1. State Key Laboratory of Multiphase Flow in Power Engineering, Xi’an Jiaotong University, Xi’an 710049, China
2. Institute of Chemical Engineering and Technology, Xi’an Jiaotong University, Xi’an 712000, China
hujiang@mail.xjtu.edu.cn
bohuang@mail.xjtu.edu.cn
Show less
History+
Received
Accepted
Published
2023-05-23
2023-07-06
2023-12-15
Issue Date
Revised Date
2023-09-15
PDF
(1907KB)
Abstract
Dehydrogenation of formic acid (FA) is considered to be an effective solution for efficient storage and transport of hydrogen. For decades, highly effective catalysts for this purpose have been widely investigated, but numerous challenges remain. Herein, the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloys over the whole composition range were successfully prepared and used to catalyze FA hydrogen production efficiently near room temperature. Small PdAu nanoparticles (5–10 nm) were well-dispersed and supported on the activated carbon to form PdAu solid solution alloys via the eco-friendly slow synthesis methodology. The physicochemical properties of the PdAu alloys were comprehensively studied by utilizing various measurement methods, such as X-ray diffraction (XRD), N2 adsorption–desorption, high angle annular dark field-scanning transmission electron microscope (HAADF-STEM), X-ray photoelectrons spectroscopy (XPS). Notably, owing to the strong metal-support interaction (SMSI) and electron transfer between active metal Au and Pd, the Pd0.5Au0.5 obtained exhibits a turnover frequency (TOF) value of up to 1648 h−1 (313 K, nPd+Au/nFA = 0.01, nHCOOH/nHCOONa = 1:3) with a high activity, selectivity, and reusability in the FA dehydrogenation.
Since the 21st century, the world faces two major challenges, energy crisis and environment deterioration. Finding and exploring safe, clean, and renewable energies has become a hot research topic. As an ideal carbon-zero fuel, hydrogen is viewed as the most promising alternatives for changing the traditional fossil energy system because of a variety of advantages (high-efficiency, recyclability, cleanness, etc.) [1,2]. Particularly, H2 has been widely used in proton exchange membrane fuel cells because its product is only water [3]. Chemical hydrogen storage materials are widely preferred due to their high weight and volumetric hydrogen density compared to classical pressurization and cryogenic liquefaction technologies [4,5]. For decades, the chemical H2 storage materials: formic acid (FA, HCOOH) [6,7], methanol (CH3OH) [8,9], sodium borohydride (NaBH4) [10,11], hydrazine (N2H4) [12,13], hydrous hydrazine (N2H4·H2O) [14–16], and hydrazine borane (NH3BH3) [17,18] have been comprehensively investigated and developed for transportation and generation of hydrogen [19,20].
As one of the common liquid biomass products, formic acid (FA) possesses a relatively high hydrogen storage content, density, and volume. Therefore, it has received widespread attention in the application of portable hydrogen storage devices because of its low price, chemical stability, and easy storage and transportation [7,21,22]. Moreover, it is easily charged and discharged through sharing current infrastructures of liquid fuels, which greatly promotes its production and application. As the reverse reaction of CO2 hydrogenation [23,24], it can be decomposed into H2 and CO2 by dehydrogenation (HCOOH → CO2 + H2, ΔG298 = −48.4 kJ/mol) or by dehydration (HCOOH → CO + H2O, ΔG298 = −28.5 kJ/mol), but the last is undesirable because CO can poison the catalysts in fuel cells [25,26]. Therefore, rapid development of highly active catalysts accompanied with high selectivity and efficient H2 generation is of paramount for FA-based hydrogen storage.
Many works have been made to investigate the homogeneous and heterogeneous catalysts in FA hydrogenation. For homogeneous catalysts, Schneider and its coworkers [27] reported that (iPrPNHP)Fe(CO)H-(COO) and Lewis acid cocatalyst obtain an particularly high turnover frequency (TOF) value of 196728 h−1 at 353 K. However, homogeneous catalysts still face many problems such as special additives, separation, and recovery, which further limit the practical application of catalysts. To date, a large number of researchers have explored the heterogeneous catalysts for FA dehydrogenation. As the most competitive catalyst for FA hydrogenation, Pd-based catalysts have high activities and median adsorption energy values compared to other metal-based catalysts [4,28–30]. But Pd-based single metal catalyst exhibit an acceptable H2 generation rate at a relatively high temperature (333–433 K) accompany with the addition of sodium and potassium salts. Additionally, the production of CO byproducts could cause the inactivation of mono-metal Pd-based catalyst [29,31–37]. Therefore, many Pd-based binary and ternary metal (PdAu [38–44], PdAg [45–47], PdCu [5], PdCo [48], PdAuCo [49], PdAgCo [50] and PdAu-MnOx [51,52]) catalysts exhibit a higher reactivity as well as reusability compared with Pd mono-metal catalysts. Although Au alone is inactive for FA decomposition, PdAu [53,54] bimetal catalysts could greatly enhance the catalytic performance due to their synergistic effect. The charge transfer between Au and Pd not only plays an important role in regulating the chemical environment nearby the active center Pd, but also inhibits the generation of CO and prevents catalysts poisoning [55,56]. Moreover, many significant works are employed to increase the structure-reactivity relationship of PdAu bimetal catalysts including alloy structures [38,57], core shell structures [58], and contracture strong metal-support interaction (SMSI). For the abovementioned several structures, the uniformly mixing PdAu bimetals with smaller particle sizes poses a huge challenge [59,60], directly affecting the FA dehydrogenation activity.
Recently, new ultra-small solid-solution alloys have been synthesized, which aroused lots of attention due to their many innovative properties [61,62]. The greatest advantage of solid solution alloys is the control of their electronic structure at the atomic level in the bulk state and the influence of most physicochemical properties via altering the metals compositions in the solution [63]. By completely mixing of metals A and B, the AB alloy possesses not only the properties of A or B, but also a unique property because of the synergistic capacity of A and B. Although many immiscible alloys including PdAu [64], CuRu [65], RhNi [66], AuRu [63], etc. have been prepared, the synthesis of solid solution alloys nanoparticles (NPs) with large differences in reduction potential remains challenging. For noble metal alloy NPs, the most critical control step is the co-reduction rate of double metals, otherwise core shell or isolated single metal NPs can be obtained [67,68]. Hence, it is challenging to co-reduce bimetal ions with a large gap in reduction potential energy. The reduction potentials are shown in Table S1 in Electronic Supplementary Material. Difference in the reduction rates including Au and Pd that possess the large gap in noble metals (Pd2+ + 2e−→ Pd, E0 = 0.951 V vs. Au3+ + 3e−→ Au, E0 = 1.498 V) [63,69].
In this work, the slow synthesis methodology was carefully adopted to synthesize PdAu solid solution alloy so that all components are distributed randomly and homogeneously. Based on this idea, activated carbon (AC) with a developed porous structure was used as basic support to deposit ultrafine PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) solid solution alloy NPs. PdAu alloy NPs were successfully obtained and catalyzed FA dehydrogenation. By regulating various composition ratios, the optimal Pd0.5Au0.5 catalyst exhibits the highest activity (TOF = 1648 h−1) for HCOOH/HCOONa (formic acid/formate (FA/SF)) dehydrogenation at 313 K. Furthermore, reaction temperature, FA/SF ratios, metal loading quantity and stability of catalysts were also tested.
2 Experiment
2.1 Chemicals
The Na2PdCl4 (99% purity), HAuCl4·3H2O (99% purity), NaBH4 (98% purity), HCOOH (88% purity), and HCOONa (99.5% purity) were obtained from Aladdin Chemistry Co., Ltd., while the AC was obtained from Macklin Chemicals Co., Ltd. All the chemicals were not further purified.
2.2 Catalyst characterizations
The crystal structures of PdAu alloy NPs were detected via XRD using a SHIMADZU XRD-6000 with Cu Kα radiation (λ = 0.15406 nm). The 2θ data from 30° to 90° were 0.02°-intervals. The physical properties of catalysts including SBET (m2/g), Vp (cm3/g), and Dp (nm) were determined using the nitrogen-sorption method on a constant volume adsorption equipment (Micrometrics ASAP 2020 Plus HD88, USA). The pore volume was analyzed by the Barrett-Joyner-Halenda (BJH). The microstructure and morphology were tested by the HAADF-STEM and EDX-mapping on the Thermo Scientific Talos-F200X. Thermo Scientific K-Alpha with monochromatic Al K-Alpha (1486.7 eV) was carried to record X-ray photoelectron spectroscopy. The atomic ratios of Pd and Au was recorded using an inductively coupled plasma mass spectroscopy (ICP-MS) on a NexIONTM 350D instrument (PerkinElmer, USA).
2.3 Catalyst preparation
As shown in Fig.1, the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloy NPs with an appropriate composition and a metal loading of 5% were prepared. Taking Pd0.5Au0.5 synthesis as an example, Na2PdCl4 (29 mg, 0.1 mmol) and HAuCl4·3H2O (0.1 mmol) were added in 60 mL H2O and stirred for 30 min as metal precursor solutions. NaBH4 (152 mg, 4 mmol) was added in 60 mL ethyleneglycol (EG). AC (580 mg) was dissolved in 100 mL H2O. After that, the metal precursor solution and EG solution were added dropwise to the AC mixed solution at a rate of 1.2 mL/min [70]. After the dropping process, the mixed solution was stirred for 30 min. The sediment obtained was collected by centrifugation and washed over 5 times with H2O. The sediment collected by centrifugation was washed with H2O for 5 times. Finally, the product was dried at 60 °C under vacuum. The PdxAu1−x (x = 0, 0.2, 0.4, 0.6, 0.8, 1) alloy NPs and x% Pd0.5Au0.5 (x = 1, 5, 10) with different metal loadings were also obtained by altering the atomic ratios of Pd2+ and Au2+.
2.4 Catalyst evaluation
The synthesized catalyst was dispersed in deionized water in a round-bottomed flask and placed in a water bath at the preset temperature. Flask was connected to a reflux and a gas dropped where the gas produced from the FA dehydrogenation was collected. Subsequently, FA/SF in a certain molar ratio (1:3, 2:3, 1:8, 1:1, 3:2, and 3:1) was injected to the flask and started to react with vigorous stirring. Moreover, the activation energy (Ea) of reaction was calculated by conducting the catalytic reaction over a temperature range of 293–313 K. From the slope of each line, the rate constants K were obtained at preset temperatures. Finally, the TOF values were obtained using Eq. (1).
where Patm is the atmospheric pressure (101325 Pa), Vgas is the gas volume produced from the FA dehydrogenation reaction which has reached 20%, R is the universal gas constant, T is preset temperature (313 K), nPdAu is the total mole amount of PdAu atoms, and t is FA decomposition time (min). Based on ICP-MS analyses, the total mole amount of PdAu atoms on the AC are shown in Table S2.
3 Results and discussion
3.1 XRD
To investigate the structures of the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloy NPs, the XRD patterns were obtained as shown in Fig.2. The lattice constants and crystal sizes of all the PdAu NPs were summarized in Tab.1. For Au NPs, apparent diffraction peaks appear at 38.3°, 44.3°, 64.7°, and 77.7° are assigned to the (111), (200), (220) and (331) planes of the fcc-Au (JCPDS No. 65-8601), showing the structural formation of Au NPs. Compared to the Au NPs, the diffraction peaks of Pd NPs are weaker, exhibiting that the particle size of Pd NPs should be smaller. With the increase in Au mixing content, the diffraction peaks of PdxAu1−x alloy NPs shift to lower Bragg angles (Fig.2), but remain between the fcc-Au and fcc-Pd (JCPDS No. 05-0681) phases. According to the nominal composition, the Pd and Au atoms are randomly attributed to each lattice site. The lattice constant and crystal size of PdAu alloys decrease significantly with the mixing of palladium because of the smaller ionic radii of Pd compared to those of Au (Tab.1). These obvious characterizations, including lattice constant, diffraction peaks shift, and mean crystal size, strongly suggest the formations of the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloys in the whole composition range.
3.2 N2 adsorption/desorption measurements
N2 sorption isotherms and corresponding pore size distributions plot for PdAu alloys are shown in Fig.3. All PdAu alloy NPs exhibit typical IV isotherms with H1 hysteresis loops, which is characteristic of mesoporous structure. The values of SBET, Vp, and Dp of all the PdAu alloy NPs are listed in Tab.1. The AC support used provides a large specific surface area for the anchorage of the PdAu alloy. The surface area (1166 m2/g) and total pore volume (0.622 cm3/g) of Pd0.5Au0.5 far outperform those of other PdAu alloy NPs. When the Pd/Au ions ratio is unequal, the total pore volume and specific surface area decrease significantly. Such a reduction may be related to the aggregation of larger PdAu alloy NPs blocking pores and/or causing some structural rearrangements [71]. In general, the large surface and high porosity of PdAu alloys facilitate the active point between catalyst and FA, thus promoting FA dehydrogenation.
3.3 High angle annular dark field-scanning transmission electron microscope (HAADF-STEM)
The synthesized PdAu alloy NPs were characterized by HAADF-STEM analysis. The mean diameter of the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloy NPs were 10.9, 8.1, 7.3, 6.7, 6.5, 5.9, and 6.3 nm, respectively (Fig.4 and S1). These particle sizes were measured by averaging at least 100 nanoparticles. The mean particle sizes of Au NPs are relatively larger than that of other PdxAu1−x (x = 0.2, 0.4, 0.5, 0.6, 0.8, 1) NPs. This may be related to the fast nucleating nature of Au [61]. From the ICP-MS analysis, the atomic ratios of Pd and Au in alloys, shown in Table S2, were in consistent with the original ratios.
To clarify the Pd and Au ions of the mixing state in the PdxAu1−x alloy NPs, STEM EDX mappings of Pd-L and Au-L were performed. In Fig.4 and S2, the elemental maps of Pd and Au show that the two elements are randomly and homogeneously distributed in the whole area of NPs. The EDX line-scan profiles are shown in Fig.5 and S3, which indicate that the metal composition is transferred from Au-rich to Pd-rich. The Pd and Au ratios based on the EDX area-scan data are also listed in Table S2, which agree well with the theoretical Pd/Au atomic ratios in preparation. The EDX analysis suggests that the Pd/Au atomic ratios in the Pd0.5Au0.5 NPs is 49.3:50.7.
3.4 XPS measurements
As shown in Fig.6, the chemical states and surface electronic properties of the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloy NPs were investigated via X-ray photoelectron spectra. The remarkable XPS peaks of C 1s for PdAu alloy NPs at 284.8 and 286.4 eV are shown in Fig. S4 and assigned to the C–C/C=C and C–O, respectively [4,72]. Meanwhile, Pd0.2Au0.8 (Pd0 3d3/2 at 341.3 eV and Pd0 3d5/2 at 336.0 eV) possesses a higher BEs of Pd0 3d3/2 (340.9 eV) and Pd0 3d5/2 (335.9 eV) for PdxAu1–x (x = 0.4, 0.5, 0.6, 0.8, 1) alloy NPs. The deviation of BEs in the direction of high energy indicate that there exists an SMSI effect between PdAu and AC due to the introduction of Au [73]. In addition, the other two peaks at 343.3 and 337.5/338.2 eV is expected to Pd2+ 3d3/2 and Pd2+ 3d5/2. This may be caused by electronic interactions between the Pd d-orbitals and AC [74,75]. Au0 4f binding energies for the PdxAu1−x (x = 0.2, 0.4, 0.5, 0.6, 0.8) shift lower values compared to those of Au NPs (87.9 and 84.3 eV). The above XPS spectra of Pd 3d and electronegativity difference for Pd (2.2) and Au (2.4) also demonstrate that Pd loses a few electrons to Au in PdAu alloys. For the Pd0.5Au0.5, the BEs for Au0 4f are located at 87.8 and 84.2 eV. The XPS results indicate that there exists an SMSI effect between PdAu alloy and AC. It is worth noting that Pd 3d and Au 4f binding energy shift positively as the Au content increasing. The BEs of Pd/Au ratio is related to the interactions between Pd and Au. Due to the formation of non-homogeneous junctions in the electronic structure of Pd and the SMSI effect [76], the adsorption of formate can be enhanced from electron-rich Pd. These phenomena remarkably enhance the catalytic FA dehydrogenation [38,72].
3.5 FA dehydrogenation in FA/SF solution
Catalytic performances over the PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloys obtained toward the FA dehydrogenation with SF were measured in a burette system. The AC support and Au NPs showed a poor hydrogenation production (Fig.7(a)). As a synergistic promoter of metals, the contribution of Au to the decomposition of FA is assigned to the SMSI effect at the heterogeneous interfaces between the PdAu alloy and the AC support. As shown in Fig.7(a) and S5, the Pd0.5Au0.5 has the highest catalytic activity (TOF = 1648 h−1) of all the PdxAu1−x (x = 0, 0.2, 0.4, 0.6, 0.8, 1) NPs. The excellent catalytic performance of Pd0.5Au0.5 due to the ultrafine and polydispersity PdAu NPs on the surface of AC and electrons transfer efficiently from Pd to Au over PdAu alloy NPs. Meanwhile, the large specific surface area of AC also improves the dispersion of active metal sites, allowing for a greater exposure to the FA dehydrogenation. Moreover, Au doping allows the Pd to be the electron-state, thus enhancing the catalytic performances for FA decomposition efficiently. In FA decomposition, SF is used as an accelerator to increase the reaction activity of hydrogen evolution reaction [19,77]. The generated gas compositions from the aqueous FA/SF solution over the Pd0.5Au0.5 are shown in Fig. S6. The optimal ratio of FA/SF for Pd0.5Au0.5 is 1:3 at 313 K and it showed a conversion of around 100% for FA decomposition near room temperature (Fig.7(c) and S6). Additionally, optimum metal loading of PdAu on the AC is 5 wt.% (mass fraction) in FA dehydrogenation (Fig.7(d)).
The Arrhenius plot and corresponding kinetic parameters over Pd0.5Au0.5 and Pd NPs catalysts were shown in Fig.7(b) and S7. Through the rate versus time, the Arrhenius plot shows a slope of −6.8144, indicating that FA decomposition is nearly a first order reaction relative to the reaction temperature. In addition, the apparent activation energy (Ea) for Pd NPs and Pd0.5Au0.5 alloy NPs of FA/SF is obtained from 293 to 313 K. The above analysis indicates that the rate of H2 production increases with rising reaction temperature. According to the Arrhenius equation, the Ea of Pd0.5Au0.5 is calculated at 56.65 kJ/mol lower than that of Pd NPs at 64.66 kJ/mol. As displayed in Fig.8, the superior catalytic performance of the Pd0.5Au0.5 alloys can be ascribed to the well-distribution of bimetal active sites and synergy effect between Pd and Au that reduce activation energy obviously and boost FA decomposition.
3.6 Recyclability of optimized Pd0.5Au0.5 catalyst for FA dehydrogenation
As shown in Fig.9, the stability and recyclability of Pd0.5Au0.5 for FA decomposition at 313 K were also measured. Although the gas produced slightly decreased, the complete conversion of the Pd0.5Au0.5 catalyst was still maintained after 5 cycles, thus displaying a high stability of Pd0.5Au0.5 catalyst. The STEM EDX mappings of Pd and Au were conducted to show the catalysts recovered after 5 cycles (Fig.10). Before and after the catalytic decomposition of FA, there exists no significant change in size and morphology of Pd0.5Au0.5 alloy NPs, indicating the excellent structure stability of the catalyst.
4 Conclusions
In this study, PdxAu1−x (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1) alloy NPs over the whole composition range with various metal loading (x = 1, 5, 10 wt.%) were synthesized via a facile slow synthesis methodology. As support, the AC with large specific surface areas enlarge the polydispersity of PdAu alloys. In all PdAu alloy NPs, the as-synthesized Pd0.5Au0.5 showed the best activity, stability and recyclability for H2 production in FA/SF mixed solution at 313 K, affording an excellent initial TOF value of up to 1648 h−1 with a lower Ea (56.65 kJ/mol). The optimal activity of Pd0.5Au0.5 was mainly due to the ultra-small alloy NPs, particle size polydispersity, and uniform mixing states of Pd and Au. The catalyst was also confirmed to catalyze FA dehydrogenation near room temperature, which could encourage the wide utilization of FA as one of the prospective recyclable liquidus H2 storage materials.
PeplowM. Hydrogen economy looks out of reach. Nature, 2004, https://doi.org/10.1038/news041004-13
[2]
Singh S K, Zhang X B, Xu Q. Room-temperature hydrogen generation from hydrous hydrazine for chemical hydrogen storage. Journal of the American Chemical Society, 2009, 131(29): 9894–9895
[3]
Scofield M E, Liu H, Wong S S. A concise guide to sustainable PEMFCs: Recent advances in improving both oxygen reduction catalysts and proton exchange membranes. Chemical Society Reviews, 2015, 44(16): 5836–5860
[4]
Li S J, Zhou Y T, Kang X. . A simple and effective principle for a rational design of heterogeneous catalysts for dehydrogenation of formic acid. Advanced Materials, 2019, 31(15): 1806781
[5]
Wang W, He T, Yang X. . General synthesis of amorphous PdM (M = Cu, Fe, Co, Ni) alloy nanowires for boosting HCOOH dehydrogenation. Nano Letters, 2021, 21(8): 3458–3464
[6]
Yadav M, Xu Q. Liquid-phase chemical hydrogen storage materials. Energy & Environmental Science, 2012, 5(12): 9698
[7]
Grad O, Mihet M, Dan M. . Au/reduced graphene oxide composites: Eco-friendly preparation method and catalytic applications for formic acid dehydrogenation. Journal of Materials Science, 2019, 54(9): 6991–7004
[8]
Nielsen M, Alberico E, Baumann W. . Low-temperature aqueous-phase methanol dehydrogenation to hydrogen and carbon dioxide. Nature, 2013, 495(7439): 85–89
[9]
Yi N, Saltsburg H, Flytzani-Stephanopoulos M. Hydrogen production by dehydrogenation of formic acid on atomically dispersed gold on Ceria. ChemSusChem, 2013, 6(5): 816–819
[10]
Shen J, Chen W, Lv G. . Hydrolysis of NH3BH3 and NaBH4 by graphene quantum dots-transition metal nanoparticles for highly effective hydrogen evolution. International Journal of Hydrogen Energy, 2021, 46(1): 796–805
[11]
Wang Y, Liu X. Catalytic hydrolysis of sodium borohydride for hydrogen production using magnetic recyclable CoFe2O4-modified transition-metal nanoparticles. ACS Applied Nano Materials, 2021, 4(10): 11312–11320
[12]
Zhang Z, Lu Z, Tan H. . CeOx-modified RhNi nanoparticles grown on rGO as highly efficient catalysts for complete hydrogen generation from hydrazine borane and hydrazine. Journal of Materials Chemistry. A, Materials for Energy and Sustainability, 2015, 3(46): 23520–23529
[13]
Hong X, Yao Q, Huang M. . Bimetallic NiIr nanoparticles supported on lanthanum oxy-carbonate as highly efficient catalysts for hydrogen evolution from hydrazine borane and hydrazine. Inorganic Chemistry Frontiers, 2019, 6(9): 2271–2278
[14]
Jiang Y, Kang Q, Zhang J. . High-performance nickel–platinum nanocatalyst supported on mesoporous alumina for hydrogen generation from hydrous hydrazine. Journal of Power Sources, 2015, 273: 554–560
[15]
Dai H, Qiu Y, Dai H. . A study of degradation phenomenon of Ni–Pt/CeO2 catalyst towards hydrogen generation from hydrous hydrazine. International Journal of Hydrogen Energy, 2017, 42(26): 16355–16361
[16]
Huang W, Liu X. The “on–off” switch for on-demand H2 evolution from hydrous hydrazine over Ni8Pt1/C nano-catalyst. Fuel, 2022, 315: 123210
[17]
Zheng J, Zhou H, Wang C. . Current research progress and perspectives on liquid hydrogen rich molecules in sustainable hydrogen storage. Energy Storage Materials, 2021, 35: 695–722
[18]
Wang Y, Liu X. Enhanced catalytic performance of cobalt ferrite by a facile reductive treatment for H2 release from ammonia borane. Journal of Molecular Liquids, 2021, 343: 117697
[19]
Sun Q, Wang N, Xu Q. . Nanopore-supported metal nanocatalysts for efficient hydrogen generation from liquid-phase chemical hydrogen storage materials. Advanced Materials, 2020, 32(44): 2001818
[20]
Yang L, Hua X, Su J. . Highly efficient hydrogen generation from formic acid-sodium formate over monodisperse AgPd nanoparticles at room temperature. Applied Catalysis B: Environmental, 2015, 168–169: 423–428
[21]
Gao Y, Hu E, Yin G. . Pd nanoparticles supported on CeO2 nanospheres as efficient catalysts for dehydrogenation from additive-free formic acid at low temperature. Fuel, 2021, 302: 121142
[22]
Suh M P, Park H, Prasad T. . Hydrogen storage in metal–organic frameworks. Chemical Reviews, 2012, 112(2): 782–835
[23]
Lee H, Kang D, Pyen S. . Production of H2-free CO by decomposition of formic acid over ZrO2 catalysts. Applied Catalysis A, General, 2017, 531: 13–20
[24]
Faroldi B M, Conesa J M, Guerrero-Ruiz A. . Efficient nickel and copper-based catalysts supported on modified graphite materials for the hydrogen production from formic acid decomposition. Applied Catalysis A, General, 2022, 629: 118419
[25]
Lyu Y, Xie J, Wang D. . Review of cell performance in solid oxide fuel cells. Journal of Materials Science, 2020, 55(17): 7184–7207
[26]
Wei K, Wang X, Budiman R. . Progress in Ni-based anode materials for direct hydrocarbon solid oxide fuel cells. Journal of Materials Science, 2018, 53(12): 8747–8765
[27]
Bielinski E A, Lagaditis P O, Zhang Y. . Lewis acid-assisted formic acid dehydrogenation using a pincer-supported iron catalyst. Journal of the American Chemical Society, 2014, 136(29): 10234–10237
[28]
Deng M, Ma J, Liu Y. . Pd nanoparticles confined in pure Silicalite-2 zeolite with enhanced catalytic performance for the dehydrogenation of formic acid at room temperature. Fuel, 2023, 333: 126466
[29]
Mori K, Dojo M, Yamashita H. Pd and Pd–Ag nanoparticles within a macroreticular basic resin: An efficient catalyst for hydrogen production from formic acid decomposition. ACS Catalysis, 2013, 3(6): 1114–1119
[30]
Peng W, Liu S, Li X. . Robust hydrogen production from HCOOH over amino-modified KIT-6-confined PdIr alloy nanoparticles. Chinese Chemical Letters, 2022, 33(3): 1403–1406
[31]
Bulushev D, Beloshapkin S, Plyusnin P. . Vapour phase formic acid decomposition over PdAu/γ-Al2O3 catalysts: Effect of composition of metallic particles. Journal of Catalysis, 2013, 299: 171–180
[32]
Nasiri R, Gholipour B, Nourmohammadi M. . Mesoporous hybrid organosilica for stabilizing Pd nanoparticles and aerobic alcohol oxidation through Pd hydride (Pd–H2) species. International Journal of Hydrogen Energy, 2023, 48(17): 6488–6498
[33]
Alamgholiloo H, Rostamnia S, Hassankhani A. . Formation and stabilization of colloidal ultra-small palladium nanoparticles on diamine-modified Cr-MIL-101: Synergic boost to hydrogen production from formic acid. Journal of Colloid and Interface Science, 2020, 567: 126–135
[34]
Doustkhah E, Rostamnia S, Imura M. . Thiourea bridged periodic mesoporous organosilica with ultra-small Pd nanoparticles for coupling reactions. RSC Advances, 2017, 7(89): 56306–56310
[35]
Ahadi A, Rostamnia S, Panahi P. . Palladium comprising dicationic bipyridinium supported periodic mesoporous organosilica (PMO): Pd@Bipy–PMO as an efficient hybrid catalyst for Suzuki–Miyaura cross-coupling reaction in water. Catalysts, 2019, 9(2): 140
[36]
Farajzadeh M, Alamgholiloo H, Nasibipour F. . Anchoring Pd-nanoparticles on dithiocarbamate-functionalized SBA-15 for hydrogen generation from formic acid. Scientific Reports, 2020, 10(1): 18188
[37]
Yang Y, Xu H, Cao D. . Hydrogen production via efficient formic acid decomposition: Engineering the surface structure of Pd-based alloy catalysts by design. ACS Catalysis, 2019, 9(1): 781–790
[38]
Wang Z, Liang S, Meng X. . Ultrasmall PdAu alloy nanoparticles anchored on amine-functionalized hierarchically porous carbon as additive-free catalysts for highly efficient dehydrogenation of formic acid. Applied Catalysis B: Environmental, 2021, 291: 120140
[39]
Gu X, Lu Z, Jiang H. . Synergistic catalysis of metal–organic framework-immobilized Au–Pd nanoparticles in dehydrogenation of formic acid for chemical hydrogen storage. Journal of the American Chemical Society, 2011, 133(31): 11822–11825
[40]
Pritchard J, Kesavan L, Piccinini M. . Direct synthesis of hydrogen peroxide and benzyl alcohol oxidation using Au−Pd catalysts prepared by sol immobilization. Langmuir, 2010, 26(21): 16568–16577
[41]
Al-Nayili A, Albdiry M. AuPd bimetallic nanoparticles supported on reduced graphene oxide nanosheets as catalysts for hydrogen generation from formic acid under ambient temperature. New Journal of Chemistry, 2021, 45(22): 10040–10048
[42]
Sanchez F, Bocelli L, Motta D. . Preformed Pd-based nanoparticles for the liquid phase decomposition of formic acid: Effect of stabiliser, support and Au–Pd ratio. Applied Sciences, 2020, 10(5): 1752
[43]
Dong A, Jiang Q, Zhou Y. Au3Pd1 intermetallic compound as single atom catalyst for formic acid decomposition with highly hydrogen selectivity. International Journal of Hydrogen Energy, 2023, 48(76): 29542–29551
[44]
Barlocco I, Capelli S, Lu X. . Disclosing the role of gold on palladium—Gold alloyed supported catalysts in formic acid decomposition. ChemCatChem, 2021, 13(19): 4210–4222
[45]
Feng C, Wang Y, Gao S. . Hydrogen generation at ambient conditions: AgPd bimetal supported on metal–organic framework derived porous carbon as an efficient synergistic catalyst. Catalysis Communications, 2016, 78: 17–21
[46]
Dai H, Xia B, Wen L. . Synergistic catalysis of AgPd@ZIF-8 on dehydrogenation of formic acid. Applied Catalysis B: Environmental, 2015, 165: 57–62
[47]
Gholipour B, Zonouzi A, Shokouhimehr M. . Integration of plasmonic AgPd alloy nanoparticles with single-layer graphitic carbon nitride as Mott-Schottky junction toward photo-promoted H2 evolution. Scientific Reports, 2022, 12(1): 13583
[48]
Feng T, Wang J, Gao S. . Covalent triazine frameworks supported CoPd nanoparticles for boosting hydrogen generation from formic acid. Applied Surface Science, 2019, 469: 431–436
[49]
Cheng J, Gu X, Liu P. . Controlling catalytic dehydrogenation of formic acid over low-cost transition metal-substituted AuPd nanoparticles immobilized by functionalized metal–organic frameworks at room temperature. Journal of Materials Chemistry. A, Materials for Energy and Sustainability, 2016, 4(42): 16645–16652
[50]
Cao N, Tan S, Luo W. . Ternary CoAgPd nanoparticles confined inside the pores of MIL-101 as efficient catalyst for dehydrogenation of formic acid. Catalysis Letters, 2016, 146(2): 518–524
[51]
Yan J, Wang Z, Gu L. . AuPd–MnOx/MOF–graphene: An efficient catalyst for hydrogen production from formic acid at room temperature. Advanced Energy Materials, 2015, 5(10): 1500107
[52]
Jin M, Oh D, Park J. . Mesoporous silica supported Pd-MnOx catalysts with excellent catalytic activity in room-temperature formic acid decomposition. Scientific Reports, 2016, 6(1): 33502
[53]
Sun X, Ding Y, Feng G. . Carbon bowl-confined subnanometric palladium-gold clusters for formic acid dehydrogenation and hexavalent chromium reduction. Journal of Colloid and Interface Science, 2023, 645: 676–684
[54]
Luo Y, Yang Q, Nie W Y. . Anchoring IrPdAu nanoparticles on NH2-SBA-15 for fast hydrogen production from formic acid at room temperature. ACS Applied Materials & Interfaces, 2020, 12(7): 8082–8090
[55]
Zhong H, Iguchi M, Chatterjee M. . Formic acid-based liquid organic hydrogen carrier system with heterogeneous catalysts. Advanced Sustainable Systems, 2018, 2(2): 1700161
[56]
Zhong H, Iguchi M, Chatterjee M. . Interconversion between CO2 and HCOOH under basic conditions catalyzed by PdAu nanoparticles supported by amine-functionalized reduced graphene oxide as a dual catalyst. ACS Catalysis, 2018, 8(6): 5355–5362
[57]
Szumełda T, Drelinkiewicz A, Lalik E. . Carbon-supported Pd100−XAuX alloy nanoparticles for the electrocatalytic oxidation of formic acid: Influence of metal particles composition on activity enhancement. Applied Catalysis B: Environmental, 2018, 221: 393–405
[58]
Tedsree K, Li T, Jones S. . Hydrogen production from formic acid decomposition at room temperature using a Ag-Pd core-shell nanocatalyst. Nature Nanotechnology, 2011, 6(5): 302–307
[59]
Zhao X, Xu D, Liu K. . Remarkable enhancement of PdAg/rGO catalyst activity for formic acid dehydrogenation by facile boron-doping through NaBH4 reduction. Applied Surface Science, 2020, 512: 145746
[60]
Zhao X, Dai P, Xu D. . Ultrafine palladium nanoparticles anchored on NH2-functionalized reduced graphene oxide as efficient catalyst towards formic acid dehydrogenation. International Journal of Hydrogen Energy, 2020, 45(55): 30396–30403
[61]
Kusada K, Wu D, Nanba Y. . Highly stable and active solid-solution-alloy three-way catalyst by utilizing configurational-entropy effect. Advanced Materials, 2021, 33(16): 2005206
[62]
Zhang Q, Kusada K, Wu D. . Solid-solution alloy nanoparticles of a combination of immiscible Au and Ru with a large gap of reduction potential and their enhanced oxygen evolution reaction performance. Chemical Science, 2019, 10(19): 5133–5137
[63]
Zhang Q, Kusada K, Wu D. . Selective control of fcc and hcp crystal structures in Au-Ru solid-solution alloy nanoparticles. Nature Communications, 2018, 9(1): 510
[64]
Deng C, Li Y, Sun W. . Supported AuPd nanoparticles with high catalytic activity and excellent separability based on the magnetic polymer carriers. Journal of Materials Science, 2019, 54(17): 11435–11447
[65]
Huang B, Kobayashi H, Yamamoto T. . Solid-solution alloying of immiscible Ru and Cu with enhanced CO oxidation activity. Journal of the American Chemical Society, 2017, 139(13): 4643–4646
[66]
Zhang Z J, Zhang S L, Yao Q L. . Metal–organic framework immobilized RhNi alloy nanoparticles for complete H2 evolution from hydrazine borane and hydrous hydrazine. Inorganic Chemistry Frontiers, 2018, 5(2): 370–377
[67]
Gao S, Wang L, Li H. . Core–shell PdAu nanocluster catalysts to suppress sulfur poisoning. Physical Chemistry Chemical Physics, 2021, 23(28): 15010–15019
[68]
Scott R, Wilson O, Oh S. . Bimetallic palladium-gold dendrimer-encapsulated catalysts. Journal of the American Chemical Society, 2004, 126(47): 15583–15591
[69]
Zhu C, Guo S, Dong S. PdM (M = Pt, Au) bimetallic alloy nanowires with enhanced electrocatalytic activity for electro-oxidation of small molecules. Advanced Materials, 2012, 24(17): 2326–2331
[70]
Tan Z, Haneda M, Kitagawa H. . Slow synthesis methodology-directed immiscible octahedral PdxRh1−x dual-atom-site catalysts for superior three-way catalytic activities over Rh. Angewandte Chemie International Edition, 2022, 61(23): 202202588
[71]
Chu P K, Li L H. Characterization of amorphous and nanocrystalline carbon films. Materials Chemistry and Physics, 2006, 96(2–3): 253–277
[72]
Zhang A, Xia J, Yao Q. . Pd–WO heterostructures immobilized by MOFs-derived carbon cage for formic acid dehydrogenation. Applied Catalysis B: Environmental, 2022, 309: 121278
[73]
Zhang B, Su D. Probing the metal-support interaction in carbon-supported catalysts by using electron microscopy. ChemCatChem, 2015, 7(22): 3639–3645
[74]
Guerrero-Ortega L, Ramírez-Meneses E, Cabrera-Sierra R. . Pd and Pd@PdO core–shell nanoparticles supported on Vulcan carbon XC-72R: Comparison of electroactivity for methanol electro-oxidation reaction. Journal of Materials Science, 2019, 54(21): 13694–13714
[75]
Nowicka E, Althahban S, Luo Y. . Highly selective PdZn/ZnO catalysts for the methanol steam reforming reaction. Catalysis Science & Technology, 2018, 8(22): 5848–5857
[76]
Costa L, Vasconcelos S, Pinto A. . Rh/CeO2 catalyst preparation and characterization for hydrogen production from ethanol partial oxidation. Journal of Materials Science, 2008, 43(2): 440–449
[77]
Jiang K, Xu K, Zou S. . B-doped Pd catalyst: Boosting room-temperature hydrogen production from formic acid-formate solutions. Journal of the American Chemical Society, 2014, 136(13): 4861–4864
RIGHTS & PERMISSIONS
Higher Education Press 2023
AI Summary 中Eng×
Note: Please be aware that the following content is generated by artificial intelligence. This website is not responsible for any consequences arising from the use of this content.