2. University of Chinese Academy of Sciences, Beijing 100049, China
2. 中国科学院大学, 北京 100049
With the increasing global greenhouse effect and increasingly serious global warming phenomenon, ammonia has attracted wide attention as a green energy source, because it does not produce carbon dioxide during combustion. In the meanwhile, ammonia is one of the important basic chemicals and an important fertilizer raw material, which is widely used in the manufacture of textiles, plastics and fertilizers[1]. At present, the conventional Haber-Bosch process continues to be employed for industrial ammonia synthesis, which uses Fe-based catalysts and requires harsh reaction conditions (300~500 ℃, 10~30 MPa)[2]. Due to the high energy required for ammonia synthesis, 2% of global energy is used to produce ammonia each year[3]. Therefore, it is very important to develop efficient ammonia synthesis catalysts under mild conditions. At low temperature and low pressure, the ammonia synthesis activity of Ru catalyst is better than that of Fe catalyst[4−5]. However, as a precious metal, the concentration of Ru in the Earth's crust is 1.0×10−9, so its high price impedes its further application in industry[6]. Therefore, it is necessary to improve the catalytic performance of Ru. Thus, the economical and sustainable utilization of Ru as a precious metal in ammonia synthesis can be realized.
DFT calculation and experiments show that the synthesis of ammonia on Ru-based catalyst is a structure-sensitive reaction, where in the particle size, morphology and exposed facets of Ru nanoparticles significantly influence their catalytic activity[7−10]. The B5 site is a collection of 5 Ru atoms on Ru nanoparticles and is the most active center for N2 dissociation. The small-size particles of 1.8~3.5 nm are more conducive to the formation of the B5 sites[11]. The performance of Ru based catalyst is related to the electronic structure of Ru nanoparticles. A series of studies have proved that the catalytic performance of the catalyst can be achieved by adjusting the electronic state of the Ru-based catalyst. Wei et al.[12] found that Ru loaded on CeO2 and MgO-CeO2 have different electronic states and thus exhibit different catalytic activities. Nagaoka et al.[13] modified the Ru electronic structure through the strong interaction between metal and the reduction support, thereby improving the ammonia synthesis activity.
In addition, most metals have bcc, hcp and fcc three crystalline phase structures. The coordination environment and spatial distribution of metal atoms vary greatly in different crystal phases[14]. Therefore, it is significant to regulate the crystal phase of the metal in improving the catalytic performance. In recent years, regulating crystal phase structure of Ru has gradually attracted the attention of researchers. Bulk Ru has hcp crystal structure in all temperature ranges[15]. With the development of nanotechnology, Ru with fcc crystal phase can be stabilized under both synthetic and reaction conditions. Kitagawa et al.[15] synthesized uniform sized Ru nanoparticles with different crystal phases using chemical reduction methods with different metal precursors. Recently, researches have been conducted on improving the catalytic activity by adjusting the crystal phase of Ru. Chen et al.[16] investigated the effect of Ru crystal on hydrolytic dehydrogenation of ammonia borane. The results indicated that the catalytic performance of fcc Ru is mainly affected by surface oxidation, while the activity of hcp Ru is predominantly affected by the size effect. Ma et al.[17] synthesized the water-dispersed fcc Ru catalyst with abundant open surfaces and improved its catalytic activity in Fischer-Tropsch synthesis. This is mainly because in the process of CO dissociation, fcc Ru exposed more active sites on the crystal surface than hcp Ru. Decomposing N≡N requires a high energy barrier(945 kJ·mol−1), so it is generally believed that N2 dissociation is the rate determining step in ammonia synthesis[18−19]. Xia's group[20−23] reported the preparation of fcc Ru nanocages with different surface structures and evaluated the activity of fcc Ru for N2 reduction synthesis of ammonia through DFT calculations. The results showed that N2 molecules were more strongly binding to fcc-(111) than hcp-(0001). At the same time, the activation energy barrier of fcc-(111) for N−N dissociation is decreased, indicating that fcc-(111) has a stronger effect on N2 dissociation. In addition, although the (100) surface preferentially exposed by fcc Ru has a slightly weaker binding to N, its activation energy barrier is further reduced. These results indicate that Ru nanoparticles with fcc crystal phase are potential catalysts in ammonia synthesis. However, up to now, the study of Ru crystal phase in ammonia synthesis catalyst has been confined to theoretical calculation, and the experimental part has not been confirmed.
In this study, Ru nanoparticles with fcc and hcp crystal phases were obtained via chemical reduction method by controlling the precursor and reducing agent, and the catalytic properties in ammonia synthesis were compared. The fcc Ru catalyst shows higher catalytic reaction rate(666.4 μmol·h−1·g−1) by comparing with the hcp Ru(378.9 μmol·h−1·g−1) at the reaction temperature(400 ℃) and pressure(1 MPa). This is mainly because the exposed crystal faces of fcc Ru and hcp Ru nanoparticles are different, and the dissociation ability of fcc-(111) to N2 is stronger, so it has higher ammonia synthesis activity, which is consistent with the theoretical research findings of predecessors. On this basis, the obtained Ru-based catalyst was loaded on rod-like CeO2 support to further improve its ammonia synthesis activity. At 400 ℃, the activity of fcc Ru/CeO2 is
All the reagents used in the experiment are commercially available. They can be directly used without any further purification or treatment. The purity of the gases used in the experiment was 99.999%.
1.2 Preparation of catalysts and supportFor the synthesis of fcc Ru, 83.7 mg Ru(acac)3 and 55.5 mg PVP(k29-32) were dissolved in 10 mL triethylene diethylene glycol(TEG), and then heated in an oil bath at 200 ℃ for 3 h. After cooling the reaction to room temperature, the products were washed and centrifuged with acetone. The volume ratio of acetone to the solvent used was 3. The preparation method of hcp Ru is similar to fcc Ru, except that 66.24 mg RuCl3 as precursor and 10 mL glycol (EG) as solvent.
For the synthesis of CeO2, 6.96 g Ce(NO3)2·6H2O and 19.6 g NaOH were dissolved in 5 and 35 mL deionized water respectively, and then the two solutions were mixed, and finally, the obtained mixture transferred to a 50 mL hydrothermal reactor at a height of 4/5, and stored in a vacuum oven at 100 ℃ for 24 h. The precipitate obtained after hydrothermal reaction is washed with deionized water to neutral and dried in the oven at 60 ℃. After drying, the product was placed in a crucible and roasted in a muffle furnace at 450 ℃ for 4 h at a heating rate of 2.5 ℃·min−1. After roasting, the sample was taken out and CeO2 was obtained.
For the preparation of fcc Ru/CeO2, 15 mg fcc Ru was dissolved in 15 mL deionized water, 0.5 g of prepared CeO2 was added, impregnated and stirred at room temperature for 24 h, then centrifuged and washed with deionized water. The obtained product was fcc Ru/CeO2 after drying in a 60 ℃ oven. The loading mass fraction of Ru is 3%. The preparation method of hcp Ru/CeO2 (Ru mass fraction, 3%) is similar to fcc Ru/CeO2, except that the active metal used is 15 mg hcp Ru.
1.3 Catalyst evaluationThe catalyst activity was evaluated by ammonia synthesis. The experiment was conducted in a fixed bed, and the catalyst packed in a Ф6 mm stainless steel reaction tube at a dosage of 200 mg each time. The total flow rate of N2 and H2 is 60 mL·min−1(N2/H2=3), and the reaction pressure is 1 MPa. After the catalyst stabilizes at the testing temperature, the outlet ammonia concentration was determined by chemical titration (Congo red as an indicator, the concentration of H2SO4 used was 0.05 mol·L−1).
1.4 Catalyst characterizationThe crystal structure of the catalyst was characterized by X-ray powder diffraction (XRD)(X'Pert PANalytical, Dutch), using Cu-Kα radiation(λ=
We synthesized Ru nanoparticles with different crystal phases by chemical reduction method. By adjusting the precursors and reducing agents, Ru nanoparticles with different crystal phases were prepared. The morphology and particle size distribution of fcc Ru and hcp Ru were studied using TEM. Ru nanoparticles prepared by Ru(acac)3 and TEG exhibited fcc crystal phase, while those prepared by RuCl3 and EG had hcp structure. As shown in Fig. 1(a) and (c), the diameter of fcc Ru nanoparticles is mostly 2.8 nm, while hcp Ru is 2.0 nm, which are consistent with the size distribution of B5 site of Ru catalyst in ammonia synthesis. The lattice fringes in HRTEM further confirm the difference between the two crystal phases. As displayed in Fig. 1(b) and (d), the lattice fringe of fcc Ru is 0.20 and 0.22 nm, corresponding to its (200) and (111) crystal planes, respectively. The lattice fringe of 0.23 nm in hcp Ru is the (100) crystal plane.
![]() |
Fig.1 TEM images, particle size distribution and HRTEM images of fcc Ru (a, b) and hcp Ru (c, d) |
The structure of the prepared Ru nanoparticles was characterized using high-resolution transmission electron microscopy (HRTEM) and selected area electron diffraction (SAED). As shown in Fig. 2, it can be observed that the hcp Ru nanoparticles (Fig. 2(a)) display the lamination sequence of hcp ABABAB… along the direction [100]. However, the fcc Ru nanoparticles showed the stacking order of ABCABC…(Fig. 2(c)), which shows the representing fcc [111] plane of Ru nanoparticles. In the SAED results, (111), (200), (220) and (311) planes of fcc Ru were observed, while hcp Ru showed (100) and (110) planes. Therefore, there are significant differences in the exposed crystal faces of Ru nanoparticles with different crystal phases.
![]() |
Fig.2 HRTEM and SAED images of hcp Ru (a, b) and fcc Ru (c, d) |
XRD analysis further confirmed the difference of crystal phase between the two Ru nanoparticles. From Fig. 3, it can be seen that fcc Ru and hcp Ru exhibited different crystal faces, consistent with those previously reported[22]. The diffraction peaks around 40.8º, 47.4º and 69.3º can be corresponded to (111), (200) and (220) planes of the fcc Ru, and the peaks at 38.4º, 42.2º, 44.0º, 58.3º and 69.4º can correspond to the (100), (002), (101), (102) and (110) diffractions of hcp Ru.
![]() |
Fig.3 XRD patterns of fcc Ru and hcp Ru |
Using ammonia synthesis as a model reaction, the influence of Ru crystal phase on catalytic reactions was investigated(Fig. 4). It was observed that the reaction rate (r) over fcc Ru (666.4 μmol·h−1·g−1) is about 1.76 times than that of hcp Ru (378.9 μmol·h−1·g−1) at the condition of 400 ℃ and 1 MPa(Fig. 4 (a)). The result clearly demonstrated that the crystal phase of Ru significantly affects the performance of ammonia synthesis. This is mainly because the exposed crystal surfaces of Ru nanoparticles with two crystal phases have different degrees of N2 dissociation, which leads to the difference in catalytic ammonia synthesis performance. Combined with our experimental results and the theoretical calculation of Xia et al.[18−21], the dissociation ability to N2 of fcc Ru exposed (111) and (200) is stronger than that of hcp Ru exposed (100), so the catalytic performance of fcc Ru in ammonia synthesis is superior to that of hcp Ru.
![]() |
Fig.4 Ammonia synthesis activities of (a) fcc Ru and hcp Ru (400 ℃, 1 MPa), (b) fcc Ru/CeO2 and hcp Ru/CeO2 (340−400 ℃, 1 MPa); XRD patterns of fcc Ru/CeO2 and hcp Ru/CeO2 (c), TEM image of CeO2 (d) |
To evaluate the support impact, CeO2 was selected as support for Ru nanoparticles with two crystal phases. CeO2 with nanorod morphology was synthesized (Fig. 4(d)) and fcc Ru/ CeO2 and hcp Ru/CeO2 were obtained by loading Ru of two crystal phases. Fig. 4(c) shows the XRD patterns of the two catalysts. Interestingly, both samples showed typical charact-eristic peaks of CeO2, and no diffraction peaks were observed for fcc Ru or hcp Ru species. This may be attributed to the low loading of Ru or the uniform distribution of active metal on the support. Fig. 4(b) showed that the catalytic activity of both fcc Ru/CeO2 and hcp Ru/CeO2 catalysts for ammonia synthesis reaction increased with the rise of temperature in the range of 340 to 400 ℃. In addition, because of the strong interaction between the metal and the support, it was observed that the catalytic activity of the Ru-based catalyst supported on CeO2 is higher than that of the unsupported Ru catalyst. Within in the testing temperature range, the ammonia synthesis activity of fcc Ru/CeO2 was superior than that of hcp Ru/CeO2. The activity of fcc Ru/CeO2 at 400 ℃ is
The elemental composition and electronic structure of Ru-based catalyst samples were further analyzed by XPS. Fig. 5 (a) and (b) show the Ru 3p3/2 orbital spectra of fcc Ru, hcp Ru, fcc Ru/CeO2 and hcp Ru/CeO2. Three Ru deconvolution peaks were observed at 461.5, 462.5 and 464.8 eV, corresponding to Ru0, Ru4+ (RuO2) and Rux+ (RuOx, 4<x<8)[24]. In all of these samples, the proportion of Ru0 is more than 50%, indicating that it is more dominant than the positively charged Ru. This indicates that there are much Ru nanoparticles or clusters present on the catalyst surface. When Ru nanoparticles with different crystal phases are loaded onto CeO2 support, the content of Ru0 in fcc crystalline Ru was almost unchanged, the proportion of Ru4+ increased from 28.3% to 38.5%, and the proportion of Rux+ decreased from 18.3% to 7.9%. The increase of Ru4+ content indicates the binding of Ce−O bond to Ru nanoparticles and there is an interaction between metal and support. For Ru with hcp phase, the content of Ru0 increases from 60.7% to 70.5%, the proportion of Ru4+ decreases from 25.7% to 10.3%, and the proportion of Rux+ increases from 13.6% to 19.2%. This result shows that the content of Ru0 is the most among the three Ru types, regardless of whether the Ru nanoparticles are loaded on the support or not. Metal Ru will promote the formation of weak basic sites[25], which facilitates the dissociation of N≡N. The Ce 3d spectra is shown in Fig.5(c), fitted to ten components. It can be seen that among the ten peaks, Ce3+ and Ce4+ exist simultaneously, and the four peaks labeled V0, V′, U0 and U′ belong to Ce3+, while the six peaks V, V″, V′″, U, U″ and U′″ correspond to Ce4+[26−27]. For the two samples, hcp Ru/CeO2 has a higher Ce3+/Ce4+(0.43) compared to fcc Ru/CeO2 (0.36). The oxygen vacancies in hcp Ru/CeO2 are higher than those in fcc Ru/CeO2 because the existence of Ce3+ is related to oxygen vacancies. Fig. 5(d) shows the spectra of O 1s. OL, OV and OC correspond to lattice oxygen bound with metal cations, O2− in the oxygen-deficient region, and chemically absorbed and dissociated oxygen, respectively. The oxygen vacancy can be roughly calculated by OV/OL, and the result obtained is that hcp Ru/CeO2(0.24) is higher than fcc Ru/CeO2(0.21). In general, the presence of oxygen vacancies can increase the electron density on the Ru metal surface, thus enhancing ammonia synthesis activity[28−29]. However, our results show that the fcc Ru/CeO2 with lower oxygen vacancy concentration exhibit higher ammonia synthesis activity, which indicates that the ammonia synthesis performance is more affected by the Ru crystal phase.
![]() |
Fig.5 XPS spectra of Ru 3p (a, b), Ce 3d (c) and O 1s (d) |
In order to investigate the reducibility of oxygen on the catalyst surface, H2-TPR test was conducted on the prepared catalyst. As shown in Fig. 6, there are two peaks were observed at 347 and 420 ℃ for fcc Ru, and at 391 and 477 ℃ for hcp Ru. These two peaks can be attributed to RuOx species. When Ru of different crystal phases is loaded onto CeO2 support, the reduction peaks of RuO2 was observed below 300 ℃ (109 and 268 ℃ for fcc Ru/CeO2, 117 and 266 ℃ for hcp Ru/CeO2), suggesting that there is an interaction between Ru and the support that is conducive to H2 reduction[30]. The peaks around 800 ℃ correspond to the reduction of the bulk of CeO2[31]. In addition, the hcp Ru/CeO2 exhibits two reduction peaks at 363 and 588 ℃, mainly due to the reduction of oxygen on the surface of CeO2[28]. Therefore, there is an interaction between the metal and the support, so that the catalyst supported on CeO2 has a higher ammonia synthesis activity.
![]() |
Fig.6 H2-TPR profiles of fcc Ru and hcp Ru (a), fcc Ru/CeO2 and hcp Ru/CeO2 (b) |
The electronegativity of the support has a significant impact on the performance of ammonia synthesis catalysts. The CO2-TPD profiles of different catalysts are shown in Fig. 7(a) and (b). The CO2 desorption temperatures of fcc Ru and hcp Ru occurred at 220, 510 ℃ and 220, 530 ℃, respectively. How-ever, there are three desorption peaks of fcc Ru/CeO2(137, 354 and 550 ℃), while hcp Ru/CeO2 has only two peaks(147 and 332 ℃). Generally speaking, the desorption temperature and number of desorption peaks of CO2 are related to the intensity and quantity of basic sites, respectively[14]. The CO2-TPD shows that fcc Ru/CeO2 is the strongest among the four catalysts, which is conducive to the electron transfer to Ru, thereby promoting N2 activation and N≡N dissociation, so fcc Ru/CeO2 has a higher ammonia synthesis activity.
![]() |
Fig.7 CO2-TPD (a, b), N2-TPD (c, d) and NH3-TPD (e, f) profiles of fcc Ru and hcp Ru (a, c, e), fcc Ru/CeO2 and hcp Ru/CeO2 (b, d, f) |
Further understanding of the effect of Ru exposed crystal faces on ammonia synthesis activity was obtained through N2-TPD. As shown in Fig. 7(c) and (d), N2 desorption occurred at 240, 502 ℃(fcc Ru) and 229, 534℃(hcp Ru), respectively, which indicating that fcc Ru has a strong adsorption capacity for N2. When loaded on CeO2, the number of N2 desorption peaks increased, and the desorption of N2 at fcc Ru/CeO2 and hcp Ru/CeO2 both occur at 150, 350 and 550 ℃. These results indicate that the interaction between the metal and the support can promote the adsorption of N2.
The adsorption and desorption of NH3 are also important factors in evaluating the activity of ammonia synthesis. The NH3-TPD profiles of different catalysts are presented in Fig. 7 (e) and (f). The desorption of NH3 by fcc Ru and hcp Ru occurs between 230 and 450 ℃. After loading on CeO2, the desorption temperature is advanced, and the first peak of NH3 desorption is observed at 155 and 140 ℃, respectively. There are five NH3 desorption peaks in fcc Ru/CeO2, while hcp Ru/CeO2 only has two peaks. The dissociation adsorption energy of N2 is linearly related to the adsorption energy of NHx[32]. The stronger the dissociation ability of the metal to N2, the more difficulty of subsequent hydrogenation to ammonia, and the weaker the adsorption energy of the metal, the more difficult it is for the dissociation of N2, and the lower the ammonia synthesis activity. Only when N2 dissociation and NHx desorption reach a relative equilibrium state, can higher ammonia synthesis activity be obtained. The Ru-based catalyst with fcc crystal phase has strong desorption of N2, so the desorption of NH3 is relatively weak. For Ru-based catalysts with hcp crystal phase, the activation capacity of N2 is poor, but the desorption capacity of NH3 is stronger. Therefore, the dissociation of N2 and desorption of NH3 of Ru-based catalysts with fcc crystal phase may be closer to the equilibrium state, and thus have higher ammonia synthesis activity.
3 ConclusionsIn conclusion, Ru nanoparticles with fcc and hcp crystal phases were prepared by chemical reduction method by regulating metal precursors and reducing agents. The research results indicate that exposed crystal planes have a certain impact on catalytic activity. The dissociation ability to N2 of fcc Ru exposed plane (111) and (200) is better than that of hcp Ru exposed plane (100). The fcc Ru catalyst shows a higher ammonia synthesis rate(666.4 μmol·h−1·g−1) comparing with the hcp Ru(378.9 μmol·h−1·g−1) at the reaction temperature(400 ℃) and pressure(1 MPa). On this basis, the obtained Ru-based catalyst was loaded on rod-like CeO2 support to further improve its ammonia synthesis activity. The catalytic activity of fcc Ru/CeO2 was higher than that of hcp Ru/CeO2 in the test temperature range. At 400 ℃, the activity of fcc Ru/CeO2 is
[1] |
Reversible ammonia-based and liquid organic hydrogen carriers for high-density hydrogen storage: Recent progress[J]. Int J Hydrogen Energy, 2019, 44(15): 7746–7767.
DOI:10.1016/j.ijhydene.2019.01.144 |
[2] |
Progress in green ammonia synthesis technology: Catalytic behavior of ammonia synthesis catalysts[J]. Adv Sustain Syst, 2024, 8(8): 2300618.
DOI:10.1002/adsu.202300618 |
[3] |
Sustainable ammonia production[J]. ACS Sustain Chem Eng, 2017, 5(11): 9527–9527.
DOI:10.1021/acssuschemeng.7b03719 |
[4] |
Ruthenium nanocatalysts for ammonia synthesis: A review[J]. Chem Eng Commun, 2015, 202(4): 420–448.
DOI:10.1080/00986445.2014.923995 |
[5] |
The ammonia-synthesis catalyst of the next generation: Barium-promoted oxide-supported ruthenium[J]. Angew Chem Int Edit, 2001, 40(6): 1061–1063.
DOI:10.1002/1521-3773(20010316)40:6<1061::AID-ANIE10610>3.0.CO;2-B |
[6] |
Primary steps in catalytic synthesis of ammonia[J]. J Vac Sci Technol A, 1983, 1(2): 1247–1253.
DOI:10.1116/1.572299 |
[7] |
Dissociative adsorption of N2 on Ru(0001): A surface reaction totally dominated by steps[J]. J Catal, 2000, 192(2): 381–390.
DOI:10.1006/jcat.2000.2858 |
[8] |
Surface science based microkinetic analysis of ammonia synthesis over ruthenium catalysts[J]. J Catal, 2000, 192(2): 391–399.
DOI:10.1006/jcat.2000.2857 |
[9] |
Efficient Ru/AC catalysts prepared by co-impregnation for ammonia synthesis[J]. J Mol Catal (China), 2013, 27(6): 530–538.
DOI:10.16084/j.cnki.issn1001-3555.2013.06.005 |
[10] |
Effect of CeO2 support basicity on the catalytic activity of Ru/CeO2 catalyst for ammonia synthesis[J]. J Mol Catal (China), 2018, 32(4): 349–358.
DOI:10.16084/j.cnki.issn1001-3555.2018.04.009 |
[11] |
Structure sensitivity of supported ruthenium catalysts for ammonia synthesis[J]. J Mol Catal A-Chem, 2000, 163(1/2): 19–26.
DOI:10.1016/S1381-1169(00)00396-4 |
[12] |
Highly efficient Ru/MgO-CeO2 catalyst for ammonia synthesis[J]. Catal Commun, 2010, 12(4): 251–254.
DOI:10.1016/j.catcom.2010.09.024 |
[13] |
Efficient ammonia synthesis over a Ru/La0.5Ce0.5O1.75 catalyst pre-reduced at high temperature[J]. Chem Sci, 2018, 9(8): 2230–2237.
DOI:10.1039/C7SC05343F |
[14] |
Crystallographic dependence of CO activation on cobalt catalysts: HCP versus FCC[J]. J Am Chem Soc, 2013, 135(44): 16284–16287.
DOI:10.1021/ja408521w |
[15] |
Discovery of face-centered-cubic ruthenium nanoparticles: Facile size-controlled synthesis using the chemical reduction method[J]. J Am Chem Soc, 2013, 135(15): 5493–5496.
DOI:10.1021/ja311261s |
[16] |
Effect of Ru crystal phase on the catalytic activity of hydrolytic dehydrogenation of ammonia borane[J]. J Power Sources, 2018, 396: 148–154.
DOI:10.1016/j.jpowsour.2018.06.028 |
[17] |
Chemical insights into the design and development of face-centered cubic ruthenium catalysts for Fischer–Tropsch synthesis[J]. J Am Chem Soc, 2017, 139(6): 2267–2276.
DOI:10.1021/jacs.6b10375 |
[18] |
Study of potassium promoter effect for Ru/AC catalysts for ammonia synthesis[J]. Catal Sci Technol, 2013, 3(5): 1367–1374.
DOI:10.1039/c3cy20830c |
[19] |
Ammonia synthesis using a stable electride as an electron donor and reversible hydrogen store[J]. Nat Chem, 2012, 4(11): 934–940.
DOI:10.1038/nchem.1476 |
[20] |
Synthesis and characterization of Ru cubic nanocages with a face-centered cubic structure by templating with Pd nanocubes[J]. Nano Lett, 2016, 16(8): 5310–5317.
DOI:10.1021/acs.nanolett.6b02795 |
[21] |
Facile synthesis of Ru-based octahedral nanocages with ultrathin walls in a face-centered cubic structure[J]. Chem Mater, 2017, 29(21): 9227–9237.
DOI:10.1021/acs.chemmater.7b03092 |
[22] |
Synthesis of Ru icosahedral nanocages with a face-centered-cubic structure and evaluation of their catalytic properties[J]. ACS Catal, 2018, 8(8): 6948–6960.
DOI:10.1021/acscatal.8b00910 |
[23] |
Ru octahedral nanocrystals with a face-centered cubic structure, {111} facets, thermal stability up to 400 ℃, and enhanced catalytic activity[J]. J Am Chem Soc, 2019, 141(17): 7028–7036.
DOI:10.1021/jacs.9b01640 |
[24] |
Investigation of selective chemisorption of fcc and hcp Ru nanoparticles using X-ray photoelectron spectroscopy analysis[J]. J Catal, 2019, 380: 247–253.
DOI:10.1016/j.jcat.2019.10.004 |
[25] |
Influence of CeO2 supports prepared with different precipitants over Ru/CeO2 catalysts for ammonia synthesis[J]. Solid State Sci, 2020, 99: 105983.
DOI:10.1016/j.solidstatesciences.2019.105983 |
[26] |
A novel synergetic effect between Ru and CeO2 nanoparticles leads to highly efficient photothermocatalytic CO2 reduction by CH4 with excellent coking resistance[J]. Solar RRL, 2022, 6(6): 2101064.
DOI:10.1002/solr.202101064 |
[27] |
Structure-dependent electronic properties of nanocrystalline cerium oxide films[J]. Phys Rev B, 2003, 68(3): 035104.
DOI:10.1103/PhysRevB.68.035104 |
[28] |
New insights into the support morphology-dependent ammonia synthesis activity of Ru/CeO2 catalysts[J]. Catal Sci Technol, 2017, 7(1): 191–199.
DOI:10.1039/C6CY02089E |
[29] |
Effect of La2O3 on Ru/CeO2-La2O3 catalyst for ammonia synthesis[J]. Catal Lett, 2009, 133(3/4): 382–387.
DOI:10.1007/s10562-009-0177-7 |
[30] |
Highly durable Ru catalysts supported on CeO2 nanocomposites for CO2 methanation[J]. Appl Catal A-Gen, 2019, 577: 35–43.
DOI:10.1016/j.apcata.2019.03.011 |
[31] |
Ru/CeO2 catalysts for combustion of mixture of light hydrocarbons: Effect of preparation method and metal salt precursors[J]. Appl Catal A-Gen, 2018, 549: 161–169.
DOI:10.1016/j.apcata.2017.09.036 |
[32] |
Recent progress in heterogeneous ammonia synthesis[J]. Chinese Sci Bull, 2019, 64(11): 1114–1128.
DOI:10.1360/N972019-00079 |