concentrations as indicated under otherwise identical conditions. (c) TEM images showing the structural evolution of s-mesoPd nanocubes with the reaction time. Download figure Download PowerPoint We then investigated the effect of Cl− ions during the synthesis of s-mesoPd nanocubes (Figure 2b). In the absence of extra Cl− ions from KCl, the product consists of spherical polycrystalline mesoporous Pd (p-mesoPd) nanoparticles ( Supporting Information Figures S4a–S4c). With the extra addition of Cl− ions, the overall product morphology gradually evolves from spheres to cubes. The optimal Cl− concentration is found to be 4.0 mM, giving rise to s-mesoPd nanocubes with the {100} enclosure as characterized in detail above. EDS and XPS analyses reveal the presence of Cl residue on the product surface after the synthesis ( Supporting Information Figure S5). Adding too many Cl− ions (16 mM) affords smaller polycrystalline nanocubes albeit with the same {100} enclosure ( Supporting Information Figures S4d–S4f). It should be noted that K+ cations do not play a role in the facet selection as replacing KCl with an equimolar amount of NaCl, CaCl2 or NH4Cl results in similar s-mesoPd nanocubes ( Supporting Information Figure S6). However, replacing Cl− ions with Br− ions or adding OH− compromises the nanocubic morphology ( Supporting Information Figure S7). Based on the above experimental observations, we propose that s-mesoPd nanocubes are formed through a concurrent surfactant-templating and facet-selective growth mechanism. CTAC is directly responsible for the formation of abundant mesoporous channels. During the reaction, amphiphilic CTAC molecules interact with PdCl42− through the Coulombic attraction, self-assemble to one-dimensional (1D) cylindrical micelles that further organize into a three-dimensional (3D) mesophase. The reduction of PdCl42− by AA gives rise to the templated growth of metallic Pd nanocrystals on the surfactant micelles, creating ordered mesoporous channels upon the surfactant removal. In the meantime, Cl− ions strongly bind to Pd(100) and selectively stabilize this facet.40,41 Such a facet-selective capping effect ultimately results in overall nanocubic morphology of our product. It is the copresence of CTAC and extra Cl− ions that collectively makes possible the successful growth of our s-mesoPd nanocubes in solution. To lend further support to the proposed mechanism, we track the time-dependent nanocube growth under TEM (Figure 2c). Upon the injection of AA, single-crystalline dendritic nanoparticles of 10–25 nm are formed immediately ( Supporting Information Figures S8a–S8c). They grow bigger and gradually evolve into nanocubes in the next 60 s ( Supporting Information Figures S8d–S8i). The nanocubic morphology is maintained throughout the rest of the reaction. Finally, s-mesoPd nanocubes were obtained in 10 min. Our solution synthetic strategy is versatile. For example, by increasing the amount of the PdCl42− precursor under otherwise identical conditions, the size of s-mesoPd nanocubes can be systematically varied from 31 to 132 nm, all of which are single-crystalline and mesoporous (Figures 3a–3d and Supporting Information Figure S9). More interestingly, this method can be extended to the preparation of single-crystalline mesoporous nanocubes of bimetallic alloys (s-mesoPdM nanocubes) such as s-mesoPdCu (Figures 3e–3i) and s-mesoPdRh ( Supporting Information Figure S10). Taking s-mesoPdCu nanocubes (with a nominal Pd/Cu molar ratio of 2) as an example, we find that the coreduction of PdCl42− and Cu2+ does not noticeably alter the templating effect of CTAC and facet-selective capping effect of Cl− ions. Resultant nanocubes are observed to have a similar morphology as that of s-mesoPd nanocubes, while the EDS elemental mapping of Pd and Cu clearly evidences their uniform spatial distribution and attests to the alloy formation instead of phase segregation. Their SAXS and XRD measurements also support the formation of the bimetallic PdCu alloy with ordered mesopores ( Supporting Information Figure S11). XPS analysis evidences the electron transfer from Cu to Pd in the alloy ( Supporting Information Figure S12). The compositional tunability in s-mesoPdM nanocubes provides an additional opportunity to tailor their electronic structures and thereby their catalytic properties. Figure 3 | Size and composition control of s-mesoPd and s-mesoPdM nanocubes. (a–d) TEM images of s-mesoPd nanocubes prepared with different CTAC concentrations. (e) STEM image of s-mesoPdCu nanocubes. (f) TEM image of a s-mesoPdCu nanocube and corresponding (g) SAED pattern and (h) high-resolution TEM image. (i) STEM image of a s-mesoPdCu nanocube and corresponding elemental mapping. Download figure Download PowerPoint The unique combination of mesoporosity and single-crystallinity renders our materials promising candidates for many applications. As a proof of concept, we here investigated the electrocatalytic performances of s-mesoPd and s-mesoPdCu nanocubes for CO2RR to formate, and compared them with p-mesoPd nanoparticles and regular Pd nanoparticles (Pd NPs). Pd is the only known material that can enable the selective CO2 reduction to formate at close-to-zero overpotential, but unfortunately suffers from very limited selectivity and stability under increasing overpotential (η > 200 mV) due to CO poisoning.34,36,38 We reason that this longstanding challenge might be alleviated by taking advantage of the large electrochemical surface areas and rich undercoordinated sites of single-crystalline mesoporous nanocubes as well as the electronic effect from alloying Pd and Cu. Figure 4a depicts the polarization curves of s-mesoPdCu nanocubes in N2- or CO2-saturated 0.1 M KHCO3. These two curves bifurcate at ∼0 V, which signals the CO2RR onset. The cathodic current density quickly rises beyond the onset potential in the presence of CO2, and reaches 8.2 mA cm−2 at −0.3 V, which is over three times larger than that measured in N2 (2.6 mA cm−2). Polarization curves of other samples are shown in Supporting Information Figures S13a–S13c. Figure 4 | Electrocatalytic CO2RR performances. (a) Polarization curves of s-mesoPdCu in N2- and CO2-saturated 0.1 M KHCO3. (b) Chronoamperometric curves of s-mesoPdCu nanocubes at different potentials under CO2. (c) Potential-dependent formate selectivity and Faradaic efficiency of s-mesoPdCu. (d) Comparison of the formate selectivity of s-mesoPdCu, s-mesoPd, p-mesoPd, and Pd NPs. (e) Long-term chronoamperometric stability of s-mesoPdCu, s-mesoPd, p-mesoPd, and Pd NPs at −0.4 V. Download figure Download PowerPoint For the product analysis and quantification, we performed chronoamperometric (i∼t) analysis at a few selected working potentials between 0 and −0.5 V for different catalyst samples under study. It is worth highlighting that s-mesoPdCu exhibits stable chronoamperometric responses over the potential regime examined (Figure 4b). Negligible gas products (H2 and CO) are detected. Formate is identified to be the only product from CO2 reduction. Its Faradaic efficiency is analyzed to be 80% at 0 V, which grows to and maintains >90% between −0.1 and −0.5 V (Figure 4c). Corresponding formate partial current density is calculated to increase from 0.3 mA cm−2 at 0 V to 8.1 mA cm−2 at −0.49 V. In stark contrast, we find that the cathodic current density starts to decline once the working potential is biased <−0.2 V for Pd NPs and p-mesoPd, and <−0.3 V for s-mesoPd owing to the catalyst poisoning by trace CO from CO2RR ( Supporting Information Figures S13d–S13f). Figure 4d compares the formate Faradaic efficiency of four different electrocatalysts under study within their respective stable potential windows. Even though all of them exhibit more or less comparable selectivity (and activity as shown in Supporting Information Figure S13g–S13i) in the low overpotential regime, our s-mesoPdCu clearly stands out for its unique capability to maintain the selectivity and stability under working potential as cathodic as −0.5 V—well beyond the conventional stable region for CO2RR on Pd-based materials and far superior to any close competitors reported in the literature ( Supporting Information Table S1). To better demonstrate the performance advantage of s-mesoPdCu, long-term stability was assessed at −0.4 V for 15,000 s, as shown in Figure 4e. The cathodic current density of Pd NPs quickly drops from 5 to 0 mA cm−2 within the first 4000 s, in agreement with the previous observations.36,42,43 The introduction of mesoporosity significantly promotes the CO tolerance and working stability, thanks to the enlarged catalyst surface areas that dilute CO coverage. At the end of 15,000 s, s-mesoPd and p-mesoPd retain the current density of 1.9 and 1.3 mA cm−2, respectively. Their average formate Faradaic efficiency is measured to be 100% and 94%, respectively. The slightly improved stability and selectivity of s-mesoPd over p-mesoPd is attributed to its single-crystalline nature that benefits the charge transfer during electrocatalysis. Finally, alloying Pd with Cu further enhances the stability. s-mesoPdCu still delivers 5.7 mA cm−2 even after 15,000 s with 100% average formate selectivity. We believe that in addition to its advantageous nanostructure, the electronic effect from alloying Pd and Cu is responsible for the observed enhancement. The introduction of a secondary metal with low work function such as Cu is expected to lower the Pd d-band center (in line with the electron transfer from Cu to Pd as revealed from XPS), consequently weakening the CO binding on the alloy catalyst and dramatically promoting its CO tolerance and catalytic stability. After the stability test, s-mesoPdCu is observed to preserve the single-crystalline mesoporous structure as well as the bimetallic nature ( Supporting Information Figure S14). The Pd/Cu molar ratio is measured to only slightly increase to 2.13. Conclusion Here, we developed a facile solution method for the preparation of s-mesoPd and s-mesoPdM nanocubes. The products featured abundant mesoporous channels and, very surprisingly, single-crystallinity throughout each nanocube. We proposed that this unique nanostructure was formed as the collective result of the templating effect of CTAC and the facet-selective capping effect of Cl− ions as supported by a range of control experiments. The large surface areas and penetrating mesoporous channels of our products greatly benefited their catalytic applications. In CO2-saturated 0.1 M KHCO3, the best sample—s-mesoPdCu-catalyzed electrochemical CO2 reduction to formate with great selectivity (>90%) and stability even under the working potential as cathodic as −0.5 V, which was previously believed to be too harsh for Pd-based CO2RR. The electronic effect from alloying Pd and Cu was found to play an additional role in stabilizing the catalyst. Our study demonstrates the great potential and new opportunity for single-crystalline and mesoporous metallic nanostructures. Supporting Information Supporting Information is available and includes schematic self-assembled structures of different surfactants; STEM images, TEM images, SAED patterns, EDS mapping, SAXS patterns, XRD patterns, and XPS spectra of different samples; polarization curves, chronoamperometric responses, and potential-dependent formate selectivity of different samples; and comparison of the formate Faradaic efficiency of our best sample with literature results. Conflict of Interest There is no conflict of interest to report. Funding Information B.L. thanks the Natural Science Foundation of Jiangsu Province (no. BK20180723), the Open Project of State Key Laboratory of Supramolecular Structure and Materials (no. sklssm20
最长约 10秒,即可获得该文献文件

科研通智能强力驱动
Strongly Powered by AbleSci AI
更新
大幅提高文件上传限制,最高150M (2024-4-1)

科研通是完全免费的文献互助平台,具备全网最快的应助速度,最高的求助完成率。 对每一个文献求助,科研通都将尽心尽力,给求助人一个满意的交代。
实时播报
浅眸流年完成签到,获得积分10
1秒前
陈文学完成签到 ,获得积分10
1秒前
1秒前
可爱的小丸子完成签到,获得积分10
1秒前
Orange应助逍遥采纳,获得10
1秒前
景代丝发布了新的文献求助30
1秒前
善良的远锋完成签到,获得积分10
2秒前
2秒前
3秒前
fafa完成签到 ,获得积分10
3秒前
4秒前
yedo发布了新的文献求助10
5秒前
qiaokizhang发布了新的文献求助10
6秒前
6秒前
Lyy完成签到,获得积分10
6秒前
MIST完成签到,获得积分10
8秒前
9秒前
风雪梧桐完成签到,获得积分10
9秒前
酸葡萄完成签到,获得积分10
10秒前
阳佟天川完成签到,获得积分10
11秒前
林小鱼完成签到,获得积分10
11秒前
熙胜发布了新的文献求助10
11秒前
11秒前
那啥完成签到 ,获得积分0
11秒前
12秒前
蒋若风完成签到,获得积分10
12秒前
11完成签到 ,获得积分10
12秒前
麻雀完成签到 ,获得积分10
13秒前
无花果应助liu欣慰采纳,获得10
13秒前
快乐每一天完成签到,获得积分10
13秒前
上山石头完成签到,获得积分10
14秒前
xx发布了新的文献求助10
15秒前
宇文宛菡完成签到 ,获得积分10
15秒前
Neshama发布了新的文献求助30
15秒前
Ysheng发布了新的文献求助10
15秒前
酸汤肥牛完成签到,获得积分10
16秒前
研友_n2QwRL完成签到,获得积分10
17秒前
sschen完成签到,获得积分10
17秒前
kk完成签到 ,获得积分10
17秒前
小钟发布了新的文献求助10
17秒前
高分求助中
The three stars each : the Astrolabes and related texts 1070
Manual of Clinical Microbiology, 4 Volume Set (ASM Books) 13th Edition 1000
Teaching Social and Emotional Learning in Physical Education 900
Boris Pesce - Gli impiegati della Fiat dal 1955 al 1999 un percorso nella memoria 500
[Lambert-Eaton syndrome without calcium channel autoantibodies] 500
少脉山油柑叶的化学成分研究 500
Recherches Ethnographiques sue les Yao dans la Chine du Sud 500
热门求助领域 (近24小时)
化学 材料科学 医学 生物 有机化学 工程类 生物化学 纳米技术 物理 内科学 计算机科学 化学工程 复合材料 遗传学 基因 物理化学 催化作用 电极 光电子学 量子力学
热门帖子
关注 科研通微信公众号,转发送积分 2401626
求助须知:如何正确求助?哪些是违规求助? 2101144
关于积分的说明 5297835
捐赠科研通 1828783
什么是DOI,文献DOI怎么找? 911554
版权声明 560333
科研通“疑难数据库(出版商)”最低求助积分说明 487293