Next Article in Journal
CuS-Based Nanostructures as Catalysts for Organic Pollutants Photodegradation
Next Article in Special Issue
Preparation of Cex-Mn0.8Fe0.2O2 Catalysts and Its Anti-Sulfur Denitration Performance
Previous Article in Journal
Insights into the Redox and Structural Properties of CoOx and MnOx: Fundamental Factors Affecting the Catalytic Performance in the Oxidation Process of VOCs
Previous Article in Special Issue
Elucidating Evidence for the In Situ Reduction of Graphene Oxide by Magnesium Hydride and the Consequence of Reduction on Hydrogen Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cyanogel-Induced Synthesis of RuPd Alloy Networks for High-Efficiency Formic Acid Oxidation

Jiangsu Key Laboratory of New Power Batteries, Jiangsu Collaborative Innovation Centre of Biomedical Functional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(10), 1136; https://doi.org/10.3390/catal12101136
Submission received: 8 September 2022 / Revised: 22 September 2022 / Accepted: 25 September 2022 / Published: 28 September 2022
(This article belongs to the Special Issue Advanced Materials for Application in Catalysis)

Abstract

:
For direct formic acid fuel cells (DFAFC), palladium (Pd)-based alloy catalysts with competitive morphology and elemental composition are essential to boost the performance of the formic acid oxidation reaction (FAOR) in the anode zone. Herein, we design and synthesize RuPdx alloy nano-network structures (ANs) via the facile wet-chemical reduction of Pd-Ru cyanogel (Pdx [Ru(CN)6]y·aH2O) as an effective electrocatalyst for the FAOR. The formation of Pd-Ru cyanogel depends on the facile coordination of K2PdCl4 and K3 [Ru(CN)6]. The unique structure of cyanogel ensures the presentation of a three-dimensional mesoporous morphology and the homogeneity of the elemental components. The as-prepared RuPd3 ANs exhibit good electrocatalytic activity and stability for the FAOR. Notably, the RuPd3 ANs achieve a mass-specific activity of 2068.4 mA mg−1 in FAOR, which shows an improvement of approximately 16.9 times compared to Pd black. Such a competitive FAOR performance of RuPd3 ANs can be attributed to the advantages of structure and composition, which facilitate the exposure of more active sites, accelerate mass/electron transfer rates, and promote gas escape from the catalyst layer, as well as enhance chemical stability.

Graphical Abstract

1. Introduction

Direct Formic Acid Fuel Cell (DFAFC) has been identified as a promising energy storage device by virtue of its low cost, high security, low-temperature adaptability, and theoretically large energy density [1,2,3,4,5]. Formic acid (HCOOH), as a fuel substance of DFAFC, is a clean and renewable energy source and can be obtained through biomass conversion processes [6,7,8,9]. As the anode reaction in a DFAFC, the formic acid oxidation reaction (FAOR, HCOOH → CO2 + H2) performs an essential function in the energy density and charge/discharge cycle stability of the cell. Moreover, as a C1 molecular electrocatalytic reaction, FAOR has substantial research significance [10,11,12]. Pd-based materials are considered to be specific catalysts for the 2e transfer step of FAOR in DFAFC (Scheme 1), possessing a lower polarisation potential and a more rapid kinetic step compared to the indirect pathway of Pt-based nanomaterials [13,14,15]. However, they have the following pressing problems. (1) Commercial Pd-based catalysts are expensive, [16] which limits the large-scale application of fuel cells [17]. (2) The FAOR performance of commercial catalysts is not yet sufficient and has considerable opportunities for improvement [18]. Accordingly, the design of catalysts with low cost and high activity/stability is a high priority for researchers [19,20,21].
The current strategies for enhancing the activity/stability of FAOR in Pd-based metal catalysts focus on electronic and structural modulation [22,23,24]. Electronic modulation is commonly utilized to enhance the intrinsic activity of catalysts [25,26,27]. The alloying/metallic of Pd with other 3d transition metals can achieve both cost reduction and modulation of the valence electrons in the Pd 3d orbitals [28], thus, improving the ability of catalytic activity and anti-toxicity [29], Further, the alloying of Ru with Pd allows for superior resistance to COads poisoning, while reducing the dosage of Pd metal. As a relatively inexpensive noble metal, Ru is only 1/5 the price of Pd. In previous research, alloying Ru with Pd has proven to be effective in improving these properties [30], because the -OH species generated on Ru can facilitate the oxidative removal of toxic intermediates, and the d-band centre of Pd will be shifted by Ru.
In terms of structural control, the architecture of the catalyst directly affects the exposition of the active site in the electrocatalytic reaction and the ability of the produced gas to escape [31,32], having a direct impact on the long-term stability and kinetic conversion efficiency of catalytic reactions, as well as causes the loss of mass transfer during DFAFC operation. Mesoporous-enriched three-dimensional (3D) materials typically have a large specific surface area and channels, which provides appreciable catalytically active sites and mass transport pathways in electrocatalytic reactions [33,34,35]. Moreover, large amounts of product gases are released during the DFAFC operation, whereas the 0 D structure does not facilitate the escape of CO2 and H2, further avoiding the blocking of oxygen by the product gas under high current density. Pd-based alloy catalyst layers with 3D porous structures exhibit greater competitiveness than 0D, 1D and 2D nanomaterials when operating in DFAFC.
The cyanogel-induced synthesis of nanoelectrocatalysts can achieve a combination of these two advantages. The cyanogel-induced synthesis of Pd-based alloy catalysts has been demonstrated in previous studies [36], due to the special coordination pattern of cyanogel that allows the Pd and the other 3d-transition (M) atoms (such as Fe [37], Co [38], Ni [39,40,41], Rh [42], Pt [43] and Cu [44], etc.) to occupy specific positions along the cyanogel skeleton, thus, enabling a predictable morphological and electronic structure during liquid phase reduction, resulting in better long-term stability in electrochemical tests. In addition, the conditions for the synthesis of cyanogel are in line with the concept of low carbon emissions and environmental friendliness. The strategy results in a stable colloidal structure at ambient temperature and pressure. The reduction process is rapid, usually taking only a few seconds to obtain a phase homogeneous alloy, and the high feed conversion rate of the cyanogel reduction reaction results in almost eliminating feedstock waste, as well as allowing for high throughput production.
In this work, for the first time, a series of yellow, orange and tawny Ru-Pd-based cyanogels are conveniently available by using potassium chloropalladite (K2PdCl4) and potassium hexacyanoruthenate (K3[Ru(CN)6]) as precursors. Notably, this is the first report on the usage of Ru-based cyano inorganic complexes for the synthesis of cyanogels. The molar ratio of Ru/Pd in the cyanogels can easily be adjusted by controlling the feeding ratios, and this feature of Pdx[Ru(CN)6]y·aH2O cyanogels facilitates the subsequent construction of alloy products with controllable compositions and adjustable properties. Ultimately, the RuPdx alloy was successfully reduced by NaBH4. The RuPd3 ANs exhibited superior catalytic activity and stability in the electrocatalytic FAOR (Scheme S1).

2. Results and Discussion

2.1. Physical Characterisation of Pdx [Ru(CN)6]y·aH2O Cyanogels

As observed, the stable Pdx[Ru(CN)6]y·aH2O cyanogels can be obtained at a series of feeding ratios of K2PdCl4 and K3[Ru(CN)6] from 1:1, 2:1, and 3:1 (Figure S1). According to the principle of cyanogel formation, the ruthenium cyanide ion ([Ru(CN)6]3−) rapidly coordinates with chloropalladite ion ([PdCl4]2−) and forms a Ru-CN-Pd bond. The aurantiacus Pdx [Ru(CN)6]y·aH2O cyanogel is completed in an instant. The RuPdx alloy networks were grown on a cyanogel skeleton by using an excess of fresh aqueous NaBH4 solution as a reducing agent at room temperature.
The crystal structure of the prepared Pdx[Ru(CN)6]y·aH2O cyanogels were characterized by X-ray diffraction (XRD) tests (Figure 1a). The results show that Pdx[Ru(CN)6]y·aH2O cyanogels exhibit the amorphous peak shape of the classical Prussian-like structure in terms of crystal structure after freeze-drying into aero-cyanogels, which matches the peak positions of other Prussian-like structures reported in the previous research [45]. Furthermore, the series of peaks that are situated on 28.33, 40.49, 50.15, 58.60, 66.35 and 73.70 belong to the standard card KCl (JCPDS 97-015-4214), suggesting the by-product prefers to be KCl, rather than KCN, in the formation of the Pdx[Ru(CN)6]y·aH2O cyanogel:
y K 3 [ R u ( C N ) 6 ] + x K 2 P d C l 4 + a H 2 O P d x [ R u ( C N ) 6 ] y · a H 2 O + ( 2 x + 3 y + n ) K C l
The coordination was further verified by the Fourier transform infrared (FT-IR) spectra (Figure 1b). The sharp peak, stretching the vibration peaks of the cyanogroup, at 2096 cm−1, indicates the formation of a Ru-CN-Pd bond, which further suggests that Ru is linked to Pd by a cyanogroup-bridging coordination [46]. The peaks at 3469 and 1610 cm−1 belong to the hydroxyl stretching and bending vibrational peaks of crystallized water, respectively [44].

2.2. Physical Characterisation of As-Prepared RuPd3 ANs, RuPd2 and RuPd

The RuPd3 ANs and contrast samples (RuPd2 and RuPd) were obtained by a liquid–chemical reduction process, using fresh NaBH4 solution. The atomic content of Pd and Ru was confirmed by energy dispersive spectroscopy (EDS, Figure S3), consisting with the inductively coupled plasma mass spectrometry (ICP-MS) test results (Ru/Pd = 51.1/48.9, 35.4/64.6 and 23.2/76.8). Moreover, the elemental ratio of RuPdx is close to the feeding ratio in the preparative process, indicating that the catalyst prepared by the cyanogel method can achieve a high reaction conversion rate, as well as effectively avoiding the waste of noble metals in the synthetic process. As characterised by XRD patterns (Figure 2a), the 2-Theta diffraction peaks of RuPd3 ANs situated on 39.71, 45.62 and 67.43 are attributed to the (111), (200) and (220) facets, respectively, which are between Pd (JCPDS 97-006-4914) and Ru (JCPDS 97-004-1515) standard cards, suggesting the formation of RuPdx alloy. Compared to RuPd2 and RuPd, the diffraction peaks show a clear leftward shift between the Ru and Pd JCPDS cards, as well as towards the Pd side, which further demonstrates that both Ru and Pd elemental ratios can be changed to form stable alloy phases rather than heterogeneous structures/interfaces. Clearly shown in Figure S2, the lattice parameter appeared on a straight fitted line based on a solid solution between Pd and Ru, in the light of Vegard’s Law: d Ru P d = x d P d + ( 1 x ) · d R u , confirming the well-defined alloy formation in RuPd3 ANs.
Further, an XPS measurement (Figure 2b–d and Figure S4) was performed to investigate the changes in the surface properties of the catalysts under different element ratios. According to the XPS analysis (Figure 2b), the surface ratios of Ru for RuPd3, RuPd2 and RuPd are approximately 0.45, 0.30 and 0.24, respectively, which are consistent with the EDS and ICP results (Table 1), demonstrating that the cyanogel-induced synthesis of RuPdx alloys networks are all homogeneous physical phases rather than core-shell structures.
Moreover, changes in elemental composition can also play a role in the modulation of the electronic structure of the alloy phase, as was demonstrated by comparing the Ru and Pd elemental binding energies of RuPd3 ANs, RuPd2 and RuPd. For RuPd3 ANs, two peaks appeared at the binding energy of 334.94 and 340.24 eV, which were ascribed as Pd 3d5/2 and Pd 3d3/2, respectively (Figure 2c), indicating the negative shift of 0.35 and 0.65 eV to a lower binding energy with respect to that of RuPd2 and RuPd. It is implied by the reduction in binding energy that the electron donation from Ru to Pd and the d-band centre of Pd may be downward-shifted, which favours a weakening of the interaction between the Pd surface and the absorber, so as to boost the electrocatalytic activity in FAOR [28,47]. By contrast, the binding energy of Ru 3p is elevated, with the RuPd3 ANs showing a 0.46 and 1.05 eV positive shift compared to RuPd2 and RuPd, respectively (Figure 2d). In summary, with the increase in Pd content, the binding energy of Ru gradually increases, while Pd shows the opposite shift, further verifying that Ru has a significant modulating effect on the electronic structure of Pd and shows binding energy which is favourable to the electrocatalytic reaction of Pd-based materials at Ru/Pd = 1:3.

2.3. Microscopic Morphological Characterisation

The morphological and structural characterisation of as-prepared RuPd3 ANs were achieved by a scanning electron microscope (SEM) and high-resolution transmission electron microscopy (HRTEM). The 3D porous structure of RuPd3 ANs was confirmed by SEM (Figure S5). Confirmed by the magnified SEM (Figure 3a) and wide-ranged HRTEM images (Figure 3b), the nano-network structure incorporates abundant mesoporous channels, which provide more active sites for the FAOR as well as facilitate the escape of gases from the catalyst layer [45,48]. The surface area of the catalyst for RuPd3 ANs was calculated by Brunauer–Emmett–Teller (BET) to be 245.78 m2 g−1 from the adsorption-desorption isotherm of N2 (Figure S6a). The pore size distribution graph further confirms the porous structure of RuPd3 ANs with a mesoporous pore size of approximately 4.8 nm (Figure S6b). As shown in Figure 3c, the further enlarged HRTEM image revealed clear lattice stripes, which further confirmed the well-crystallised quality of the RuPd3 ANs. The lattice distance of RuPd3 ANs measured by the HRTEM (Figure 3d) is 0.228 nm, which is between Pd (111) and Ru (111), consistent with results from XRD patterns. This further verified the Ru content to be approximately 25%. EDS elemental mapping shows a homogeneous distribution of Ru and Pd along the nano-network (Figure 3e), indirectly verifying the ordered arrangement of the RuPd3 alloy formed on the basis of the cyanogel skeleton.

2.4. Electrochemical Analysis

Cyclic voltammetry tests of RuPd3 ANs, RuPd2, RuPd and Pd black were carried out in a 0.5 M H2SO4 aqueous solution to investigate their electrochemical properties (Figure S7). By calculating the oxygen reduction peak, the electrochemically active surface area (ECSA) of RuPd3 ANs, RuPd2, RuPd and Pd black can be estimated to be 8.1, 7.3, 3.7 and 6.6 m2 g−1, respectively (Figure 4a). The reduction peak potential on RuPd3 ANs is negatively shifted by 56, 166 and 82 mV compared to that of RuPd2, RuPd and Pd black catalysts, respectively, indicating that RuPd3 ANs can provide –OH species at a lower potential, which is conducive to the 2 e transfer pathway of FAOR. It also provides further evidence that the elevated content of Ru/Pd = 1:3 is more conducive to the electrocatalytic FAOR. The electrocatalytic activities for the FAOR of these four catalysts were assessed by CV curves (Figure 4b,c) conducted in N2-saturated 0.5 M H2SO4 + 0.5 M HCOOH at 50 mV s−1. Compared to the Pd black, both the RuPd3 ANs and RuPd2 take on much larger mass-normalized peak current densities for FAOR. The improved FAOR activities of these catalysts could be ascribed to the modification of the electronic structure, which is the reduction in the electronic density of states near the Fermi level. The RuPd3 ANs exhibit superior mass electrocatalytic activity with a peak current density of 2068.4 mA mgPd−1, which was 2.9, 7.6 and 16.9 times higher than that of RuPd2 (705.8 mA mgPd−1), RuPd (277.3 mA mgPd−1), and Pd black (122.2 mA mgPd−1), respectively. The FAOR mass specific activity of RuPd3 ANs is higher than most reported Pd-based electrocatalysts (Table S1), demonstrating that RuPd3 ANs can achieve higher current densities at lower potentials. Moreover, compared to RuPd2 and RuPd catalysts, RuPd3 ANs have lower oxidation potentials for FAOR, with potential advantages of 77 and 222 mV, respectively. In the comparison of area specific activity, RuPd3 ANs display 1.4 times the current density of Pd Black, as well as a negative shift of 82 mV of onset potential for FAOR, indicating that RuPd3 ANs possess the best intrinsic activity. From the Tafel plots in Figure 4e, the RuPd3 ANs present a higher output current density than RuPd2, RuPd and Pd black at the same polarization overpotential, suggesting favourable charge transfer and promoted reaction kinetics of FAOR. The electrochemical impedance spectroscopy (EIS) of RuPd3 ANs and Pd black (Figure S8) shows that RuPd3 ANs exhibit a smaller semi-circular arc in the region of kinetic control (high-frequency zone) than that of Pd black, which indicates that RuPd3 ANs have less charge transfer resistance (Rct) than Pd black in acidic condition. The chronoamperometric measurements shown in Figure S9 were brought into effect in 0.5 M H2SO4 + 0.5 M HCOOH electrolyte at a ceaseless potential of 0.6 V (vs. RHE). After 6000 s, commercial Pd black decreased to 0.4 mA cm−2, while the RuPd3 ANs showed a comparative decrease of 40.13% to 15.1 mA cm−2. To further investigate the electrochemical behaviour of RuPd3 ANs and Pd black during FAOR, 200 cycles of an accelerated endurance test (ADTs) were executed in 0.5 M H2SO4 + 0.5 M HCOOH electrolyte. As shown in Figure S10a, with the increase in the CV cycles, the current density enhances slightly, and then gradually stabilizes after 60 cycles. Compared to the rapid decay of the Pd black counterpart (Figure S10b), RuPd3 ANs exhibit better FAOR stability. Over the next 140 cycles of testing, both RuPd3 ANs and Pd black all show varying degrees of attenuation. At 200th cycles, the mass-normalized activity of RuPd3 ANs decays to 855.0 mA mgPd−1 (41.35% of the current density in the first cycles), which is still 7.0 times higher than that of Pd black. The Pd black decays to 18.2 mA mgPd−1 in the 200th cycle, which is only 14.9% of the current density in the first cycle. Notably, after 60 cycles of ADTs, the FAOR curve for Pd black showed two distinct oxidation peaks, suggesting an indirect pathway to electrocatalytic FAOR after the poisoning of Pd black, while that for RuPd3 ANs maintained its initial electrochemical characteristics, suggesting a clear advantage of RuPd3 ANs over Pd black in terms of their anti-CO toxicity performance. The morphology of the RuPd3 ANs can be well maintained after ADTs testing. As shown in Figure S11, the morphology of RuPd3 ANs was not significantly changed, and the mesoporous-enriched 3D structure did not collapse/agglomerate, benefiting from the structural stability derived from the robust cyanogel skeleton.

3. Materials and Methods

3.1. Electrochemical Analysis

Potassium hexacyanoruthenate (K3[Ru(CN)6]) was supplied by Alfa Aesar Co., Ltd. (Shanghai, China). Potassium chloropalladite (K2PdCl4) was purchased from D.B. Chemical Reagent Co., Ltd. (Shanghai, China). Formic acid (HCOOH), ethanol (C2H5OH), sulfuric acid (H2SO4) and sodium borohydride (NaBH4) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Pd black was purchased from Johnson Matthey Corporation. All the reagents were not further purified.

3.2. Methods—Synthesis of RuPdx Alloy Networks

In a typical synthesis, K3[Ru(CN)6] solution (0.3 mL, 0.1 M) was added to K2PdCl4 (0.3, 0.6 and 0.9 mL, 0.1 M) solution in continuous oscillation. Concussion should stop immediately after mixing. Pdx[Ru(CN)6]y·aH2O cyanogel quickly formed in a matter of minutes. Then, the newly prepared excess NaBH4 (5.0 mL, 0.2 M) was added to Pdx[Ru(CN)6]y·aH2O cyanogel. After two hours, the black RuPd, RuPd2 and RuPd3 ANs were obtained by ultrasonic pulverisation in a cell pulveriser for 4 min, followed by centrifugal washing several times.

3.3. Characterisation

High-resolution transmission electron microscopy (HRTEM) and HADDF-STEM images were obtained on the Talos F200x G2 microscope operated with an accelerating voltage of 200 kV (Wuhan, China). The scanning electron microscopy (SEM) was characterized by Hitachi S4800 scanning electron microscopy (Nanjing, China). EDS measurements were conducted on Oxford Instruments ULTIM MAX 170 (Nanjing, China). The concentrations of metals were then measured with a Thermo Scientific Plasma Quad inductively coupled plasma mass spectrometry (ICP-MS, (Nanjing, China) and EDS (Nanjing, China). XPS measurements were conducted on a Thermo VG Scientific ESCALAB 250 spectrometer with an Al Kα radiator (Zibo, China). The binding energy was calibrated by means of the C 1s peak energy of 284.6 eV. XRD analyses were performed on a Model D/max-rC X-ray diffractometer employing Kα radiation (λ = 0.15406 nm) (Nanjing, China). BET measurements were required on MicrotracBEL BELSORP-max (Zibo, China) with adsorptive of N2. The apparatus temperature and adsorption temperature are 0 °C and 77 K, respectively.

3.4. Electrochemical Measurements

Electrochemical measurements were performed on a CHI 760E analyser (Chenhua Co. Shanghai, China) with a conventional three-electrode system. The glassy carbon electrode loaded with catalyst was used as the working electrode, the graphite rod as the counter electrode, and the saturated calomel electrode (SCE) as the reference electrode for the FAOR tests. In electrochemical testing, 4.0 mg of electrocatalysts was dispersed into a mixture of 0.7 mL of alcohol, 1.2 mL of deionized water and 0.1 mL of Nafion solution (5 wt%), and the above substances were mixed by sonication thoroughly for 30 min. Then, 10 μL of the acquired catalyst ink (2 mg mL−1) was dropped onto the surface of the glass carbon electrode of 3.0 mm in diameter. Thus, the mass loading of the catalyst on the GC electrode is 0.285 mg cm−2. The catalysts were pre-activated in a cyclic voltammetric window in 0.5 M H2SO4 solution before the ADTs were tested.

4. Conclusions

In summary, we have developed a new strategy for the construction of Pdx[Ru(CN)6]y·aH2O cyanogels, using K3 [Ru(CN)6] as metal salt precursors, by utilizing the principles of cyanogel synthesis, as well as RuPdx alloy network-like catalysts, which were further obtained at ambient temperature and pressure. During electrocatalytic FAOR testing, the RuPd3 ANs catalyst demonstrated a competitive electrocatalytic mass/area specific activity, electron/proton transfer rates, and operational long-term stability in comparison to commercial Pd black catalyst and other element ratio component of RuPdx alloys (1:1 and 1:2). The competitive FAOR activity can be attributed to the advantage of a stable 3D nano-network structure, and the modulating effect of electronic effects due to the specific RuPd3 component ratio. The RuPd3 ANs show good prospects for the catalyst layer in DFAFC. In addition, this work provides guidance for the subsequent synthesis of cyanogels using ruthenium cyanide as the central ligand.

Supplementary Materials

The following supplementary information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12101136/s1, Table S1: FAOR activity comparisons of RuPd3 ANs with the previous reported catalysts [2,10,11,29,30,44,49,50,51,52,53,54,55,56,57,58,59,60,61]; Scheme S1: Graphical abstract; Figure S1: Photographs of Pdx[Ru(CN)6]y·aH2O cyanogels; Figure S2: The plot of lattice parameter versus Pd atomic percentage of RuPd3 ANs; Figure S3: The EDS spectrum of (a) RuPd, (b) RuPd2 and (c) RuPd3 ANs, respectively; Figure S4: The XPS survey of RuPd3 ANs; Figure S5: The large-scale SEM images of RuPd3 ANs; Figure S6: (a) N2 adsorption-desorption isotherms and (b) Brunauer–Emmett–Teller (BET) pore calculation of RuPd3 ANs; Figure S7: Cyclic voltammetry curves of RuPd3 ANs, RuPd2, RuPd and Pd black in 0.5 M H2SO4 solution; Figure S8: Nyquist curves of RuPd3 ANs and Pd Black in 0.5 M H2SO4 + 0.5 M HCOOH electrolyte; Figure S9: Chronoamperometric curves obtained at 0.6 V (vs. RHE) in N2-saturated 0.5 M H2SO4 + 0.5 M HCOOH solution; Figure S10: The 1st, 60th and 200th FAOR curves for accelerated durability tests (ADTs) tests in N2-saturated 0.5 M H2SO4 + 0.5 M HCOOH solution, compared with (a) RuPd3 ANs and (b) Pd black; Figure S11: The TEM image of RuPd3 ANs after 200 cycles ADTs test.

Author Contributions

Conceptualization, Q.L., J.Z., D.S. and Y.T.; Synthesis and preparation of catalysts, Q.L., W.Y., J.Z. and Y.R.; Characterization, Q.L., W.Y., J.L. and X.Z.; writing—original draft, Q.L.; writing—review and editing, Q.L., D.S. and Y.T.; funding acquisition, Q.L. and Y.T. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge the financial support from the National Natural Science Foundation of China (Grant No. 21875112 and Grant No. 22072067), the Natural Science Foundation of Jiangsu Province (Grant No. BK20171473) and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (KYCX22_1551). The authors also thank the support from the National and Local Joint Engineering Research Centre of Biomedical Functional Materials, a project sponsored by the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the suggestions made by Gengtao Fu, Zhijuan Li and Chuang Fan for this work, and Mingxing Gong for his support in HRTEM technology, as well as Anzhou Yang for his support in the graphical abstract.

Conflicts of Interest

There are no conflict of interest to declare.

References

  1. Lv, F.; Zhang, W.; Sun, M.; Lin, F.; Wu, T.; Zhou, P.; Yang, W.; Gao, P.; Huang, B.; Guo, S. Au Clusters on Pd Nanosheets Selectively Switch the Pathway of Ethanol Electrooxidation: Amorphous/Crystalline Interface Matters. Adv. Energy Mater. 2021, 11, 2100187. [Google Scholar] [CrossRef]
  2. Yang, L.; Li, G.; Chang, J.; Ge, J.; Liu, C.; Vladimir, F.; Wang, G.; Jin, Z.; Xing, W. Sea urchin-like Aucore@Pdshell electrocatalysts with high FAOR performance: Coefficient of lattice strain and electrochemical surface area. Appl. Catal. B 2020, 260, 118200. [Google Scholar] [CrossRef]
  3. Feng, L.; Cui, Z.; Yan, L.; Xing, W.; Liu, C. The enhancement effect of MoOx on Pd/C catalyst for the electrooxidation of formic acid. Electrochim. Acta 2011, 56, 2051–2056. [Google Scholar] [CrossRef]
  4. Zhang, J.; Wu, L.; Xu, L.; Sun, D.; Sun, H.; Tang, Y. Recent advances in phosphorus containing noble metal electrocatalysts for direct liquid fuel cells. Nanoscale 2021, 13, 16052–16069. [Google Scholar] [CrossRef] [PubMed]
  5. Jiao, K.; Xuan, J.; Du, Q.; Bao, Z.; Xie, B.; Wang, B.; Zhao, Y.; Fan, L.; Wang, H.; Hou, Z.; et al. Designing the next generation of proton-exchange membrane fuel cells. Nature 2021, 595, 361–369. [Google Scholar] [CrossRef]
  6. Wang, T.; Jiang, Y.; He, J.; Li, F.; Ding, Y.; Chen, P.; Chen, Y. Porous palladium phosphide nanotubes for formic acid electrooxidation. Carbon Energy 2022, 4, 283–293. [Google Scholar] [CrossRef]
  7. Wang, Y.; Jiang, X.; Fu, G.; Li, Y.; Tang, Y.; Lee, J.; Tang, Y. Cu5Pt Dodecahedra with Low-Pt Content: Facile Synthesis and Outstanding Formic Acid Electrooxidation. ACS Appl. Mater. Interfaces 2019, 11, 34869–34877. [Google Scholar] [CrossRef] [PubMed]
  8. Meng, Q.; Yang, X.; Wang, X.; Xiao, M.; Li, K.; Jin, Z.; Ge, J.; Liu, C.; Xing, W. Preparation Strategy Using Pre-Nucleation Coupled with In Situ Reduction for a High-Performance Catalyst towards Selective Hydrogen Production from Formic Acid. Catalysts 2022, 12, 325. [Google Scholar] [CrossRef]
  9. Ma, Z.; Legrand, U.; Pahija, E.; Tavares, J.R.; Boffito, D.C. From CO2 to Formic Acid Fuel Cells. Ind. Eng. Chem. Res. 2020, 60, 803–815. [Google Scholar] [CrossRef]
  10. Hossain, S.K.S.; Alwi, M.M.; Saleem, J.; Al-Hashem, H.T.; McKay, G.; Mansour, S.; Ali, S.S. Bimetallic Pd-Co Nanoparticles Supported on Nitrogen-Doped Reduced Graphene Oxide as Efficient Electrocatalysts for Formic Acid Electrooxidation. Catalysts 2021, 11, 910. [Google Scholar] [CrossRef]
  11. Teng, Z.; Li, M.; Li, Z.; Liu, Z.; Fu, G.; Tang, Y. Facile synthesis of channel-rich ultrathin palladium-silver nanosheets for highly efficient formic acid electrooxidation. Mater. Today Energy 2021, 19, 100596. [Google Scholar] [CrossRef]
  12. Fang, Z.; Chen, W. Recent advances in formic acid electro-oxidation: From the fundamental mechanism to electrocatalysts. Nanoscale Adv. 2021, 3, 94–105. [Google Scholar] [CrossRef] [PubMed]
  13. Rettenmaier, C.; Aran-Ais, R.; Timoshenko, J.; Rizo, R.; Jeon, H.; Kuhl, S.; Chee, S.; Bergmann, A.; Cuenya, B.R. Enhanced Formic Acid Oxidation over SnO2-decorated Pd Nanocubes. ACS Catal. 2020, 10, 14540–14551. [Google Scholar] [CrossRef] [PubMed]
  14. Chang, J.; Feng, L.; Liu, C.; Xing, W.; Hu, X. An effective Pd-Ni(2)P/C anode catalyst for direct formic acid fuel cells. Angew. Chem. Int. Ed. 2014, 53, 122–126. [Google Scholar] [CrossRef]
  15. Bhaskaran, R.; Abraham, B.G.; Chetty, R. Recent advances in electrocatalysts, mechanism, and cell architecture for direct formic acid fuel cells. WIREs Energy Environ. 2021, 11, e419. [Google Scholar] [CrossRef]
  16. Liu, Q.; Wang, J.; Zhang, J.; Yan, Y.; Qiu, X.; Wei, S.; Tang, Y. In situ immobilization of isolated Pd single-atoms on graphene by employing amino-functionalized rigid molecules and their prominent catalytic performance. Catal. Sci. Technol. 2020, 10, 450–457. [Google Scholar] [CrossRef]
  17. Chen, D.; Pei, S.; He, Z.; Shao, H.; Wang, J.; Wang, K.; Wang, Y.; Jin, Y. High Active PdSn Binary Alloyed Catalysts Supported on B and N Codoped Graphene for Formic Acid Electro-Oxidation. Catalysts 2020, 10, 751. [Google Scholar] [CrossRef]
  18. Jiang, X.; Xiong, Y.; Wang, Y.; Wang, J.; Li, N.; Zhou, J.; Fu, G.; Sun, D.; Tang, Y. Treelike two-level PdxAgy nanocrystals tailored for bifunctional fuel cell electrocatalysis. J. Mater. Chem. A 2019, 7, 5248–5257. [Google Scholar] [CrossRef]
  19. Xia, B.; Wu, H.; Yan, Y.; Wang, H.; Wang, X. One-pot synthesis of platinum nanocubes on reduced graphene oxide with enhanced electrocatalytic activity. Small 2014, 10, 2336–2339. [Google Scholar] [CrossRef]
  20. Ye, S.; Feng, J.; Li, G. Pd Nanoparticle/CoP Nanosheet Hybrids: Highly Electroactive and Durable Catalysts for Ethanol Electrooxidation. ACS Catal. 2016, 6, 7962–7969. [Google Scholar] [CrossRef]
  21. Peng, L.; Wei, Z. Catalyst Engineering for Electrochemical Energy Conversion from Water to Water: Water Electrolysis and the Hydrogen Fuel Cell. Engineering 2020, 6, 653–679. [Google Scholar] [CrossRef]
  22. Yu, H.; Zhou, T.; Wang, Z.; Xu, Y.; Li, X.; Wang, L.; Wang, H. Defect-Rich Porous Palladium Metallene for Enhanced Alkaline Oxygen Reduction Electrocatalysis. Angew. Chem. Int. Ed. 2021, 60, 12027–12031. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, L.; Wang, F.; Wang, S.; Huang, H.; Meng, X.; Ouyang, Y.; Yuan, W.; Guo, C.; Li, C. Layered and Heterostructured Pd/PdWCr Sheet–Assembled Nanoflowers as Highly Active and Stable Electrocatalysts for Formic Acid Oxidation. Adv. Funct. Mater. 2020, 30, 2003933. [Google Scholar] [CrossRef]
  24. Xia, B.; Chen, Y. Preface to special issue on electrocatalysis for sustainable energy. Chin. J. Catal. 2022, 43, 1379. [Google Scholar] [CrossRef]
  25. Wang, X.; Tang, Y.; Lee, J.; Fu, G. Recent advances in rare-earth-based materials for electrocatalysis. Chem. Catal. 2022, 2, 967–1008. [Google Scholar] [CrossRef]
  26. Wang, X.; Zhu, Y.; Li, H.; Lee, M.; Tang, Y.; Fu, G. Rare-Earth Single-Atom Catalysts: A New Frontier in Photo/Electrocatalysis. Small Methods 2022, 6, 2200413. [Google Scholar] [CrossRef]
  27. Wang, X.; Wang, J.; Wang, P.; Li, L.; Zhang, X.; Sun, D.; Li, Y.; Tang, Y.; Wang, Y.; Fu, G. Engineering 3d-2p-4f Gradient Orbital Coupling to Enhance Electrocatalytic Oxygen Reduction. Adv. Mater. 2022, 2206540. [Google Scholar] [CrossRef]
  28. Cao, K.; Yang, H.; Bai, S.; Xu, Y.; Yang, C.; Wu, Y.; Xie, M.; Cheng, T.; Shao, Q.; Huang, X. Efficient Direct H2O2 Synthesis Enabled by PdPb Nanorings via Inhibiting the O–O Bond Cleavage in O2 and H2O2. ACS Catal. 2021, 11, 1106–1118. [Google Scholar] [CrossRef]
  29. Yan, X.; Hu, X.; Fu, G.; Xu, L.; Lee, J.; Tang, Y. Facile Synthesis of Porous Pd3Pt Half-Shells with Rich “Active Sites” as Efficient Catalysts for Formic Acid Oxidation. Small 2018, 14, e1703940. [Google Scholar] [CrossRef]
  30. Liu, S.; Wang, Z.; Zhang, H.; Yin, S.; Xu, Y.; Li, X.; Wang, L.; Wang, H. B-Doped PdRu nanopillar assemblies for enhanced formic acid oxidation electrocatalysis. Nanoscale 2020, 12, 19159–19164. [Google Scholar] [CrossRef]
  31. Ren, J.; Yao, Y.; Yuan, Z. Fabrication strategies of porous precious-metal-free bifunctional electrocatalysts for overall water splitting: Recent advances. Green Energy Environ. 2021, 6, 620–643. [Google Scholar] [CrossRef]
  32. Li, Z.; Li, M.; Wang, X.; Fu, G.; Tang, Y. The use of amino-based functional molecules for the controllable synthesis of noble-metal nanocrystals: A minireview. Nanoscale Adv. 2021, 3, 1813–1829. [Google Scholar] [CrossRef]
  33. Zaman, S.; Tian, X.; Su, Y.; Cai, W.; Yan, Y.; Qi, R.; Douka, A.; Chen, S.; You, B.; Liu, H.; et al. Direct integration of ultralow-platinum alloy into nanocarbon architectures for efficient oxygen reduction in fuel cells. Sci. Bull. 2021, 66, 2207–2216. [Google Scholar] [CrossRef]
  34. Zhang, H.; Wang, H.; Xu, Y.; Zhuo, S.; Yu, Y.; Zhang, B. Conversion of Sb2Te3 hexagonal nanoplates into three-dimensional porous single-crystal-like network-structured Te plates using oxygen and tartaric acid. Angew. Chem. Int. Ed. 2012, 51, 1459–1463. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, D.; Jiang, X.; Lin, Z.; Zeng, X.; Zhu, Y.; Wang, Y.; Gong, M.; Tang, Y.; Fu, G. Ethanol-Induced Hydrogen Insertion in Ultrafine IrPdH Boosts pH-Universal Hydrogen Evolution. Small 2022, 18, 2204063. [Google Scholar] [CrossRef]
  36. Lin, S.; Yao, Y.; Zhang, L.; Feng, J.; Wang, J. Cyanogel and its derived-materials: Properties, preparation methods, and electrochemical applications. Mater. Today Energy 2021, 20, 100701. [Google Scholar] [CrossRef]
  37. Wan, J.; Liu, Z.; Yang, X.; Cheng, P.; Yan, C. Cyanogel-Derived Synthesis of Porous PdFe Nanohydrangeas as Electrocatalysts for Oxygen Reduction Reaction. Nanomaterials 2021, 11, 3382. [Google Scholar] [CrossRef]
  38. Burgess, C.M.; Yao, N.; Bocarsly, A.B. Stabilizing cyanosols: Amorphous cyanide bridged transition metal polymer nanoparticles. J. Mater. Chem. 2009, 19, 8846–8855. [Google Scholar] [CrossRef]
  39. Liu, X.; Xu, G.; Chen, Y.; Lu, T.; Tang, Y.; Xing, W. A strategy for fabricating porous PdNi@Pt core-shell nanostructures and their enhanced activity and durability for the methanol electrooxidation. Sci. Rep. 2015, 5, 7619. [Google Scholar] [CrossRef]
  40. Liu, H.; Liu, X.; Li, Y.; Jia, Y.; Tang, Y.; Chen, Y. Hollow PtNi alloy nanospheres with enhanced activity and methanol tolerance for the oxygen reduction reaction. Nano Res. 2016, 9, 3494–3503. [Google Scholar] [CrossRef]
  41. Liu, Z.; Yang, X.; Cui, L.; Shi, Z.; Lu, B.; Guo, X.; Zhang, J.; Xu, L.; Tang, Y.; Xiang, Y. High-Performance Oxygen Reduction Electrocatalysis Enabled by 3D PdNi Nanocorals with Hierarchical Porosity. Part. Part. Syst. Charact. 2018, 35, 1700366. [Google Scholar] [CrossRef]
  42. Bai, J.; Xue, Q.; Zhao, Y.; Jiang, J.; Zeng, J.; Yin, S.; Chen, Y. Component-Dependent Electrocatalytic Activity of Ultrathin PdRh Alloy Nanocrystals for the Formate Oxidation Reaction. ACS Sustain. Chem. Eng. 2018, 7, 2830–2836. [Google Scholar] [CrossRef]
  43. Liu, X.; Zhang, Y.; Gong, M.; Tang, Y.; Lu, T.; Chen, Y.; Lee, J. Facile synthesis of corallite-like Pt–Pd alloy nanostructures and their enhanced catalytic activity and stability for ethanol oxidation. J. Mater. Chem. A 2014, 2, 13840–13844. [Google Scholar] [CrossRef]
  44. Liu, Q.; Li, Z.; Zhou, X.; Xiao, J.; Han, Z.; Jiang, X.; Fu, G.; Tang, Y. Cyanogel--Induced PdCu Alloy with Pd–Enriched Surface for Formic Acid Oxidation and Oxygen Reduction. Adv. Energy Sustain. Res. 2022, 2200067. [Google Scholar] [CrossRef]
  45. Liu, Z.; Yang, X.; Lu, B.; Shi, Z.; Sun, D.; Xu, L.; Tang, Y.; Sun, S. Delicate topotactic conversion of coordination polymers to Pd porous nanosheets for high-efficiency electrocatalysis. Appl. Catal. B 2019, 243, 86–93. [Google Scholar] [CrossRef]
  46. Shi, H.; Fang, Z.; Zhang, X.; Li, F.; Tang, Y.; Zhou, Y.; Wu, P.; Yu, G. Double-Network Nanostructured Hydrogel-Derived Ultrafine Sn-Fe Alloy in Three-Dimensional Carbon Framework for Enhanced Lithium Storage. Nano Lett. 2018, 18, 3193–3198. [Google Scholar] [CrossRef]
  47. Liu, H.; Li, J.; Wang, L.; Tang, Y.; Xia, B.Y.; Chen, Y. Trimetallic PtRhNi alloy nanoassemblies as highly active electrocatalyst for ethanol electrooxidation. Nano Res. 2017, 10, 3324–3332. [Google Scholar] [CrossRef]
  48. Ahmad, A.; Khan, S.; Tariq, S.; Luque, R.; Verpoort, F. Self-sacrifice MOFs for heterogeneous catalysis: Synthesis mechanisms and future perspectives. Mater. Today 2022, 55, 137–169. [Google Scholar] [CrossRef]
  49. Zhang, H.; Cen, K.; Wen, J.; Qi, M.; Ge, X.; Liu, Y.; Sun, D.; Tang, Y. One-Pot Synthesis of Porous Palladium Nanorods for Efficient Formic Acid Electro-Oxidation Reaction. Sci. Adv. Mater. 2017, 10, 460–466. [Google Scholar] [CrossRef]
  50. Zhang, L.; Sui, Q.; Tang, T.; Chen, Y.; Zhou, Y.; Tang, Y.; Lu, T. Surfactant-free palladium nanodendrite assemblies with enhanced electrocatalytic performance for formic acid oxidation. Electrochem. Commun. 2013, 32, 43–46. [Google Scholar] [CrossRef]
  51. Qiu, X.; Zhang, H.; Dai, Y.; Zhang, F.; Wu, P.; Wu, P.; Tang, Y. Sacrificial Template-Based Synthesis of Unified Hollow Porous Palladium Nanospheres for Formic Acid Electro-Oxidation. Catalysts 2015, 5, 992–1002. [Google Scholar] [CrossRef]
  52. Qiu, X.; Zhang, H.; Wu, P.; Zhang, F.; Wei, S.; Sun, D.; Xu, L.; Tang, Y. One-Pot Synthesis of Freestanding Porous Palladium Nanosheets as Highly Efficient Electrocatalysts for Formic Acid Oxidation. Adv. Funct. Mater. 2017, 27, 1603852. [Google Scholar] [CrossRef]
  53. Xu, L.; Wang, K.; Yuan, Q. One-pot synthesis of lotus-shaped Pd–Cu hierarchical superstructure crystals for formic acid oxidation. RSC Adv. 2016, 6, 96079–96083. [Google Scholar] [CrossRef]
  54. Xu, J.; Zhao, M.; Yamaura, S.-I.; Jin, T.; Asao, N. Core–shell Pd–P@Pt nanoparticles as efficient catalysts for electrooxidation of formic acid. J. Appl. Electrochem. 2016, 46, 1109–1118. [Google Scholar] [CrossRef]
  55. Wang, H.; Li, Y.; Li, C.; Wang, Z.; Xu, Y.; Li, X.; Xue, H.; Wang, L. Hyperbranched PdRu nanospine assemblies: An efficient electrocatalyst for formic acid oxidation. J. Mater. Chem. A 2018, 6, 17514–17518. [Google Scholar] [CrossRef]
  56. Yang, F.; Zhang, Y.; Liu, P.-F.; Cui, Y.; Ge, X.-R.; Jing, Q.-S. Pd–Cu alloy with hierarchical network structure as enhanced electrocatalysts for formic acid oxidation. Int. J. Hydrogen Energy 2016, 41, 6773–6780. [Google Scholar] [CrossRef]
  57. Shen, T.; Chen, S.; Zeng, R.; Gong, M.; Zhao, T.; Lu, Y.; Liu, X.; Xiao, D.; Yang, Y.; Hu, J.; et al. Tailoring the Antipoisoning Performance of Pd for Formic Acid Electrooxidation via an Ordered PdBi Intermetallic. ACS Catal. 2020, 10, 9977–9985. [Google Scholar] [CrossRef]
  58. Yang, G.; Chen, Y.; Zhou, Y.; Tang, Y.; Lu, T. Preparation of carbon supported Pd–P catalyst with high content of element phosphorus and its electrocatalytic performance for formic acid oxidation. Electrochem. Commun. 2010, 12, 492–495. [Google Scholar] [CrossRef]
  59. Shen, T.; Lu, Y.; Gong, M.; Zhao, T.; Hu, Y.; Wang, D. Optimizing Formic Acid Electro-oxidation Performance by Restricting the Continuous Pd Sites in Pd–Sn Nanocatalysts. ACS Sustain. Chem. Eng. 2020, 8, 12239–12247. [Google Scholar] [CrossRef]
  60. Xu, Y.; Yu, S.; Ren, T.; Li, C.; Yin, S.; Wang, Z.; Li, X.; Wang, L.; Wang, H. A quaternary metal–metalloid–nonmetal electrocatalyst: B, P-co-doping into PdRu nanospine assemblies boosts the electrocatalytic capability toward formic acid oxidation. J. Mater. Chem. A 2020, 8, 2424–2429. [Google Scholar] [CrossRef]
  61. Goswami, C.; Saikia, H.; Tada, K.; Tanaka, S.; Sudarsanam, P.; Bhargava, S.K.; Bharali, P. Bimetallic Palladium–Nickel Nanoparticles Anchored on Carbon as High-Performance Electrocatalysts for Oxygen Reduction and Formic Acid Oxidation Reactions. ACS Appl. Energy Mater. 2020, 3, 9285–9295. [Google Scholar] [CrossRef]
Scheme 1. The direct pathway involved in the HCOOH electrooxidation mechanism.
Scheme 1. The direct pathway involved in the HCOOH electrooxidation mechanism.
Catalysts 12 01136 sch001
Figure 1. (a) X-ray diffraction (XRD) pattern and (b) Fourier transform infrared (FT-IR) spectrum of Pdx[Ru(CN)6]y·aH2O aero-cyanogel.
Figure 1. (a) X-ray diffraction (XRD) pattern and (b) Fourier transform infrared (FT-IR) spectrum of Pdx[Ru(CN)6]y·aH2O aero-cyanogel.
Catalysts 12 01136 g001
Figure 2. (a) XRD pattern; (b) X-ray photoelectron spectroscopy (XPS) spectrum of surface element ratio; (c) XPS spectrum of Pd 3d; and (d) Ru 3p for RuPd3 ANs (Orange), RuPd2 (Blue) and RuPd (Green).
Figure 2. (a) XRD pattern; (b) X-ray photoelectron spectroscopy (XPS) spectrum of surface element ratio; (c) XPS spectrum of Pd 3d; and (d) Ru 3p for RuPd3 ANs (Orange), RuPd2 (Blue) and RuPd (Green).
Catalysts 12 01136 g002
Figure 3. (a) Scanning electron microscope (SEM) image; (bd) high-resolution transmission electron microscope (HRTEM) images of RuPd3 ANs; (e) EDS mapping spectrum of RuPd3 ANs.
Figure 3. (a) Scanning electron microscope (SEM) image; (bd) high-resolution transmission electron microscope (HRTEM) images of RuPd3 ANs; (e) EDS mapping spectrum of RuPd3 ANs.
Catalysts 12 01136 g003
Figure 4. Comparison of the formic acid oxidation reaction (FAOR) performances of RuPd3 ANs, RuPd2, RuPd and Pd black. (a) Cyclic voltammetry (CV) curves (enlarged image) in 0.5 M H2SO4 solution; (b) mass-normalized and (c) area-normalized FAOR curves in N2-saturated 0.5 M H2SO4 + 0.5 M HCOOH solution; (d) Tafel plots.
Figure 4. Comparison of the formic acid oxidation reaction (FAOR) performances of RuPd3 ANs, RuPd2, RuPd and Pd black. (a) Cyclic voltammetry (CV) curves (enlarged image) in 0.5 M H2SO4 solution; (b) mass-normalized and (c) area-normalized FAOR curves in N2-saturated 0.5 M H2SO4 + 0.5 M HCOOH solution; (d) Tafel plots.
Catalysts 12 01136 g004
Table 1. The element analysis (Ru content) of RuPd, RuPd2 and RuPd3 Ans.
Table 1. The element analysis (Ru content) of RuPd, RuPd2 and RuPd3 Ans.
Sample NameEDS AnalysisXPS AnalysisICP Analysis
RuPd0.470.450.51
RuPd20.300.300.35
RuPd30.120.240.23
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Q.; Yan, W.; Zhang, J.; Ren, Y.; Liu, J.; Zeng, X.; Sun, D.; Tang, Y. Cyanogel-Induced Synthesis of RuPd Alloy Networks for High-Efficiency Formic Acid Oxidation. Catalysts 2022, 12, 1136. https://doi.org/10.3390/catal12101136

AMA Style

Liu Q, Yan W, Zhang J, Ren Y, Liu J, Zeng X, Sun D, Tang Y. Cyanogel-Induced Synthesis of RuPd Alloy Networks for High-Efficiency Formic Acid Oxidation. Catalysts. 2022; 12(10):1136. https://doi.org/10.3390/catal12101136

Chicago/Turabian Style

Liu, Qicheng, Wei Yan, Jiachen Zhang, Yi Ren, Jiaqi Liu, Xin Zeng, Dongmei Sun, and Yawen Tang. 2022. "Cyanogel-Induced Synthesis of RuPd Alloy Networks for High-Efficiency Formic Acid Oxidation" Catalysts 12, no. 10: 1136. https://doi.org/10.3390/catal12101136

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop