Next Article in Journal
Enhanced Photoluminescence of Gd3Al4GaO12: Cr3+ by Energy Transfers from Co-Doped Dy3+
Next Article in Special Issue
Electrochemical Detection of Olivetol Based on Poly(L-Serine) Film Layered Copper Oxide Modified Carbon Paste Electrode (p-L-Serine/CuO/CPE)
Previous Article in Journal
Toxicity Assessment of [177Lu]Lu−iFAP/iPSMA Nanoparticles Prepared under GMP-Compliant Radiopharmaceutical Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pd3Co1 Alloy Nanocluster on the MWCNT Catalyst for Efficient Formic Acid Electro-Oxidation

College of Chemistry and Materials Engineering, Huaihua University, Huaihua 418008, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to the work.
Nanomaterials 2022, 12(23), 4182; https://doi.org/10.3390/nano12234182
Submission received: 2 November 2022 / Revised: 19 November 2022 / Accepted: 22 November 2022 / Published: 25 November 2022

Abstract

:
In this study, the Pd3Co1 alloy nanocluster from a multiwalled carbon nanotube (MWCTN) catalyst was fabricated in deep eutectic solvents (DESs) (referred to Pd3Co1/CNTs). The catalyst shows a better mass activity towards the formic acid oxidation reaction (FAOR) (2410.1 mA mgPd−1), a better anti-CO toxicity (0.36 V) than Pd/CNTs and commercial Pd/C. The improved performance of Pd3Co1/CNTs is attributed to appropriate Co doping, which changed the electronic state around the Pd atom, lowered the d-band of Pd, formed a new Pd-Co bond act at the active sites, affected the adsorption of the toxic intermediates and weakened the dissolution of Pd; moreover, with the assistance of DES, the obtained ultrafine Pd3Co1 nanoalloy exposes more active sites to enhance the dehydrogenation process of the FAOR. The study shows a new way to construct a high-performance Pd-alloy catalyst for the direct formic acid fuel cell.

1. Introduction

The development of fuel cells is one of the important ways to achieve carbon neutrality [1,2,3,4,5]. A direct formic acid fuel cell (DFAFC), with the following advantages, a high theoretical energy density, a low operating temperature and easy management, are attracting a lot of interest [6,7,8,9,10].
Palladium (Pd) has a good catalytic performance for DFAFC [11,12,13,14]. Various Pd-based bimetallic alloy catalysts (e.g., PdCu [15], PdMn [16], PdW [17], PdCO [18], PdIr [19], PdSn [20], PdNi [21]) have been prepared to improve the inert FAOR because these alloys exhibited the ability to get rid of OH species. Among them, the PdCo-alloys have attracted great attention. Zhang et al. fabricated a PdCo Nds-RGO catalyst exhibiting the excellent peak current density toward the FAOR [22]. Li et al. prepared a PdCo0.70/C catalyst showing an enhanced mass activity toward the FAOR [23]. Hence, PdCo-based catalysts have the great potential in a DFAFC. However, the instability of the catalysts caused by the dissolution of metal, the Ostwald ripening, and the corrosion of the carbon carrier have not been solved [22,24,25]. Therefore, how to improve the electrochemical catalytic activity and stability of the PdCo-based catalysts are the current focus of researchers’ efforts.
Doping [17,26,27,28] is considered to be the most effective method to improve the stability and activity of Pd-alloy catalysts because the doped metal can change the electronic state around the Pd atom, affecting the adsorption of the toxic intermediates and weakening the dissolution of Pd. Tang et al. doped Co into Pd to form an alloy, which significantly improved the activity and stability of the catalyst [29]. In addition, Pd can also form alloy catalysts with other metals (Cu, Ni, Ru) by doping, which is characterized by the improved intrinsic activity and stability toward the FAOR [30,31,32].
It is well known that increasing the loading rate of metal nanoparticles on a support is an effective method to improve the activity of the catalysts. With these advantages, a good electrical conductivity, thermal stability, electrochemical stability, a negligible vapor pressure, non-toxicity and biodegradability, DES [33,34,35] was used as a green solvent to prepare catalysts. In our previous study, a series of Pt- and Pd-based high-performance catalysts were prepared using DES as the reaction solvent [36,37,38,39,40]. Fan et al. prepared the alloy catalysts (PdSn [20], PtCo [34], PtCu [33], PtV [41], PtLa [42]) by a chemical reduction or electrochemistry method in DES, that plays an important role in controlling the shape of these nanocrystallines. These catalysts showed an excellent mass activity in the fuel cell. Zhuo et al. electrochemically synthesized the AuPt nanopowers in DES and applied them for the organic electrooxidation [43]. Wei et al., prepared concave cubic PtSm alloy nanocrystals [44], concave-disdyakis triacontahedral Pd nanocrystals [45] and cubic Pt93Ir7 [46] with high-index facets that exhibit a higher electrocatalytic activity and stability. Hammons et al. prepared Pd nanoparticles electrodeposited from DES and revealed the interaction between the particles and the solvent, showing that this special solvent can stabilize these electrodeposited Pd nanoparticles [47]. Therefore, the high-performance PdCo catalysts are expected to be prepared in DES.
MWCNTs, with a unique morphology and excellent properties, have attracted much more interest in a variety of fields [48,49,50,51]. Therefore, in DES, we doped the Co element into Pd to form the PdCo alloy structure and then loaded it on MWCNTs (Pd3Co1/CNTs) (Scheme 1). The electrochemical tests showed that this catalyst shows an excellent electrocatalytic performance. This study provides a new strategy for the development of a high-performance PdCo alloy FAOR catalyst.

2. Materials and Methods

2.1. Materials

MWCNTs (OD: 8 nm, Length: 10–30 mm, Purity ¼ 95 wt%) were purchased from the Xianfeng Nanomaterials Technology Co., Ltd., Nanjing, China. Nafion solution (5 wt%) was purchased from Sigma-Aldrich, Saint Louis, MO, USA. Choline chloride [HOC2H4N(CH3)3Cl], Urea [CO(NH2)2], C2 H5OH, PdCl2, Co(NO3)2, NaBH4, H2SO4, HCOOH and HNO3 were purchased from Shanghai Chemical Reagent Co., Ltd., Shanghai, China. All of the reagents used were analytically pure.

2.2. Materials Synthesis

Concretely, DES (mole ratio of the urea/choline chloride is 2) was stirred at 80 °C until a homogeneous, colourless liquid developed and was sealed at 60 °C, for later use [37]. All solutions are configured with DES instead of water. Then, the untreated 200 mg MWCNTs were functionalized by an acidification treatment (VH2SO4/VHNO3 = 3/1) (referred to CNTs-AO) and sealed at room temperature, for later use. A 10 mg CNTs-AO, 0.33 mL PdCl2 /DESs solution (10 mg/mL), and 0.11 mL Co(NO3)2/DESs solution (10 mg/mL) were added in 20 mL DES with an appropriate proportion (mole ratio of Pd/Co is 3). Then, 60 mg NaBH4 was added and stirred at 60 °C for three hours at room temperature. The product was washed three times with distilled water and anhydrous ethanol, repeatedly. The material was centrifuged in a centrifugal machine, and dried in a vacuum drying oven at 60 °C all night (referred to Pd3Co1/CNTs). In the control experiment, the Pd1Co1/CNT and Pd1Co3/CNT catalysts were prepared following the same steps, as above, except for the change in the amount of Co(NO3)2. Pd/CNTs were prepared following the same steps, as above, except for the addition of Co(NO3)2.

2.3. Materials Characterization

The XRD patterns were obtained using an X-ray diffractometer (Rigaku D/MAX 2500 v/pc, Japan) with a Cu K a radiation source (l ¼ 1.5406 Å). The XPS measurements were carried out using a Physical Electronics PHIQuantum 2000 system with an Al K a radiation source, and all of the XPS spectra were calibrated with the C1s line at 284.5 eV. The surface morphologies and the microstructures of the prepared catalysts were analyzed using (SEM, LEO-1530) and (HRTEM, JEOL JEM-2100) with an accelerating voltage of 200 kV. An Agilent 720 inductively coupled plasma-optical emission spectrophotometer (ICP-OES, USA) was used to determine the Pd contents of Pd/C (20%), Pd/CNTs (17.5%), Pd1Co1/CNTs (18.3%), Pd1Co3/CNTs (18.6%) and Pd3Co1/CNTs (19.1%) (Table S1: Supporting Information).

2.4. Electrochemical Measurements

According to our previously published protocol, the catalyst-modified glassy carbon electrodes (GC, 5 mm in diameter) were prepared [37]. Using a traditional three-electrode system in an electrochemical workstation (CHI 830b05049, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China) at 25 ℃, all of the electrochemical measurements were performed. The counter electrode and the reference electrode were a platinum foil and a saturated calomel electrode (SCE), respectively. The working electrode was modified as follows: 0.05, 0.3, and 1.0 µm alumina powder was used to polish the GC. With the existence of redistilled water (950 µL) and a Nafion solution (0.5 wt%, 50 μL), the catalysts (2 mg) were dispersed in this solution above (1000 μL). A suspension (10 μL) was trickled over the GC electrode at room temperature. The Pd loading of Pd/C, Pd/CNTs, Pd1Co1/CNTs, Pd1Co3/CNTs and Pd3Co1/CNTs were 23.8, 24.2, 24.3, 24.5, and 24.9 μg cm−2.
In a N2-saturated 0.5 M H2SO4 solution, we studied the electrochemical behaviors of all catalysts. The corresponding formic acid electrocatalytic properties of all catalysts were studied in 1.0 M HCOOH + 0.5 M H2SO4 solution. Furthermore, we used CO stripping experiments to investigate the CO toxicity of all catalysts, in terms of the following steps. Firstly, to remove the dissolved O2 in the solution, we degassed the electrolyte by purging with a pure N2 airflow for 15 min. Secondly, to allow the saturated adsorption of CO, we degassed the pure CO airflow bubbling for 15 min at a scan rate of 50 mV s−1 while maintaining the potential sweep in the range of −0.2 and 0.0 V. Then, to remove the dissolved CO and to avoid the disturbance of O2 in the air, we degassed the electrolyte by purging with pure N2 for 30 min. Finally, to oxidize the preadsorbed CO on the surface of the catalysts, we executed the CO stripping voltammograms in the range of −0.2 V and 1.0 V. All of the currents in the electrochemical experiments were expressed by the normalized current per milligram of the Pd metal loading on the working electrode. The electrolyte was purged with pure N2 for 15 min before each measurement, and impeded the disturbance of O2 in the air with a flux of N2 flow over the electrolyte.

3. Results and Discussion

Figure 1 shows the XRD patterns of the Pd3Co1/CNT and Pd/CNT catalysts. The 2θ (26.2°) is attributed to C(002) of MWCNTs-AO [37]. The Pd(111), (200), (220) and (311) diffraction peaks of Pd/CNTs at 39.0°, 44.3°, 66.1° and 79.8°, while all corresponding Pd peaks of Pd3Co1/CNTs shifted to higher 2θ values of 39.2°, 44.6°, 66.3° and 80.1°, suggesting the formation of the Pd3Co1 alloy [18,52,53]. Meanwhile, (Figure S1: Supporting Information) shows the XRD patterns of the Pd1Co1/CNT and Pd1Co3/CNT catalysts. The Pd(111), (200), (220) and (311) diffraction peaks still appear, which indirectly indicates that Co doped into the interior of Pd to form an alloy.
Figure 2 shows the TEM (a), HRTEM (b–c) images and the HAADF-STEM element mapping (d–g) results (Pd, Co) and elements of Pd3Co1/CNTs. Figure 2a,b shows Pd3Co1 nanoclusters successfully loaded on the surface of MWCNTs. As shown in Figure 2c, the spacings of Pd(111) crystal planes in Pd3Co1/CNTs to be 0.206 nm, which is different from the Pd(111)(0.222 nm) of Pd/CNTs (Figure S2: Supporting Information). This may be attributed to the doping of Co atoms with a small radius and compressed in the lattice constant of Pd [22]. Figure 2d-g shows that Pd and Co atoms are evenly distributed in the nanoclusters, which also indirectly proves the formation of alloys. In addition, in order to further determine the formation of the alloy structure, the TEM and HRTEM images and the HAADF-STEM element mapping of Pd3Co1/CNTs were also re-performed in another region, and the results showed that Pd and Co formed an alloy structure (Figure S3: Supporting Information). Figures S4 and S5: Supporting Information show the EDX line-profiles and the spot scanning of the Pd3Co1 nanoparticle (where Pd is in red and Co in blue), which further proved the existence of the alloy and the molar ratio of Pd and Co is (3.07/1). The average nanoparticles’ sizes of Pd3Co1/CNTs (Figure 2a), Pd3Co1/C (Figure S6: Supporting Information), Pd1Co1/CNTs (Figure S7: Supporting Information), Pd1Co1/CNTs (Figure S7: Supporting Information) and Pd/CNTs (Figure S8: Supporting Information) are 4.3, 4.5, 4.9, 5.3 and 7.5 nm (200 particles). Furthermore, in order to explore the carrier’s influence on the catalysts, Figure S9: Supporting Information shows the TEM and HRTEM images of Pd3Co1/C. It can be seen that there is no significant change in the size of the nanoparticles when carbon black is used as the carrier.
Figure 3a shows the XPS survey spectra of Pd3Co1/CNTs. The signals corresponding to C 1s (284.6 eV), O 1s (531.2 eV), Pd 3d (338.2 eV), and Co 2p (783.2 eV) are observed. The two splitting peaks of C1s (C-C, C-O) are attributed to MWCNTs-AO (Figure 3b) [38]. Figure 3c shows the peaks of Co (0) and Co (+2), which are the evidence of Co. Figure 3d and Table S2: Supporting Information show a slight positive change (0.3 eV) in the Pd (0) peak position from Pd/CNTs to Pd3Co1/CNTs, indicating the strong charge transfer interaction between the Co and Pd atoms to form the PdCo bond [54,55], which was the important reason for improving the stability and anti-CO toxicity. Furthermore, the presence of Co (+2) and Pd (+2) is attributed to the oxidation of oxygen in the air [36].
Figure 4a shows the cyclic voltammograms (CV) of the Pd3Co1/CNTs, Pd/CNTs and Pd/C catalysts in the 0.5 M H2SO4 solution. The associated ECSA was calculated by integrating the area associated with the hydrogen adsorption region [56], according to:
ECSA= QH/210 (µC cm−2)/WPd
wherein QH is the integrated charge within the hydrogen adsorption area in the CV curves after subtracting the charge from the double-layer region. It should be noted that the data were obtained from the CV curves with 210 (μC/cm2) as the conversion factor. WPd is the mass of Pd. Therefore, the ECSA of Pd3Co1/CNTs was calculated to be 41.2 m2 g−1, which is higher than those of Pd/CNTs (33.4 m2 g−1), and Pd/C (31.1 m2 g−1). The larger ECSA of Pd3Co1/CNTs is attributed to the smaller Pd3Co1 particles. In order to explore the effect of the atomic ratio on the catalyst performance, we tested the catalytic performance of different Pd/Co catalysts for the FAOR (Figure 4b). In the forward scan, the anodic peak was assigned to the formic acid oxidation, and another anodic peak in the reverse scan was associated with the oxidation of the intermediate carbonaceous species formed during the forward scan [57,58]. For Pd3Co1/CNTs, the peak current density in the forward scan is 2410.1 mA mgPd−1, higher than Pd1Co1/CNTs (1752.3 mA mgPd−1), Pd1Co3/CNTs (1255.1 mA mgPd−1), Pd/CNTs (1262.8 mA mgPd−1) and Pd/C (819.2 mA mgPd−1) (Figure 4b,c). In order to investigate whether the high activity is achieved by the alloying effect or the supporting effect, we tested the catalytic activity of the carbon-supported Pd3Co1 alloy catalyst (Pd3Co1/C) towards the FAOR (Figure S10: Supporting Information). The peak current density of Pd3Co1/C is 1919.2 mA mgPd−1, higher than Pd1Co1/CNTs (1752.3 mA mgPd−1), Pd1Co3/CNTs (1255.1 mA mgPd−1), Pd/CNTs (1262.8 mA mgPd−1) and Pd/C (819.2 mA mgPd−1), but lower than Pd3Co1/CNTs (2410.1 mA mgPd−1), which revealed that the high activity was achieved through the synergistic effect of the alloy effect and the support effect. The alloy effect is beneficial to enhance the toxicity resistance of the catalyst and the addition of MWCNTs is beneficial to improve the electron transport. To evaluate the long-term performance of all catalysts, the chronoamperometric (CA) measurements were performed at +0.3 V for 7200 s. As shown in Figure 4d, in the initial period, all curves showed a fast current decay that indicated the poisoning of the electrocatalysts, due to the formation of intermediate species [26]. Then, after 7200 s, the Pd3Co1/CNT catalyst, still maintaining higher current density (95.1 mA mgPd−1), was almost 2.6 and 29.7 times greater than Pd/CNTs (36.3 mA mgPd−1), and Pd/C (3.2 mA mgPd−1), which illustrated that Pd3Co1/CNTs exhibit a better stability.
We used CO stripping experiments to investigate the CO toxicity of all catalysts. Figure 5 shows the CO stripping voltammograms and the subsequent CV curves. The hydrogen adsorption/desorption were completely suppressed in the low potential region due to the COads species on the active sites of the catalysts [59]. Following the removal of COads, the peaks associated with hydrogen adsorption/desorption reappeared. The onset potential of the adsorbed CO oxidation of Pd3Co1/CNTs was negatively shifted to 0.36 V, and the corresponding potentials were 0.55 and 0.67 V for Pd/CNTs and Pd/C, indicating that Pd3Co1/CNTs have an excellent CO oxidation ability.
In conclusion, compared with Pd/C, Pd/CNTs, Pd1Co1/CNTs and Pd1Co3/CNTs, Pd3Co1/CNTs exhibited 1.38-, 1.92-, 1.91- and 2.94-fold enhancement in mass activity toward the FAOR, respectively. The onset potential of the adsorbed CO oxidation of Pd3Co1/CNTs was negatively shifted to 0.36 V and the corresponding potentials were 0.55, and 0.67 V for Pd/CNTs, and Pd/C. Furthermore, after 7200 s, the Pd3Co1/CNT catalyst still maintained a higher current density (95.1 mA mgPd−1) that was 2.6-, and 29.7- times greater than Pd/CNTs (36.3 mA mgPd−1) and Pd/C (3.2 mA mgPd−1). The excellent catalytic performance and stability of Pd3Co1/CNTs can be attributed to the bi-functional mechanism and the obtained Pd3Co1 alloy nanoclusters. In the bifunctional mechanism, Pd is responsible for the adsorption and oxidative dehydrogenation of formic acid and Co provides the adsorbed hydroxyl group (OHads) at a more negative potential, in comparison with Pd to oxidize the intermediate [22,60]. Moreover, the special Pd3Co1 alloy nanoclusters can provide abundant catalytic active sites, reduce the charge-transfer impedance, improving the efficiency of the material transfer and further improve the catalytic activity and stability. Furthermore, the Pd3Co1/CNT catalyst presents the better FAOR mass activity, in comparison with the recent research works on Pd-based bimetallic catalysts [22,26,57,61,62,63,64,65,66,67] (Table S3: Supporting Information).

4. Conclusions

Herein, with the assistance of DES, the obtained Pd3Co1/CNT catalyst exhibited the excellent electrocatalytic performance towards the FAOR. The doping of Co atoms changed the electron configuration of Pd to form a new PdCo bond, thus affecting the adsorption of the toxic intermediates and weakening the dissolution of Pd, which were the important reasons for improving the stability and anti-CO toxicity. Furthermore, these doped Co atoms provided more co-catalytic active sites to obtain more OH from the H2O molecule, which could further interact with the toxic intermediates generated on the Pd active sites to further improve performance. This reaserch provides a new strategy to obtain the Pd-alloy high-performance catalysts for DFAFC.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12234182/s1, Figure S1. XRD patterns of Pd1Co1/CNT (a) and Pd1Co3/CNTs(b) catalysts; Figure S2. TEM and HRTEM images of the Pd/CNT catalyst; Figure S3: TEM and HRTEM images; HAADF-STEM elements mapping; the corresponding elements of Pd and Co of Pd3Co1/CNTs (another region); Figure S4. EDX line-profiles (a,b), spot scanning (c,d) of a Pd3Co1 nanoparticle (where Pd is in red and Co in blue) of Pd3Co1/CNTs. Figure S5. EDX-spot scanning and element content ratio (b,c) of a Pd3Co1 nanoparticle in Pd3Co1/CNTs. Figure S6. The corresponding particle size distribution of a Pd3Co1/C catalyst. Figure S7. TEM and HRTEM images and the corresponding particle size distribution of Pd1Co1/CNT (a–c) and Pd1Co3/CNT (d–f) catalysts. Figure S8. The corresponding particle size distribution of Pd/CNTs. Figure S9. TEM and HRTEM images of a Pd3Co1/C catalyst. Figure S10 Cyclic voltammograms curve of Pd3Co1/C in 0.5 M H2SO4 + 1.0 M HCOOH. Table S1: Elemental composition of the samples obtained from ICP; Table S2: Pd 3d peaks of Pd3Co1/CNTs and Pd/CNTs; Table S3: A recent literature survey of the activity of the FAOR electrocatalysts.

Author Contributions

Conceptualization, P.Y.; methodology, L.Z.; software, X.W.; validation, X.W. and S.D.; formal analysis, P.Y.; investigation, P.Y.; resources, P.Y.; data curation, L.Z.; writing—original draft preparation, P.Y.; writing—review and editing, P.Y.; visualization, P.Y.; supervision, Y.O.; project administration, Y.O.; funding acquisition, Y.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Hunan Province (2020JJ4073), Hunan Provincial general project of Education Department (21C0645). Key project of Huaihua University (HHUY2021-03).

Data Availability Statement

Data are available upon request; please contact Pingping Yang, [email protected].

Acknowledgments

The authors thank the the Natural Science Foundation of Hunan Province (2020JJ4073), Hunan Provincial general project of Education Department (21C0645). Key project of Huaihua University (HHUY2021-03).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, Z.; Chen, Y.; Ji, S.; Tang, Y.; Chen, W.; Li, A.; Zhao, J.; Xiong, Y.; Wu, Y.; Gong, Y.; et al. Iridium Single-Atom Catalyst on Nitrogen-Doped Carbon for Formic Acid Oxidation Synthesized Using a General Host-Guest Strategy. Nat. Chem. 2020, 12, 764–772. [Google Scholar] [CrossRef] [PubMed]
  2. Lv, F.; Huang, B.; Feng, J.; Zhang, W.; Wang, K.; Li, N.; Zhou, J.; Zhou, P.; Yang, W.; Du, Y.; et al. A Highly Efficient Atomically Thin Curved PdIr Bimetallene Electrocatalyst. Natl. Sci. Rev. 2021, 8, nwab019. [Google Scholar] [CrossRef]
  3. Ding, J.; Liu, Z.; Liu, X.; Liu, B.; Liu, J.; Deng, Y.; Han, X.; Hu, W.; Zhong, C. Tunable Periodically Ordered Mesoporosity in Palladium Membranes Enables Exceptional Enhancement of Intrinsic Electrocatalytic Activity for Formic Acid Oxidation. Angew. Chem. 2020, 132, 5130–5139. [Google Scholar] [CrossRef]
  4. Shi, Y.; Schimmenti, R.; Zhu, S.; Venkatraman, K.; Chen, R.; Chi, M.; Shao, M.; Mavrikakis, M.; Xia, Y. Solution-Phase Synthesis of PdH0.706 Nanocubes with Enhanced Stability and Activity toward Formic Acid Oxidation. J. Am. Chem. Soc. 2022, 144, 2556–2568. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, P.; Yang, X.; Liu, W.; Guo, R.; Yao, Z. Graphene-Based Electrocatalysts for Advanced Energy Conversion. Green Energy Environ. 2022. [Google Scholar] [CrossRef]
  6. Xi, Z.; Li, J.; Su, D.; Muzzio, M.; Yu, C.; Li, Q.; Sun, S. Stabilizing CuPd Nanoparticles via CuPd Coupling to WO2.72 Nanorods in Electrochemical Oxidation of Formic Acid. J. Am. Chem. Soc. 2017, 139, 15191–15196. [Google Scholar] [CrossRef]
  7. Li, H.; Zhang, Y.; Wan, Q.; Li, Y.; Yang, N. Expanded Graphite and Carbon Nanotube Supported Palladium Nanoparticles for Electrocatalytic Oxidation of Liquid Fuels. Carbon 2018, 131, 111–119. [Google Scholar] [CrossRef]
  8. Zhang, L.Y.; Wang, F.; Wang, S.; Huang, H.; Meng, X.; Ouyang, Y.; Yuan, W.; Guo, C.X.; Li, C.M. 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]
  9. Li, Y.H.; Deng, H.C.; Zhou, Z.H.; Yang, P.P.; Fei, J.J.; Xie, Y.X. Pd12Ag1 Nanoalloy on Dendritic CNFs Catalyst for Boosting Formic Acid Oxidation. Appl. Surf. Sci. 2023, 608, 155131. [Google Scholar] [CrossRef]
  10. Qian, H.; Wu, J.; Guo, Y.; Fang, W. PdAgPt Corner-Satellite Nanocrystals in Well-Controlled Morphologies and the Structure-Related Electrocatalytic Properties. Nanomaterials 2021, 11, 340. [Google Scholar] [CrossRef] [PubMed]
  11. Liu, X.; Li, Z.; Wang, K.; Zhou, L.; Zhao, X.; Jiang, W.; Li, Q.; Deng, Y. Facile Synthesis of Pd Nanocubes with Assistant of Iodide and Investigation of Their Electrocatalytic Performances Towards Formic Acid Oxidation. Nanomaterials 2019, 9, 375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Zheng, F.; Kwong, T.L.; Yung, K.F. Surfactant-Free Monodispersed Pd Nanoparticles Template for Core-Shell Pd@PdPt Nanoparticles as Electrocatalyst towards Methanol Oxidation Reaction (MOR). Nanomaterials 2022, 12, 260. [Google Scholar] [CrossRef]
  13. Kim, J.; Jang, J.S.; Peck, D.H.; Lee, B.; Yoon, S.H.; Jung, D.H. Methanol-Tolerant Platinum-Palladium Catalyst Supported on Nitrogen-Doped Carbon Nanofiber for High Concentration Direct Methanol Fuel Cells. Nanomaterials 2016, 6, 148. [Google Scholar] [CrossRef] [Green Version]
  14. Su, S.; Shi, Y.; Zhou, Y.; Wang, Y.B.; Wang, F.B.; Xia, X.H. Tailoring the Electron Density of Pd Nanoparticles through Electronic Metal-Support Interaction for Accelerating Electrocatalysis of Formic Acid. Electrochem. Commun. 2019, 107, 106540. [Google Scholar] [CrossRef]
  15. Li, X.; Liu, Y.; Zhang, J.J.; Yan, B.; Jin, C.; Dou, J.; Li, M.; Feng, X.; Liu, G. No Annealing Synthesis of Ordered Intermetallic PdCu Nanocatalysts for Boosting Formic Acid Oxidation. Chem. Mater. 2022, 34, 1385–1391. [Google Scholar] [CrossRef]
  16. Zheng, L.; Zeng, T.; Ling, M.; Wang, Y.; Yuan, W.; Zhang, L.Y. Electrocatalytic Activity Enhancement of Palladium-Manganese Nanosheet Assembled Nanobuds by Tuning Electronic Structure. Appl. Surf. Sci. 2022, 605, 154634. [Google Scholar] [CrossRef]
  17. Zhang, L.Y.; Meng, X.; Zhang, W.; Zeng, T.; Yuan, W.; Zhao, Z. Synthesis of Palladium–Tungsten Metallene-Constructed Sandwich-Like Nanosheets as Bifunctional Catalysts for Direct Formic Acid Fuel Cells. ACS Appl. Energy Mater. 2021, 4, 12336–12344. [Google Scholar] [CrossRef]
  18. Juárez-Marmolejo, L.; Pérez-Rodríguez, S.; de Oca-Yemha, M.M.; Palomar-Pardavé, M.; Romero-Romo, M.; Ezeta-Mejía, A.; Morales-Gil, P.; Martínez-Huerta, M.V.; Lázaro, M.J. Carbon Supported PdM (M = Fe, Co) Electrocatalysts for Formic Acid Oxidation. Influence of the Fe and Co Precursors. Int. J. Hydrogen Energy 2019, 44, 1640–1649. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, L.Y.; Liu, Z. Graphene Decorated with Pd4Ir Nanocrystals: Ultrasound-Assisted Synthesis, and Application as a Catalyst for Oxidation of Formic Acid. J. Colloid Interf. Sci. 2017, 505, 783–788. [Google Scholar] [CrossRef]
  20. Wang, R.X.; Fan, Y.J.; Liang, Z.R.; Zhang, J.M.; Zhou, Z.Y.; Sun, S.G. PdSn Nanocatalysts Supported on Carbon Nanotubes Synthesized in Deep Eutectic Solvents with High Activity for Formic Acid Electrooxidation. RSC Adv. 2016, 6, 60400–60406. [Google Scholar] [CrossRef]
  21. Bao, Y.; Zha, M.; Sun, P.; Hu, G.; Feng, L. PdNi/N-Doped Graphene Aerogel with Over Wide Potential Activity for Formic Acid Electrooxidation. J. Energy Chem. 2021, 59, 748–754. [Google Scholar] [CrossRef]
  22. Zhang, L.Y.; Gong, Y.; Wu, D.; Li, Z.; Li, Q.; Zheng, L.; Chen, W. Palladium-Cobalt Nanodots Anchored on Graphene: In-Situ Synthesis, and Application as an Anode Catalyst for Direct Formic Acid Fuel Cells. Appl. Surf. Sci. 2019, 469, 305–311. [Google Scholar] [CrossRef]
  23. Li, X.; Peng, X.; Wang, Y.; Xi, B.; Dou, J.; Zhang, J.J.; Liu, Y.; Jin, C. Iron- and Cobalt-Doped Palladium/Carbon Nanoparticles as Catalysts for Formic Acid Oxidation. ACS Appl. Nano Mater. 2022, 5, 12407–12412. [Google Scholar] [CrossRef]
  24. Zhang, L.Y.; Gong, Y.; Liu, H.; Yuan, W.; Liu, Z. Ultrasmall and Uniform Pt3Au Clusters Strongly Suppress Ostwald Ripening for Efficient Ethanol Oxidation. Electrochem. Commun. 2017, 84, 1–5. [Google Scholar] [CrossRef]
  25. Yan, B.; Wang, C.; Xu, H.; Zhang, K.; Li, S.; Du, Y. Facile Synthesis of a Porous Pd/Cu Alloy and Its Enhanced Performance toward Methanol and Formic Acid Electrooxidation. ChemPluschem 2017, 82, 1121–1128. [Google Scholar] [CrossRef]
  26. Zhang, L.Y.; Gong, Y.; Wu, D.; Wu, G.; Xu, B.; Bi, L.; Yuan, W.; Cui, Z. Twisted Palladium-Copper Nanochains toward Efficient Electrocatalytic Oxidation of Formic Acid. J. Colloid Interf. Sci. 2019, 537, 366–374. [Google Scholar] [CrossRef]
  27. Wang, F.; Wang, S.; Wu, D.; Huang, H.; Yuan, W.; Zhang, L.Y. Layered PdW Nanosheet Assemblies for Alcohol Electrooxidation. Appl. Surf. Sci. 2021, 537, 147860. [Google Scholar] [CrossRef]
  28. Selepe, C.T.; Gwebu, S.S.; Matthews, T.; Mashola, T.A.; Sikeyi, L.L.; Zikhali, M.; Maxakato, N.W. Effect of Sn Doping on Pd Electro-Catalysts for Enhanced Electro-Catalytic Activity towards Methanol and Ethanol Electro-Oxidation in Direct Alcohol Fuel Cells. Nanomaterials 2021, 11, 2725. [Google Scholar] [CrossRef]
  29. Ma, Y.; Li, T.; Chen, H.; Chen, X.; Deng, S.; Xu, L.; Sun, D.; Tang, Y. A General Strategy to the Synthesis of Carbon-Supported PdM (M = Co, Fe and Ni) Nanodendrites as High-Performance Electrocatalysts for Formic Acid Oxidation. J. Energy Chem. 2017, 26, 1238–1244. [Google Scholar] [CrossRef] [Green Version]
  30. Zeng, T.; Meng, X.; Huang, H.; Zheng, L.; Chen, H.; Zhang, Y.; Yuan, W.; Zhang, L.Y. Controllable Synthesis of Web-Footed PdCu Nanosheets and Their Electrocatalytic Applications. Small 2022, 18, 2107623. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, W.; He, T.; Yang, X.; Liu, Y.; Wang, C.; Li, J.; Xiao, A.; Zhang, K.; Shi, X.; Jin, M. General Synthesis of Amorphous PdM (M = Cu, Fe, Co, Ni) Alloy Nanowires for Boosting HCOOH Dehydrogenation. Nano Lett. 2021, 21, 3458–3464. [Google Scholar] [CrossRef]
  32. Zhang, X.J.; Zhang, J.M.; Zhang, P.Y.; Li, Y.; Xiang, S.; Tang, H.G.; Fan, Y.J. Highly Active Carbon Nanotube-Supported Ru@Pd Core-Shell Nanostructure as an Efficient Electrocatalyst toward Ethanol and Formic Acid Oxidation. Mol. Catal. 2017, 436, 138–144. [Google Scholar] [CrossRef]
  33. Zhong, J.; Li, L.; Waqas, M.; Wang, X.; Fan, Y.; Qi, J.; Yang, B.; Rong, C.; Chen, W.; Sun, S. Deep Eutectic Solvent-Assisted Synthesis of Highly Efficient PtCu Alloy Nanoclusters on Carbon Nanotubes for Methanol Oxidation Reaction. Electrochim Acta 2019, 322, 134677. [Google Scholar] [CrossRef]
  34. Zhang, J.M.; Sun, S.N.; Li, Y.; Zhang, X.J.; Zhang, P.Y.; Fan, Y.J. A Strategy in Deep Eutectic Solvents for Carbon Nanotube-Supported PtCo Nanocatalysts with Enhanced Performance toward Methanol Electrooxidation. Int. J. Hydrogen Energy 2017, 42, 26744–26751. [Google Scholar] [CrossRef]
  35. Wagle, D.V.; Zhao, H.; Baker, G.A. Deep Eutectic Solvents: Sustainable Media for Nanoscale and Functional Materials. Acc. Chem. Res. 2014, 47, 2299–2308. [Google Scholar] [CrossRef] [PubMed]
  36. Yang, P.; Devasenathipathy, R.; Xu, W.; Wang, Z.; Chen, D.H.; Zhang, X.; Fan, Y.; Chen, W. Pt1(CeO2)0.5 Nanoparticles Supported on Multiwalled Carbon Nanotubes for Methanol Electro-Oxidation. ACS Appl. Nano Mater. 2021, 4, 10584–10591. [Google Scholar] [CrossRef]
  37. Yang, P.P.; Zhao, P.C.; Luo, N.; Li, Y.H.; Wang, C.; Zhang, L.; Xie, Y.X.; Fei, J.J. A “Special” Anhydrous System for the Preparation of Alloyed Pd1Ce0.5 Nanonetworks Catalyst Supported on Carbon Nanotubes with High Electrochemical Oxidation Activity for Formic Acid. Int. J. Hydrogen Energy 2021, 46, 18857–18865. [Google Scholar] [CrossRef]
  38. Yang, P.; Zhou, Z.; Zheng, T.; Gu, C.; Gong, X.; Zhang, Y.; Xie, Y.; Yang, N.; Fei, J. A Novel Strategy to Synthesize Pt/CNTs Nanocatalyst with Highly Improved Activity for Methanol Electrooxidation. J. Electroanal. Chem. 2021, 897, 115557. [Google Scholar] [CrossRef]
  39. Yang, P.; Wei, L.; Xiao, X.; Zhou, Z.; Li, J.; Zhang, Y.; Xie, Y.; Yang, N.; Fei, J. Electrocatalytic Oxidation of Formic Acid on Pd/CNTs Nanocatalysts Synthesized in Special “Non-Aqueous” System. J. Electroanal. Chem. 2022, 906, 115980. [Google Scholar] [CrossRef]
  40. Yang, P.; Li, Y.; Chen, S.; Li, J.; Zhao, P.; Zhang, L.; Xie, Y.; Fei, J. One-Step Synthesis in Deep Eutectic Solvents of Pt3Sn1-SnO2 Alloy Nanopore on Carbon Nanotubes for Boosting Electro-Catalytic Methanol Oxidation. J. Electroanal. Chem. 2021, 887, 115164. [Google Scholar] [CrossRef]
  41. Zhang, J.M.; He, J.J.; Wang, X.Q.; Fan, Y.J.; Zhang, X.J.; Zhong, J.P.; Chen, W.; Sun, S.G. One-Step Synthesis in Deep Eutectic Solvents of PtV Alloy Nanonetworks on Carbon Nanotubes with Enhanced Methanol Electrooxidation Performance. Int. J. Hydrogen Energy 2019, 44, 28709–28719. [Google Scholar] [CrossRef]
  42. Xiang, S.; Wang, L.; Huang, C.C.; Fan, Y.J.; Tang, H.G.; Wei, L.; Sun, S.G. Concave Cubic PtLa Alloy Nanocrystals with High-Index Facets: Controllable Synthesis in Deep Eutectic Solvents and Their Superior Electrocatalytic Properties for Ethanol Oxidation. J. Power Sources 2018, 399, 422–428. [Google Scholar] [CrossRef]
  43. Azevedo, R.S.; de Sousa, J.R.; Araujo, M.T.; Martins Filho, A.J.; de Alcantara, B.N.; Araujo, F.; Queiroz, M.G.; Cruz, A.C.; Vasconcelos, B.H.B.; Chiang, J.O.; et al. In Situ Immune Response and Mechanisms of Cell Damage in Central Nervous System of Fatal Cases Microcephaly by Zika Virus. Sci. Rep. 2018, 8, 1. [Google Scholar] [CrossRef] [Green Version]
  44. Wei, L.; Mao, Y.J.; Liu, F.; Sheng, T.; Wei, Y.S.; Li, J.W.; Fan, Y.J.; Zhao, X.S. Concave Cubic Pt–Sm Alloy Nanocrystals with High-Index Facets and Enhanced Electrocatalytic Ethanol Oxidation. ACS Appl. Energy Mater. 2019, 2, 7204–7210. [Google Scholar] [CrossRef]
  45. Wei, L.; Xu, C.-D.; Huang, L.; Zhou, Z.-Y.; Chen, S.-P.; Sun, S.-G. Electrochemically Shape-Controlled Synthesis of Pd Concave-Disdyakis Triacontahedra in Deep Eutectic Solvent. J. Phys. Chem. 2015, 120, 15569–15577. [Google Scholar] [CrossRef]
  46. Mao, Y.J.; Wei, L.; Zhao, X.S.; Wei, Y.S.; Li, J.W.; Sheng, T.; Zhu, F.C.; Tian, N.; Zhou, Z.Y.; Sun, S.G. Excavated Cubic Platinum-Iridium Alloy Nanocrystals with High-Index Facets as Highly Efficient Electrocatalysts in N2 Fixation to NH3. Chem. Commun. 2019, 55, 9335–9338. [Google Scholar] [CrossRef]
  47. Hammons, J.A.; Muselle, T.; Ustarroz, J.; Tzedaki, M.; Raes, M.; Hubin, A.; Terryn, H. Stability, Assembly, and Particle/Solvent Interactions of Pd Nanoparticles Electrodeposited from a Deep Eutectic Solvent. J. Phys. Chem. 2013, 117, 14381–14389. [Google Scholar] [CrossRef]
  48. Luo, S.-X.L.; Liu, R.Y.; Lee, S.; Swager, T.M. Electrocatalytic Isoxazoline-Nanocarbon Metal Complexes. J. Am. Chem. Soc. 2021, 143, 10441–10453. [Google Scholar] [CrossRef]
  49. Qi, Z.; Xiao, C.; Liu, C.; Goh, T.W.; Zhou, L.; Maligal-Ganesh, R.; Pei, Y.; Li, X.; Curtiss, L.A.; Huang, W. Sub-4 Nm PtZn Intermetallic Nanoparticles for Enhanced Mass and Specific Activities in Catalytic Electrooxidation Reaction. J. Am. Chem. Soc. 2017, 139, 4762–4768. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, L.; Li, J.; Wang, C.; Tang, J.; Chen, X.; Li, Y.; Shi, J.; Zhao, P.; Xie, Y.; Fei, J. A Novel Kaempferol Electrochemical Sensor Based on Glass Carbon Electrode Modified by Poly (3, 4-Ethylenedioxythiophene) Decorated with Green Synthesized MIL-100(Fe)-Multi-Walled Carbon Nanotubes Composites. Colloids Surf. A 2022, 649, 129484. [Google Scholar] [CrossRef]
  51. Hu, Y.; Wang, C.; Zhao, P.; Zhang, L.; Fei, J.; Xie, Y. A Novel Catechin Electrochemical Sensor Based on a Two-Dimensional MOFs Material Derivative Zn Doped Carbon Nanosheets and Multi-Walled Carbon Nanotubes Composite Film. Talanta 2022, 246, 123520. [Google Scholar] [CrossRef] [PubMed]
  52. Cazares-Avila, E.; Ruiz-Ruiz, E.J.; Hernández-Ramírez, A.; Rodríguez-Varela, F.J.; Morales-Acosta, M.D.; Morales-Acosta, D. Effect of OMC and MWNTC Support on Mass Activity of PdCo Catalyst for Formic Acid Electro-Oxidation. Int. J. Hydrogen Energy 2017, 42, 30349–30358. [Google Scholar] [CrossRef]
  53. Gunji, T.; Wakabayashi, R.H.; Noh, S.H.; Han, B.; Matsumoto, F.; DiSalvo, F.J.; Abruña, H.D. The Effect of Alloying of Transition Metals (M = Fe, Co, Ni) with Palladium Catalysts on the Electrocatalytic Activity for the Oxygen Reduction Reaction in Alkaline Media. Electrochim. Acta 2018, 283, 1045–1052. [Google Scholar] [CrossRef]
  54. Zuluaga, S.; Stolbov, S. Factors Controlling the Energetics of the Oxygen Reduction Reaction on the Pd-Co Electro-Catalysts: Insight from First Principles. J. Chem. Phys. 2011, 135, 134702. [Google Scholar] [CrossRef]
  55. Suo, Y.; Zhuang, L.; Lu, J. First-Principles Considerations in the Design of Pd-Alloy Catalysts for Oxygen Reduction. Angew. Chem. Int. Edit. 2007, 46, 2862–2864. [Google Scholar] [CrossRef]
  56. Yang, B.; Zhang, W.; Hu, S.; Liu, C.; Wang, X.; Fan, Y.; Jiang, Z.; Yang, J.; Chen, W. Bidirectional Controlling Synthesis of Branched PdCu Nanoalloys for Efficient and Robust Formic Acid Oxidation Electrocatalysis. J. Colloid Interf. Sci. 2021, 600, 503–512. [Google Scholar] [CrossRef]
  57. Zhang, Z.; Gong, Y.; Wu, D.; Li, Z.; Li, Q.; Zheng, L.; Chen, W.; Yuan, W.; Zhang, L.Y. Facile Fabrication of Stable PdCu Clusters Uniformly Decorated on Graphene as an Efficient Electrocatalyst for Formic Acid Oxidation. Int. J. Hydrogen Energy 2019, 44, 2731–2740. [Google Scholar] [CrossRef]
  58. Zhang, L.Y.; Meng, X.; Wu, H.; Wang, F.; Huang, H.; Ouyang, Y.; Yuan, W.; Guo, C.X.; Li, C.M. Tungsten-Induced Synthesis of Defective Palladium–Copper–Tungsten Trimetallic Nanochains to Highly Enhance Activity for Formic acid Electrooxidation. Mater. Today Energy 2020, 18, 100558. [Google Scholar] [CrossRef]
  59. Chen, Y.; Yang, Y.; Fu, G.; Xu, L.; Sun, D.; Lee, J.M.; Tang, Y. Core–Shell CuPd@Pd Tetrahedra with Concave Structures and Pd-Enriched Surface Boost Formic Acid Oxidation. J. Mater. Chem. A 2018, 6, 10632–10638. [Google Scholar] [CrossRef]
  60. Gong, Y.; Liu, X.; Gong, Y.; Wu, D.; Xu, B.; Bi, L.; Zhang, L.Y.; Zhao, X. Synthesis of Defect-Rich Palladium-Tin Alloy Nanochain Networks for Formic Acid Oxidation. J. Colloid Interf. Sci. 2018, 530, 189–195. [Google Scholar] [CrossRef]
  61. Zheng, J.; Zeng, H.; Tan, C.; Zhang, T.; Zhao, B.; Guo, W.; Wang, H.; Sun, Y.; Jiang, L. Coral-like PdCu Alloy Nanoparticles Act as Stable Electrocatalysts for Highly Efficient Formic Acid Oxidation. ACS Sustain. Chem. Eng. 2019, 7, 15354–15360. [Google Scholar] [CrossRef]
  62. Zhang, L.Y.; Zhao, Z.L.; Yuan, W.; Li, C. M Facile one-pot surfactant-free synthesis of uniform Pd6Co nanocrystals on 3D graphene as an efficient electrocatalyst toward formic acid oxidation. Nanoscale 2016, 8, 1905–1909. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, L.; Zhai, J.J.; Jiang, K.; Wang, J.Q.; Cai, W.B. Pd–Cu/C electrocatalysts synthesized by one-pot polyol reduction toward formic acid oxidation: Structural characterization and electrocatalytic performance. Int. J. Hydrogen Energy 2015, 40, 1726–1734. [Google Scholar] [CrossRef]
  64. Sun, D.; Si, L.; Fu, G.; Liu, C.; Sun, D.; Chen, Y.; Tang, Y.; Lu, T. Nanobranched porous palladium–tin intermetallics: One-step synthesis and their superior electrocatalysis towards formic acid oxidation. J. Power Sources 2015, 280, 141–146. [Google Scholar] [CrossRef]
  65. 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]
  66. Feng, A.; Bai, J.; Shao, W.; Hong, W.; Tian, Z.Q.; Xiao, Z. Surfactant-free Pd–Fe nanoparticles supported on reduced graphene oxide as nanocatalyst for formic acid oxidation. Int. J. Hydrogen Energy 2017, 42, 15196–15202. [Google Scholar] [CrossRef]
  67. Zhu, F.; Ma, G.; Bai, Z.; Hang, R.; Tang, B.; Zhang, Z.; Wang, X. High activity of carbon nanotubes supported binary and ternary Pd-based catalysts for methanol, ethanol and formic acid electro-oxidation. J. Power Sources 2013, 242, 610–620. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration showing the preparation of Pd3Co1/CNTs.
Scheme 1. Schematic illustration showing the preparation of Pd3Co1/CNTs.
Nanomaterials 12 04182 sch001
Figure 1. XRD patterns of the Pd/CNT and Pd3Co1/CNT catalysts.
Figure 1. XRD patterns of the Pd/CNT and Pd3Co1/CNT catalysts.
Nanomaterials 12 04182 g001
Figure 2. TEM and HRTEM images (the corresponding particle size distribution); HAADF-STEM elements.
Figure 2. TEM and HRTEM images (the corresponding particle size distribution); HAADF-STEM elements.
Nanomaterials 12 04182 g002
Figure 3. (a) XPS survey spectra, (b) C (1s) spectrum, (c) Co (2p) spectrum of Pd3Co1/CNTs, (d) Pd (3d) spectrum of Pd3Co1/CNTs and Pd/CNTs.
Figure 3. (a) XPS survey spectra, (b) C (1s) spectrum, (c) Co (2p) spectrum of Pd3Co1/CNTs, (d) Pd (3d) spectrum of Pd3Co1/CNTs and Pd/CNTs.
Nanomaterials 12 04182 g003
Figure 4. (a,c) Cyclic voltammograms and (d) current-time curves of Pd3Co1/CNTs, Pd/CNTs and Pd/C; (b) cyclic voltammograms curves of Pd3Co1/CNTs, Pd1Co1/CNTs and Pd1Co3/CNTs in 0.5 M H2SO4/1.0 M HCOOH + 0.5 M H2SO4 solution.
Figure 4. (a,c) Cyclic voltammograms and (d) current-time curves of Pd3Co1/CNTs, Pd/CNTs and Pd/C; (b) cyclic voltammograms curves of Pd3Co1/CNTs, Pd1Co1/CNTs and Pd1Co3/CNTs in 0.5 M H2SO4/1.0 M HCOOH + 0.5 M H2SO4 solution.
Nanomaterials 12 04182 g004
Figure 5. CO stripping voltammograms of Pd/C, Pd/CNT and Pd3Co1/CNT catalysts in 0.5 M H2SO4 solution.
Figure 5. CO stripping voltammograms of Pd/C, Pd/CNT and Pd3Co1/CNT catalysts in 0.5 M H2SO4 solution.
Nanomaterials 12 04182 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, P.; Zhang, L.; Wei, X.; Dong, S.; Ouyang, Y. Pd3Co1 Alloy Nanocluster on the MWCNT Catalyst for Efficient Formic Acid Electro-Oxidation. Nanomaterials 2022, 12, 4182. https://doi.org/10.3390/nano12234182

AMA Style

Yang P, Zhang L, Wei X, Dong S, Ouyang Y. Pd3Co1 Alloy Nanocluster on the MWCNT Catalyst for Efficient Formic Acid Electro-Oxidation. Nanomaterials. 2022; 12(23):4182. https://doi.org/10.3390/nano12234182

Chicago/Turabian Style

Yang, Pingping, Li Zhang, Xuejiao Wei, Shiming Dong, and Yuejun Ouyang. 2022. "Pd3Co1 Alloy Nanocluster on the MWCNT Catalyst for Efficient Formic Acid Electro-Oxidation" Nanomaterials 12, no. 23: 4182. https://doi.org/10.3390/nano12234182

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