You are on page 1of 11

Applied Catalysis B: Environmental 319 (2022) 121958

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


journal homepage: www.elsevier.com/locate/apcatb

Synergistic catalysis of Ru single-atoms and zeolite boosts high-efficiency


hydrogen storage
Lixia Ge a, b, Minghuang Qiu a, Yanfeng Zhu a, b, Shuai Yang c, Wenqian Li a, b, Wanting Li a, b,
Zheng Jiang c, Xinqing Chen a, b, *
a
CAS Key Laboratory of Low-carbon Conversion Science and Engineering, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China
b
University of Chinese Academy of Sciences, Beijing 100049, China
c
Shanghai Synchrotron Radiation Facility, Zhangjiang National Lab, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China

A R T I C L E I N F O A B S T R A C T

Keywords: Liquid organic hydrogen carriers (LOHCs) are promising hydrogen carriers that play an important role in the
Liquid organic hydrogen carriers hydrogen economy. However, designing an efficient catalyst for realizing hydrogen storage with cost-effective
Zeolite and low-temperature is still a great challenge. Herein, we report a Ru single-atoms supported on *BEA zeolite
Hydrogenation
catalyst (Ru(Na)/Beta), with the assistance of hydrogen spillover, which can remarkably enhance the hydro­
Single atom
genation of N-ethylcarbazole(NEC), N-propylcarbazole(NPC) and 2-methylindole (2-MID) at lower temperatures
N-ethylcarbazole
with lower Ru content (0.5 wt%). Notably, the obtained Ru(Na)/Beta catalyst exhibits excellent activity in the
hydrogenation of NEC with the hydrogen uptake of 5.69 wt% and a conversion rate of > 99% within 1.5 h for the
6 MPa H2 at 100 ◦ C, whereas the hydrogen uptake on traditional Ru/Al2O3 is only 2.97 wt% with the conversion
rate of 67 % under the same conditions. It is found that highly dispersed Ru single-atoms boost hydrogen
activation and the strong acid sites (Brønsted and Lewis) of zeolites promote the hydrogen spillover on the
hydrogenation with N-heterocycles. Moreover, the synergistic effect of Ru single atoms and *BEA zeolite is
crucial for accelerating the hydrogenation rate and lowering the activation energy (45.7 vs. 88.3 kJ/mol)
compared with traditional Ru-based catalysts.

1. Introduction up to 5 wt% and high reaction temperatures (130–230 ◦ C). Hence, it is


an urgent issue to explore the feasibility of high-efficiency hydrogena­
Hydrogen energy, as a promising renewable energy, has attracted tion catalysts with low-cost in hydrogen storage under lower
more and more attentions to dealing with the crises of energy resources temperature.
[1–3]. However, the safe and efficient storage and transportation of Recently, atomically dispersed or single-atom catalysts (SACs) have
hydrogen has greatly restricted the development of hydrogen energy. emerged continuously and applied to the research of highly active hy­
Liquid organic hydrogen carriers (LOHCs) are considered one of the drogenation catalysts [17,18]. Besides, the hydrogenation reaction can
critical technologies to solve these problems and achieve carbon be promoted by hydrogen spillover, which provides an idea for
neutrality in decades [4–7]. So far, the most studied LOHCs are N-het­ designing efficient catalysts [19–23]. Researchers show that a frustrated
erocycles, such as N-ethylcarbazole (NEC) [8], N-propylcarbazole (NPC) Lewis pair consisting of a pair of H- and H+ ions is considered to promote
[9], 2-methylindole (2-MID) [10] and dibenzyltoluene (DBT) et. al[11]. hydrogenation for many compounds, such as unsaturated hydrocarbons,
Furthermore, the most commonly used hydrogenation catalysts are Ru, aldehydes, carbon dioxide [24–28]. γ-Al2O3 with sufficient Lewis acidic,
Rh, Pt-based catalysts [12–14]. Particularly, Ru-based catalysts are is a commonly used catalysts support in the hydrogenation of various
widely studied on the hydrogenation of LOHCs, which usually comprise LOHC compounds [12,15,29], while the hydrogen spillover on Al2O3 is
Ru nanoparticles (Ru NPs) on different supports [15,16]. We summa­ relatively slow and confined to a short distance from the metal atoms
rized the state-of-the-art on Ru-based catalysts shown in Table S1. It is [22]. These catalysts are normally confronted with the problems of
found that the current problems are the high loading of Ru metal needed relatively high noble metal loading and low reactivity at lower

* Corresponding author at: CAS Key Laboratory of Low-carbon Conversion Science and Engineering, Shanghai Advanced Research Institute, Chinese Academy of
Sciences, Shanghai 201210, China.
E-mail address: chenxq@sari.ac.cn (X. Chen).

https://doi.org/10.1016/j.apcatb.2022.121958
Received 28 June 2022; Received in revised form 2 September 2022; Accepted 5 September 2022
Available online 7 September 2022
0926-3373/© 2022 Elsevier B.V. All rights reserved.
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

temperature condition. It is well known that zeolites are an ideal support converted to the proton form via the ion exchange method. 1 g of
with large specific surface areas, well-defined microporous and adjust­ calcined zeolite was added into 10 mL 1 M NH4NO3 aqueous solution
able acidity and basicity (Lewis and Brønsted) to immobilize metal with continuous stirring at 80 ◦ C for 2 h. The solid precipitate after
species with high dispersion. A lot of hydroxyl group (-OH) on zeolites centrifugation was dried at 80 ◦ C. The process was repeated three times,
surface can activation and spillover of hydrogen species formed by and the obtained samples was calcined in air at 450 ◦ C for 4 h.
dissociated on the metal [23,30,31]. Currently, zeolite-supported metal
catalysts were widely used in various catalytic reaction, showing supe­ 2.3. Synthesis of catalysts
rior catalytic activity and product selectivity compared with other
supported catalysts [32,33]. For example, SAPO-34(CHA) and FAU ze­ First, 1 g Beta or γ-Al2O3 was dispersed in 20 mL of ethanol, stirred
olites with ultrasmall ruthenium (Ru) clusters could synergistically for 10 min. A certain amount of RuCl3 solution (9.894 mmol/mL) was
activate the AB and water molecules, which greatly promote the added dropwise to the above dispersion and stirred for 3 h at room
hydrogen production rate [34]. Silicalite-1(MFI) zeolite encapsulated Pd temperature, then centrifuge and redispersion into 30 mL of deionized
clusters for Pd@S-1 have an excellent catalytic activity for the FA water. A certain amount of NaOH (0.5 M) solution was added dropwise
dehydrogenation [35]. ZSM-5 (MFI) zeolites are extensively utilized in to the above solution to adjust the PH to 9–10, and stirred for 10 h at
the methane oxidation [36]. However, few studies have been reported so room temperature. Finally, the solution was centrifuged and washed
far on the application of the zeolite-supported metal catalysts for cata­ with deionized water, and placed in a vacuum drying at 60 ◦ C overnight.
lytic hydrogenation of LOHCs such as NEC, NPC, etc. The as-obtained catalysts are named as Ru(Na)/Beta, Ru(Na)/Al2O3
In this article, we reported a Ru single-atoms catalyst supported Beta catalysts.
zeolite by deposition precipitation promoted with alkali cations for the For comparative study, Ru/Beta and Ru/Al2O3 catalysts were pre­
hydrogenation of N-heterocycles. The detailed structures and property pared by the conventional incipient wetness impregnation method,
of catalysts were characterized by STEM, In-situ DRIFTS and EXAFS et.al. dried at 60 ◦ C overnight under a vacuum condition.
The as-obtained Ru(Na)/Beta catalyst, benefiting from the synergistic
effect between Ru single-atoms and *BEA zeolite, realized fast low- 2.4. Catalyst characterizations
temperature hydrogen storage via LOHCs, which is far superior to the
traditional Ru/Al2O3 catalyst. On one hand, the highly dispersed Ru The phase structure of the catalysts was characterized by X-ray
single-atoms enhance the hydrogen activation and hydrogenation; on powder diffraction (XRD, Rigaku Ultima IV) with a Cu Kα rays (λ =
the other hand, the strong acid sites (Brønsted and Lewis) of zeolite 1.54056 Å), operated at 40 kV, 40 mA. The patterns were collected
promote the hydrogen spillover on the surface of the zeolite and hy­ from10–90◦ at a scan speed of 4◦ min− 1. The content of Ru metal in the
drogenation with N-heterocycles. The reaction mechanism and the hy­ catalysts was determined by inductively coupled plasma atomic emis­
drogenation kinetics of N-ethylcarbaozle (NEC) established that the sion spectrometer (ICP-OES) at Optima 8000, Perkin Elmer. The
activation energy (Ea) is lowered on the Ru(Na)/Beta catalyst compared morphology and size of catalysts can be obtained by field emission
to traditional catalysts. Our study combining Ru single-atoms with ze­ scanning electron microscopy (FE-SEM, SUPRA 55 SAPPHIRE, Zeiss)
olites represents a very promising strategy for hydrogen storage in with an acceleration voltage of 5 kV. High resolution transmission
LOHCs. electron microscopy (HR-TEM) measured in JEOL, JEM-2011 with an
acceleration voltage of 200 kV can analyze the morphology of the
2. Experimental section catalyst and the interplanar spacing of specific crystal planes. Energy
dispersive X-ray spectroscopy (EDX) analyzed by Talos F200S. Double
2.1. Chemicals and materials spherical aberration-corrected high-angle annular dark-field scanning
transmission electron microscopy (AC-HAADF-STEM) images were ac­
Tetraethylammonium hydroxide solution (TEAOH, 35 wt%, Alfa quired on a JEM-ARM300F with an accelerating voltage of 300 kV. The
Aesar), sodium aluminate (NaAlO2, Shanghai Richjoint Chemical Re­ texture properties of the catalysts were measured after degassed for 10 h
agents Co, Ltd), colloidal silica (40 wt%, Aldrich), aluminum oxide under vacuum at 300 ◦ C by Micromeritics ASAP 2420 automatic
(99.99 % metals basis, γ phase, 20 nm, Aladdin), lithium hydroxide, adsorption instrument at − 196 ◦ C. H2 temperature-programmed
anhydrous (99.9 % metals basis, Macklin), sodium hydroxide (99.9 % desorption (H2-TPD) experiments were performed on a Micromeritics
metals basis, Macklin), potassium hydroxide (99.9 % metals basis, Chem Sorb 2920 instrument with a flame ionization detector (TCD)
Macklin), caesium hydroxide, (99.9 % metals basis, Macklin), ruthe­ under nitrogen flow. The type of acid sites of catalysts was tested by
nium(III) chloride hydrate(RuCl3⋅xH2O, Ru≥47.0 %, Adamas), N-Eth­ pyridine IR (Py-IR) spectra using an FTIR-650 spectrometer (Bruker,
ylcarbazole (97 %, Aladdin), 2-methylindole(98 %, Aladdin), N- Germany). After degassing under vacuum at 400 ◦ C for 2 h, the sample
propylcarbazole(98 %, Adamas). was saturated with pyridine vapor at room temperature. Before
recording the IR spectrum, the sample was degassed for 30 min at a
2.2. Synthesis of Beta zeolites temperature of 150 ◦ C, 200 ◦ C and 350 ◦ C to remove excess probe
molecules. The number of acid sites was calculated by the formula from
The Beta zeolites were synthesized under conventional hydrothermal the integrated intensities of the characteristic bands (1450 cm− 1 for
conditions at 140 ◦ C for 6 days from the starting gel with the molar Lewis acid sites and 1540 cm− 1 for Brönsted acid sites). X-ray photo­
compositions of 11.8 H2O:0.106NaOH: xAl2O3:1.0SiO2:0.36TEAOH (x electron spectroscopy (XPS) analysis was performed using a Thermo
= 0.01, 0.025, 0.04), the obtained samples were named as Beta-100, Scientific K-alpha spectrometer. The spectrometer has an Al Kα X-ray
Beta-40, Beta-25. Typically, 37.86 g of TEAOH solution and a certain source operating at 1486.6 eV. The near-surface chemical information of
amount of NaAlO2 was mixed under continuous stirring. After the the catalyst samples was analyzed by scanning Al Kα X-rays with a tube
following stirring for 10 min, 1.07 g NaOH dissolved in 6 g deionized current of 3 mA and voltage of 12 kV. Calibration was performed using a
water was added into the solution. 37.55 g of colloidal silica was then C1s binding energy of 284.6 eV.
added into the above mixture and stirred continuously for 6 h. Finally, The X-ray absorption spectra (XAS) of the Ru K-edge was performed
the reaction mixture was transferred into a 100 mL Teflon-lined stain­ on the BL14W1and BL11B beamline of Shanghai Synchrotron Radiation
less-steel autoclave and crystallized at 140 ◦ C for 6 days. The as- Facility (SSRF) with an operating voltage of 3.5 GeV and injection cur­
synthesized solid product was centrifuged, washed with deionized rent of 240 mA. Si(311) double-crystal was used as the monochromator,
water, and then dried at 80 ◦ C, followed by calcination in air at 550 ◦ C and the data was collected using a solid-state detector. The EXAFS raw
for 6 h to remove the organic template. The calcined zeolite was finally data was acquired in standard Lytle ion chamber fluorescence mode

2
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

under ambient atmosphere. Ea


In-situ diffuse reflectance infrared Fourier transform of adsorbed CO lnk = lnA − (5)
RT
(CO-DRIFTS) were performed on a thermal FTIR spectrometer (Nicolet
IS50) equipped with an MCT detector. The samples were pretreated with k is the reaction rate constant; R is universal gas constant (8.314 J/
Argon at 150 ◦ C for 1 h. After the sample was cooled to 100 ◦ C, 20 % CO/ (mol⋅K)); T is reaction temperature(K); Ea is apparent activation energy
Ar was passed into the chamber at a flow rate of 40 mL/min for (J/mol).
approximately 20 min until saturation. Next, the samples were purged
with Ar at a flow rate of 40 mL/min for 10 min to remove excess CO. All 3. Results and discussion
data were acquired over a range of 4000 − 900 cm− 1 with a resolution of
4 cm− 1, with 64 scans at 100 ◦ C. Hydrogen-deuterium exchange (H-D 3.1. Characterization of Catalysts
exchange) were also measured on a thermal FTIR spectrometer (Nicolet
IS50) equipped with an MCT detector. The catalyst was pretreated with In this study, different Beta zeolites (Si/Al ratio=25–100) were
Argon at 150 ◦ C for 1 h and cooled down to 100 ◦ C. Then the reaction synthesized under hydrothermal conditions. Then, various Ru-based
gases of H2 and D2 were pulsed into the chamber intermittently. catalysts were prepared by deposition precipitation (DP) and incipient
wetness impregnation (IWI) of RuCl3 on Beta zeolites or commercial
2.5. Catalytic tests γ-Al2O3, respectively. The obtained catalysts are named as Ru(Na)/Beta,
Ru(Na)/Al2O3, Ru/Beta, Ru/Al2O3 and the details are shown in the
The catalysts were not reduced before applied to catalytic perfor­ supporting materials. The powder X-ray diffraction (XRD) patterns of all
mance tests. In typical Liquid Organic Hydrogen Carriers (LOHCs) hy­ Beta zeolites with different Si/Al ratio exhibit typical diffraction peaks
drogenation system, the reactants N-ethylcarbazole (10 g, NEC) and the corresponding to the *BEA framework (Fig. S1). No diffraction peaks of
catalysts were performed in a stainless-steel autoclave reactor (100 mL, Ru species could be detected in the obtained Ru-based catalysts due to
SLM microform, Beijing Century Senlong experimental apparatus Co., small metal loading and uniform metal dispersion (Fig. 1a). ICP-OES
Ltd.), where the mass ratio of catalysts to the substrate can be adjusted to measurements show that the Ru contents in Ru(Na)/Beta, Ru(Na)/
a desired value (Cat/NEC=x, x represent the ratio). The reactor was Al2O3, Ru/Beta and Ru/Al2O3 are 0.42, 0.43, 0.45, 0.47 wt%, respec­
purged with H2 (99.99% H2) and pressurized to a certain pressure (4 tively (Table S2). N2 adsorption and desorption isotherms describe that
MPa, 5 MPa, 6 MPa), maintaining this state in the entire reaction. After the specific surface areas of Beta (844.41 m2/g) are much higher than
the reactor reached the desired temperature, the reaction was started Al2O3 (113.95 m2/g) (Figs. S2-S4 and Tables S2-S3). Studies have shown
together with the stirring of 600 rpm for a desired time. After the re­ that the large specific surface area of the support is also a very important
action was finished and naturally cooled to the room temperature, the factor in favor of the formation of high metal dispersion. The
product (0.1 g) was dissolved in ethanol (5 mL) for analysis using GC morphology and size of the support did not change greatly after loading
chromatograph. The detailed analysis conditions and instrument pa­ Ru species, no matter for Beta or Al2O3(Fig. S5). High resolution trans­
rameters are provided in the Supplementary Information. The compo­ mission electron microscopy (HRTEM) images of Ru(Na)/Beta and Ru
sition and concentration of the hydrogenation products calculated from (Na)/Al2O3 (Fig. 1c,d) show the interplanar spacing of 1.07 nm and
the separated peaks area of 12 H-NEC, 8 H-NEC, 4 H-NEC and NEC in GC 0.197 nm are ascribed to the (100) lattice plane of Beta and the (400)
chromatograms. The amount of hydrogen storage was calculated by the lattice plane of Al2O3. No identifiable Ru nanoparticles (NPs) on the
following equation: surface Ru(Na)/Beta and Ru(Na)/Al2O3 could be observed, indicating
that the Ru species might be dispersed in ultrasmall metal clusters or
H2 storage content (wt%) = 5⋅8 ×{(12 H-NEC [mol] ×6 H2) + (8 H-NEC single-atom level. The Cs-corrected high angle annular dark field scan­
[mol] × 4 H2)+ (4 H-NEC [mol]×2 H2)}/{( Initial NEC) [mol]×6 H2} (1) ning transmission electron microscopy (HAADF-STEM) images of Ru
(Na)/Beta are shown in Fig. 1e, displayed the presence of isolated Ru
atoms. The corresponding EDX element mapping images (Fig. 1f) further
confirmed the high dispersion of single atom Ru over the surface of Beta.
2.6. Hydrogenation kinetics determination Besides, the Ru metal dispersion of Ru-based catalysts results (Table S2)
also indicate a high dispersion in the Ru(Na)/Beta and Ru(Na)/Al2O3
Since this reaction is an liquid organic compound catalytic hydro­ catalysts.
genation reaction, the stirring speed is over 600 r/min, and the catalyst The acidity of samples was obtained by NH3-TPD and Py-IR spec­
particles are nanoscale, thus the influence of both external and internal troscopy at desorption temperatures of 150, 200 and 350 ◦ C (Fig. 1b,
diffusion on the hydrogenation reaction were basically eliminated [37]. Figs. S6-S9 and Tables S4-S6). It is noticed that the pyridine adsorption
Therefore, the NEC hydrogenation reaction can be regarded as a peak at 1450 and 1540 cm− 1 are denoted as the Lewis and Brønsted acid
zero-order reaction of H2 concentration and a first-order reaction of NEC sites, respectively.[38] Among all samples, the Lewis acid sites of the
concentration. Here, the reaction rate of the hydrogenation reaction of catalysts with Beta zeolite as the support are much higher than that of
N-ethylcarbazole over Ru(Na)/Beta and Ru(Na)/Al2O3 catalysts is esti­ the catalysts prepared with Al2O3 as the support, which is over 55-fold
mated from the following rate equation: higher than Al2O3 at 350 ◦ C. Due to the nature of Al2O3, there are no
dCNEC Brønsted acid sites in Ru(Na)/Al2O3 and Ru/Al2O3. The X-ray photo­
rNEC = = − kCNEC (2) electron spectroscopy (XPS) analyses of the Ru 3p3/2 for the prepared
dt
samples were also shown in Fig. S10 and Table S7. We can clearly see
( )
CNEC that only one characteristic peak at 463.38 eV for Ru(Na)/Beta and Ru
ln = − kt (3)
C0 (Na)/Al2O3, corresponding to the oxidation state of ruthenium metal
close to Ru(III) or Ru(IV) (Ruδ+, ca. 463.2–463.9 eV). Meanwhile, the
In addition, Arrhenius equation were used to calculate the apparent
ruthenium species of about 90.7 % and 96.0 % in Ru(Na)/Beta-spent and
activation energy of hydrogenation reaction of N-ethylcarbazole on
Ru(Na)/Al2O3-spent still maintained an the oxidation states, and a very
catalysts:
small amount of ruthenium species shifts to a smaller value with a
where C0 and CNEC are the initial reactant concentration and con­
ruthenium 3p3/2 peak at 461.28 eV. These results indicate that the Ru
centration at time t, respectively;
species of Ru(Na)/Beta and Ru(Na)/Al2O3 could maintain the oxidation
k = A • e−
Ea
ET (4) state during the hydrogenation reaction. Whereas Fig. S11 shows a peak
for Ru/Beta and Ru/Al2O3 at ~462.7 eV. Since the peak of metal

3
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

Fig. 1. (a) XRD patterns. (b) Py-IR spectra of all samples. (c) HRTEM images of Ru(Na)/Beta and (d) Ru(Na)/Al2O3. (e) representative aberration-corrected HAADF-
STEM image of Ru(Na)/Beta and (f) the corresponding EDX element mapping results.

Fig. 2. (a) Ru K-edge XANES spectra and (b) FT k3-weighted Ru K-edge EXAFS spectra. EXAFS fitting curve of (c) Ru(Na)/Beta. Wavelet transform of (d) Ru foil, (e)
RuO2, (f) Ru(Na)/Beta.

4
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

ruthenium is at ~461 eV and the peak of oxidized ruthenium is at (In-situ DRIFTS) was employed to study the chemisorption of CO mole­
~463 eV, the valence state of ruthenium is likely to the coexistence of cules as probes on different Ru-based catalysts at 100 ◦ C. As presented in
metal ruthenium and oxidized ruthenium on the Ru/Beta and Ru/Al2O3 Fig. 3a, the CO DRIFTS spectra of Ru(Na)/Beta showed two CO-
catalyst.[39,40]. adsorption (COads) peaks at 2132 and 2082 cm-1, respectively. Simi­
X-ray absorption spectroscopy (XAS) were conducted to investigate larly, Ru(Na)/Al2O3 also manifests two peaks at 2127 and 2067 cm-1,
the electronic and structural properties of Ru-based catalysts at atomic- which could be assigned to the linearly adsorbed of CO adsorption on Ru
level. Fig. 2a present the X-ray absorption near edge spectra (XANES) of atomic sites [43,44]. For Ru/Beta and Ru/Al2O3, the COads peaks of
the prepared catalysts. Notably, the X-ray absorption near edge spectra 2154 and 2122 cm-1 are also assigned to linearly adsorbed of CO on
(XANES) of all the samples shows that the absorption edges of Ru sup­ atomic dispersion Ru sites; the 2092 and 2030 cm-1 peaks of Ru/Beta
ported samples is higher than that of Ru foil reference and close to the and the 2067 cm-1 peak of Ru/Al2O3 can be assigned to the linearly
Ru(OH)3 and RuO2, revealing the oxidized state Ru species with adsorbed of CO on the rims of Ru clusters. However, a broad band at
chemical state close to Ru3+or Ru4+, which is consistent with XPS re­ 1979 cm-1 of Ru/Al2O3 tends to adsorb CO over bridge or hollow sites of
sults. Besides, the analyze of linear combination fitting (LCF, Fig. S12, Ru clusters [3,35]. These comparison results demonstrated the existence
Table S8) also confirms this. The Fourier-transform extended X-ray ab­ of isolated Ru atoms in Ru(Na)/Beta and Ru(Na)/Al2O3, whereas the Ru
sorption fine structure (FT-EXAFS) spectra and the fitting structural clusters are the main form in Ru/Beta and Ru/Al2O3, which is consistent
parameters of EXAFS spectrum of different catalysts are shown in with the STEM and EXAFS results.
Fig. 2b,c, Figs. S13–14 and Table 1, respectively. The FT-EXAFS spectra The H2 temperature programmed desorption (H2-TPD) data (Fig. 3b)
(Fig. 2b) of Ru(Na)/Beta and Ru(Na)/Al2O3 feature a prominent peak at revealed that the H2 desorption peak of Al2O3-support catalyst at the
~1.5 Å from the first coordination shell of Ru related to the Ru–O temperatures of below 120 ◦ C, which is attributed to homolytically
scattering, and the absence of Ru-Ru peak indicated that the presence of adsorbed H2 species on the Ru single-atoms. Differently, Beta-support
isolated Ru atoms in the samples. The average coordination number and catalysts show a high-temperature H2 desorption peak (~155 ◦ C),
bond length of Ru-O of Ru(Na)/Beta and Ru(Na)/Al2O3 were calculated which are likely to result from the heterolytically adsorbed H2 species on
to be 4.7 ± 0.7, 2.00 ± 0.01 Å and 4.4 ± 0.6, 1.99 ± 0.01 Å, respec­ catalysts. This heterolytically adsorption of H2 over Beta-support cata­
tively. For Ru/Beta and Ru/Al2O3, apart from the peak at ~1.5 Å, an lysts means the stronger adsorption of hydrogen species than Al2O3-
additional peak at ~2.85 Å associated with Ru-Ru emerges, indicating support catalysts by homolytically adsorbed [45]. Moreover, the
the formation of Ru metal clusters.[41] In addition, the Ru-Ru average adsorption strength of H2 on catalysts prepared by the deposition pre­
coordination number and mean bond length were 8.4 ± 5.2, 3.40 cipitation method was significantly higher than those of the catalysts
± 0.02 Å and 8.5 ± 3.9, 3.42 ± 0.02 Å, respectively. Compared with the prepared by the incipient wetness impregnation method, which may be
Ru foil, the relatively low Ru-Ru coordination number indicates the attributed to the high dispersion of Ru atoms on the catalyst surface
presence of Ru clusters. Besides, the longer Ru-Ru distance of Ru/Beta [46]. Besides, the hydrogen spillover of the catalyst prepared with Beta
and Ru/Al2O3 imply the loose Ru-Ru bonding.[42] As illustrated in zeolite is much higher than that of the catalyst prepared with Al2O3
Fig. 2d-f and Fig. S15, the detailed wavelet transform extended X-ray support (>200 ◦ C), which means Beta zeolite can significantly promote
absorption fine structure (WT-EXAFS) was analyzed to examine the the transfer of activated hydrogen from Ru species to the zeolite
dispersion of Ru species. Only one maximum peak at ~1.3 Å was compared to Al2O3[30].
observed for Ru(Na)/Beta and Ru(Na)/Al2O3, which is attributed to the In situ FTIR spectroscopy further proved that the hydrogen activation
Ru-O coordination. For Ru/Beta and Ru/Al2O3, another maximum peak mechanism on the catalysts by the H-D isotope labeling experiment. As
at ~3.0 Å appeared, corresponding to Ru-Ru contribution. In summary, shown in Fig. 4a, the IR bands at 3745, 3688, 3652 cm-1 are assigned to
these data demonstrate that the exist of single dispersed Ru atoms over the different stretching mode of hydroxyls groups(-OH) observed upon
Ru(Na)/Beta and Ru(Na)/Al2O3, whereas Ru nanoclusters over Ru/Beta H2 activation on different catalysts at 100 ◦ C, indicating the combina­
and Ru/Al2O3. tion of protons with the hydroxyl groups at the zeolites and alumina
In-situ diffuse reflectance infrared Fourier transform spectroscopy surface. Meanwhile, the strong IR band at ~1632 cm-1 corresponding to
the Ru-H species, which clearly proves that the heterolytic dissociation
of hydrogen on the Ru(Na)/Beta catalyst [47,48]. In the H/D exchange
Table 1 experiments (Fig. 4b), abundant surface –ODx at 2750 and 2610 cm-1
EXAFS fitting parameters at the Ru K-edge for various samples (Ѕ20 =0.70). and Ru-D at ~1439 cm-1 were found, especially a strong intensity of
Samples Shell Na R (Å)b σ2× 103 ΔE0 R Ru-D on Ru(Na)/Beta was observed [17,46]. However, there are only a
(Å2)c (eV)d factor weak peak displayed at ~3750 cm-1 in Beta zeolite in the H− D isotope
Ru foil Ru- 12 * 2.67 3.1 ± 0.6 3.2 0.009 labeling experiment, which is attributed to the interaction of hydrogen
Ru ± 0.01 ± 0.9 with hydroxyl groups on the surface of zeolite, and no characteristic
RuO2 Ru-O 6.0 1.96 3.9 ± 1.2 1.6 0.011 peaks were found in Al2O3. That is, the Ru(Na)/Beta catalyst with Ru
± 0.5 ± 0.01 ± 1.0
single-atoms dispersion and more Brønsted and Lewis acid sites has
Ru- 4.6 3.14 10.8 ± 4.1 1.5
Ru ± 2.0 ± 0.01 ± 2.2
stronger heterolytic dissociation and hydrogen activation. The ther­
Ru(OH)3 Ru-O 6.1 2.01 10.3 ± 2.9 3.0 0.014 mogravimetric analysis (Fig. S16) demonstrated that the weight loss
± 1.3 ± 0.02 ± 2.6 (3.5 wt%) of Ru(Na)/Al2O3 and Ru(Na)/Beta likely caused by the
Ru(Na)/ Ru-O 4.7 2.00 9.1 ± 2.7 -1.7 0.014 dehydration of surface –OHx species.
Beta ± 0.7 ± 0.01 ± 1.6
Ru(Na)/ Ru-O 4.4 1.99 1.5 ± 1.4 -0.4 0.019
Al2O3 ± 0.6 ± 0.01 ± 1.6 3.2. Catalytic performances
Ru/Beta Ru-O 5.4 1.97 1.2 ± 2.1 -2.0 0.016
± 0.9 ± 0.01 ± 2.2 The catalytic performance of different catalysts in hydrogenation of
Ru- 8.4 3.40 4.6 ± 4.2 -8.7
N-ethylcarbazole (NEC) was investigated as a model reaction under
Ru ± 5.2 ± 0.02 ± 1.3
Ru/Al2O3 Ru-O 5.4 1.96 1.3 ± 2.1 -0.5 0.015 relatively mild conditions. The products for hydrogenation of NEC were
± 1.0 ± 0.01 ± 2.1 analyzed by GC-MS and GC. Fig. S17 describes the time-dependent
Ru- 8.5 3.42 3.6 ± 0.1 -11.7 products of hydrogen uptake of NEC on the Ru(Na)/Beta catalyst. We
Ru ± 3.9 ± 0.02 ± 1.4 can clearly see that the hydrogenation process of NEC is finished
[a]
N: coordination numbers; [b] R: bond distance; [c] σ2: Debye-Waller factors; [d] through three main stages, and the intermediates and products mainly
ΔE0: the inner potential correction. R factor: goodness of fit. include tetrahydro-N-ethylcarbazole(4 H-NEC), octahydro-

5
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

Fig. 3. (a) In-situ DRIFTS CO chemisorption (100 ℃) and (b) H2-TPD profiles of different catalysts.

Fig. 4. Dihydrogen activation on catalysts (a) in-situ FTIR spectra of H2 activation and (b) D2 activation on Ru(Na)/Beta, Ru/Beta, Beta, Ru(Na)/Al2O3, Ru/Al2O3
and Al2O3 recorded at 100 ℃.

ethylcarbazole (8 H-NEC) and dodeca-hydroethylcarbazole(12 H-NEC). 98.28 % of the theoretical hydrogen storage of NEC. Interestingly, the
However, the intermediates of 6 H-NEC and 10 H-NEC is hardly found catalytic hydrogenation activity of Ru(Na)/Al2O3 prepared by the
by GC analysis.[49] In the initial stage of the hydrogenation reaction, deposition precipitation method was improved compared with
the concentration of NEC decreased rapidly and was hydrogenated to Ru/Al2O3, the conversion rate increased to 97.24 %, and the hydrogen
4 H-NEC, while 4 H-NEC was continuously hydrogenated to 8 H-NEC storage capacity increased to 4.12 wt%. The experimental results show
and 12 H-NEC. In a word, the hydrogenation process is a continuous that the atomically dispersed Ru and more Lewis acid sites (seen in
reaction. During the reaction, the 4 H-NEC concentration reached the HAADF-STEM and Py-IR) may be attributed to improve the catalytic
highest value at about 10 min. When the reaction time was prolonged to activity. Compared to Ru/Al2O3, Ru/Beta also has better performance in
40 min, the concentration of 8 H-NEC peaked and started to decrease. the hydrogenation of NEC, the conversion rate of 99.27% and the
Finally, NEC and intermediates are almost completely converted to hydrogen storage capacity of 5.41 wt%, which is slightly lower than Ru
12 H-NEC after 90 min. Therefore, the hydrogenation reaction pathway (Na)/Beta. These results demonstrated that Ru(Na)/Beta with high
of NEC can be described as a stepwise saturation process of two double specific surface area, higher number of acid sites and atomic dispersion
bonds, resulting in stepwise production of intermediates 4 H-NEC, of Ru atoms has high catalytic activity in the hydrogenation of NEC.
8 H-NEC and 12 H-NEC with three isomers of perhydrogen compounds Furthermore, it is found that Beta zeolite (Si/Al ratio of 40) as a support
(Fig. 5a). These results are consistent with the results of Eblagon and for the preparation of catalysts, which has better catalytic performance
co-workers.[49]. in 100 ◦ C, 6 MPa than other Beta zeolites with different Si/Al ratio in
Ru/Al2O3 is the most commonly used catalyst in the catalytic hy­ Fig. S18. The reason is possibly that Ru(Na)/Beta (Si/Al ratio of 40) has
drogenation of N-ethylcarbazole, and the calculation of theoretical suitable acidity compared to the Ru(Na)/Beta (Si/Al ratio of 25 and 100)
hydrogen storage of NEC is 5.80 wt%.[13] Fig. 5b shows that the cata­ from Fig. S8 and Table S4&S5. the When Beta zeolite (Si/Al ratio of 40)
lytic performance of Ru/Al2O3 afforded the conversion of 66.95 %, the was used as the support, we explored the effects of different metal
12 H-NEC yield of 40.36% and the hydrogen storage capacity of 2.97 wt loadings and other reaction conditions on the catalytic hydrogenation
%, which is only 51.20 % of the theoretical hydrogen storage of NEC. In performFance of NEC. Fig. S19 shows that the full hydrogenation can
sharp contrast, the obatined Ru(Na)/Beta exhibited the excellent per­ almost achieved at metal loading of 0.5 wt%. When the metal loading
formance with the conversion of 99.19 %, the 12 H-NEC yield of 96.62% decreased to 0.2 wt%, the NEC conversion rate is only 34.24%, and the
and the hydrogen storage capacity of 5.69 wt%, which is as high as hydrogen storage capacity is only 1.11 wt%. However, the catalytic

6
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

Fig. 5. (a) The hydrogenation process and intermediates and products of N-ethylcarbazole (NEC). (b) Catalytic performances of as-prepared catalysts (100 ◦ C,
6 MPa H2, Cat/NEC=0.1). (c) Ru(Na)/Beta at different reaction temperature(T = 80–120 ◦ C, 6 MPa H2, Cat/NEC=0.1). (d) Comparison of catalytic performance for
the hydrogenation of N-ethylcarbazole (NEC) with various catalysts. (e) The catalytic performance of Ru(Na)/Beta catalyst for hydrogenation of different LOHCs.

activity of 1.0Ru(Na)/Beta began to decrease, the hydrogen storage (Na)/Beta exhibited outstanding catalytic performance on the hydro­
capacity from 5.69 wt% to 5.57 wt%, which may be due to the fact that genation of NEC, compared with the catalysts prepared by other zeolites
some Ru nanoparticles began to aggregate. Different bases as pre­ as the supports, such as MCM-41, Naβ, ZSM-5, SAPO-11. Moreover,
cipitants (such as K, Li, Cs, et.al) were used prepare catalysts, it is found pristine Beta zeolite without Ru loading hardly catalyze the NEC hy­
that there is no obvious effect on the catalytic performance of Ru drogenation reaction (0% yield). In a word, Ru(Na)/Beta (0.5 wt% of
(X)/Beta in hydrogenation of NEC (Table S9). Next, we used the Ru Ru) catalyst obtained in this work achieves the goal of low-temperature
(Na)/Beta catalyst to study the effect of different reaction conditions on and efficient hydrogen storage, which has higher economic value than
the catalytic hydrogenation performance of NEC (Fig. 5c, S20–21). It is the previous studies (Table S1, Fig. 5d). In order to evaluate the recycle
clearly seen that the catalytic activity of Ru(Na)/Beta increases rapidly stability of the catalyst of Ru(Na)/Beta for the hydrogenation of NEC,
before 100 ◦ C, and the catalytic activity is almost unchanged when the the recycle hydrogenation results are shown in Fig. S23. After three
temperature is higher than 100 ◦ C, with complete hydrogenation of NEC recycles over Ru(Na)/Beta, the conversion rate of NEC still reach 99.63
(100% yield) at 110 ◦ C and 120 ◦ C. The hydrogen pressure and catalyst %, the yield of 12-NEC is 89.02%, and the hydrogen storage capacity is
ratio have a greater impact on the hydrogenation performance of NEC, 5.57 wt% at the reaction condition of 100 ◦ C, 6 MPa H2. This result
and higher pressure and catalyst mass ratio are more conducive to the shows that the prepared catalyst Ru(Na)/Beta has a good stability and
hydrogenation of NEC. When the hydrogen pressure is higher than recyclability.
6 MPa, the catalytic performance hardly changes, so we explore the Furthermore, the hydrogenation performance of other LOHCs (N-
hydrogenation performance of different catalysts under 6 MPa H2. In propylcarbazole and 2-methylindole) over Ru(Na)/Beta were explored.
addition, we explored the catalytic performance of catalysts prepared N-propylcarbazole (NPC) is also a promising candidate for LOHCs,
with different zeolites as supports for the hydrogenation of NEC at which has a theoretical hydrogen storage capacity of 5.43 wt%. The
100 ◦ C and 6 MPa H2. As shown in Fig. S22 and Table S10,Ru experimental results indicate that Ru(Na)/Beta catalyst can achieve

7
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

almost complete hydrogenation (100 % yield) within 1 h under 100 ◦ C The catalytic activity of heterogeneous catalysts is closely related to
and 6 MPa H2 (Fig. 5e). Apparently, the catalytic activity of Ru(Na)/ the interaction between the reactants/intermediates and the catalyst
Beta catalyst for hydrogenation of NPC was further improved than the surface according to the Sabatier principle [51]. Thus, all intermediates
previously reported catalysts(Table S1).[50] In addition, the and product distribution with reaction time at different temperatures
as-prepared Ru(Na)/Beta catalyst were also used for 2-methylindole under the same conditions were investigated in this work between Ru
(2-MID) hydrogenation under 120 ◦ C and 6 MPa H2. As shown in (Na)/Beta and Ru(Na)/Al2O3. Fig. 6 show that the hydrogenation per­
Fig. 5e, the conversion of 2-MID reached 98.96% over Ru(Na)/Beta after formance of the two catalysts increases gradually with the increase of
2 h, and the hydrogen uptake was 5.58 wt%, which is comparable to the temperature. Particularly, Ru(Na)/Beta shows higher catalytic activity
hydrogenation performance of 5 wt% Ru/Al2O3 under 160 ◦ C and compared with Ru(Na)/Al2O3 at low temperature (90 ◦ C), the conver­
7 MPa H2 [10]. In short, the Ru(Na)/Beta catalyst exhibits high catalytic sion of NEC is 96.51 % vs. 28.66 %. This means that the hydrogen
activity in catalyzing various LOHCs hydrogenation reactions. storage of NEC at low reaction temperature is more efficient over Ru

Fig. 6. Time-dependent product distribution in the hydrogenation of NEC over (a–c) Ru(Na)/Beta and (d–f) Ru(Na)/Al2O3. (Cat/NEC=0.1, 6 MPa H2 and tem­
perature of (a,d) 90 ℃, (b,e) 100 ℃, and (c,f) 120 ℃.

8
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

(Na)/Beta than Ru(Na)/Al2O3. Both catalysts can achieve almost com­ and suitable acid sites could be sufficient to utilize hydrogen spillover to
plete conversion of NEC (>97 %) at the temperature of 100–120 ◦ C. promoting the hydrogenation of NEC.
However, the Ru(Na)/Beta catalyst can also simultaneously achieve
almost complete conversion of the reactants(NEC) to the perhydro 4. Conclusions
product (12 H-NEC) within 90 min (yield >96.62 %). According to the
measured conversion of NEC, the first-order rate constant (k1) under In summary, Ru single-atoms catalyst supported on *BEA zeolite
different temperature could be obtained (Fig. 7). Ru(Na)/Beta exhibited with suitable acidities was firstly designed and prepared for hydrogen
3.1–11.1-fold higher k1 value than Ru(Na)/Al2O3 at the temperature of storage on liquid organic hydrogen carriers (LOHCs) (mainly, NEC). As
90–120 ◦ C. Meanwhile, the apparent activation energy Ea of hydroge­ compared with conventional Ru/Al2O3 catalysts, the single-atoms
nation of NEC over Ru(Na)/Beta and Ru(Na)/Al2O3 catalysts can be dispersed Ru sites and adjacent acid sites of *BEA zeolite could syner­
obtained by calculation which were 45.7 and 88.3 kJ/mol, respectively. gistically activate the hydrogen and fast promote the hydrogenation of
Therefore, Ru(Na)/Beta is obviously superior to Ru(Na)/Al2O3 in the NEC at lower temperature. The obtained Ru(Na)/Beta catalyst exhibited
hydrogenation of NEC, especially at low temperature. superior hydrogen uptake of 5.69 wt% with a conversion rate of > 99%
Based on these characterizations and previous studies,[52] the in 1.5 h at 100 ℃, which is 98.28% of the theoretical hydrogen storage
mechanism of the hydrogenation of NEC over the Ru(Na)/Beta are of NEC. The hydrogen capacity of Ru(Na)/Beta are much higher than
proposed in Scheme 1a. The reaction occurs mainly through two path­ that Ru-based and other metal-based heterogeneous catalysts reported
ways:(1) N-Ethylcarbazole molecules can be directly adsorbed and react under the same conditions. With the help of various characterization
with H2 on the Ruδ+ surface; (2) the activated H* radical generated by technologies, it is found that single-atoms dispersed Ru(Na)/Beta cata­
H2 dissociation on Ruδ+ surface reacts with N-Ethylcarbazole molecules lysts show a higher heterolytically activation H2 species capacity and
on metal-support interface or integrated with OH species on the surface strength of hydrogen spillover. The kinetics results of these catalysts
of Beta zeolite. In this way, Beta zeolite acts as a receiver and transferrer revealed that the hydrogenation activation energies of NEC over Ru
of H* radicals, promoting the role of hydrogen spillover and increasing (Na)/Beta can be significantly reduced. Consequently, our strategy
the rate of hydrogenation on the surface of zeolite. As shown in Scheme provides insights towards the synergetic catalysis of zeolite-supported
1b, abundant Si-OH groups on the Beta surface (including surface de­ metal catalysts for fast hydrogen storage into aromatic LOHCs under
fects and Brønsted acids) play an important role in anchoring and mild conditions.
transferring H* radicals in the hydrogen spillover process. In addition,
the surface area of support would be highly important due to the larger Supporting Information
molecular dynamics size of NEC, and the catalytic reaction via hydrogen
spillover can only take place on the surface of the support. In the case of Additional characterization of materials, catalytic hydrogenation
Ru(Na)/Beta, Beta zeolite possesses extremely high external surface area results of NEC, and references.

Fig. 7. Calculation of the k1 values for (a) Ru(Na)/Beta and (b) Ru(Na)/Al2O3; Arrhenius plot to calculate the Ea values using the estimated k1 values for (c) Ru(Na)/
Beta and (d) Ru(Na)/Al2O3.

9
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

Scheme 1. (a) The hydrogenation pathways of NEC and (b) mechanism of transfer of H species on Ru(Na)/Beta.

CRediT authorship contribution statement References

Lixia Ge: Formal analysis, Investigation, Methodology, Writing - [1] L. Schlapbach, A. Züttel, Hydrogen-storage materials for mobile applications,
Nature 414 (2001) 353–358.
original draft, Writing - review & editing. Minghuang Qiu: Formal [2] U. Eberle, M. Felderhoff, F. Schüth, Chemical and physical solutions for hydrogen
analysis, Methodology. Yanfeng Zhu: Investigation. Wenqian Li: storage, Angew. Chem. Int. Ed. 48 (2009) 6608–6630.
Investigation. Xinqing Chen: Supervision, Methodology, Funding [3] C. Dong, Z. Gao, Y. Li, M. Peng, M. Wang, Y. Xu, C. Li, M. Xu, Y. Deng, X. Qin,
F. Huang, X. Wei, Y.-G. Wang, H. Liu, W. Zhou, D. Ma, Fully exposed palladium
acquisition, Writing - review & editing. cluster catalysts enable hydrogen production from nitrogen heterocycles, Nat.
Catal. (2022).
[4] E. Gianotti, M. Taillades-Jacquin, J. Rozière, D.J. Jones, High-purity hydrogen
Declaration of Competing Interest generation via dehydrogenation of organic carriers: a review on the catalytic
process, ACS Catal. 8 (2018) 4660–4680.
[5] Y.-Q. Zou, N. von Wolff, A. Anaby, Y. Xie, D. Milstein, Ethylene glycol as an
The authors declare that they have no known competing financial efficient and reversible liquid-organic hydrogen carrier, Nat. Catal. 2 (2019)
interests or personal relationships that could have appeared to influence 415–422.
the work reported in this paper. [6] M. Reuß, T. Grube, M. Robinius, P. Preuster, P. Wasserscheid, D. Stolten, Seasonal
storage and alternative carriers: a flexible hydrogen supply chain model, Appl.
Energy 200 (2017) 290–302.
Data Availability [7] D. Teichmann, W. Arlt, P. Wasserscheid, R. Freymann, A future energy supply
based on liquid organic hydrogen carriers (LOHC), Energy Environ. Sci. 4 (2011)
Data will be made available on request. 2767–2773.
[8] S. Fei, B. Han, L. Li, P. Mei, T. Zhu, M. Yang, H. Cheng, A study on the catalytic
hydrogenation of N-ethylcarbazole on the mesoporous Pd/MoO3 catalyst, Int. J.
Acknowledgements Hydrog. Energy 42 (2017) 25942–25950.
[9] C. Li, M. Yang, Z. Liu, Z. Zhang, T. Zhu, X. Chen, Y. Dong, H. Cheng, Ru–Ni/Al2O3
bimetallic catalysts with high catalytic activity for N-propylcarbazole
The authors acknowledge supports from Natural Science Foundation hydrogenation, Catal. Sci. Technol. 10 (2020) 2268–2276.
of China, China (22078354, 21776295) and the Youth Innovation Pro­ [10] L. Li, M. Yang, Y. Dong, P. Mei, H. Cheng, Hydrogen storage and release from a new
promising Liquid Organic Hydrogen Storage Carrier (LOHC): 2-methylindole, Int.
motion Association of CAS, China (No.2017355).
J. Hydrog. Energy 41 (2016) 16129–16134.
[11] P.M. Modisha, J.H.L. Jordaan, A. Bösmann, P. Wasserscheid, D. Bessarabov,
Appendix A. Supporting information Analysis of reaction mixtures of perhydro-dibenzyltoluene using two-dimensional
gas chromatography and single quadrupole gas chromatography, Int. J. Hydrog.
Energy 43 (2018) 5620–5636.
Supplementary data associated with this article can be found in the
online version at doi:10.1016/j.apcatb.2022.121958.

10
L. Ge et al. Applied Catalysis B: Environmental 319 (2022) 121958

[12] H. Jorschick, A. Bulgarin, L. Alletsee, P. Preuster, A. Bösmann, P. Wasserscheid, [33] Q. Sun, N. Wang, J. Yu, Advances in catalytic applications of zeolite-supported
Charging a liquid organic hydrogen carrier with wet hydrogen from electrolysis, metal catalysts, Adv. Mater. 33 (2021) 2104442.
ACS Sustain. Chem. Eng. 7 (2019) 4186–4194. [34] Q. Sun, N. Wang, R. Bai, Y. Hui, T. Zhang, D.A. Do, P. Zhang, L. Song, S. Miao,
[13] H. Liu, C. Zhou, W. Li, W. Li, M. Qiu, X. Chen, H. Wang, Y. Sun, Ultralow Rh J. Yu, Synergetic effect of ultrasmall metal clusters and zeolites promoting
bimetallic catalysts with high catalytic activity for the hydrogenation of N- hydrogen generation, Adv. Sci. 6 (2019) 1802350.
Ethylcarbazole, ACS Sustain. Chem. Eng. 9 (2021) 5260–5267. [35] Q. Sun, B.W.J. Chen, N. Wang, Q. He, A. Chang, C.M. Yang, H. Asakura, T. Tanaka,
[14] W. Xue, H. Liu, B. Mao, H. Liu, M. Qiu, C. Yang, X. Chen, Y. Sun, Reversible M.J. Hulsey, C.H. Wang, J. Yu, N. Yan, Zeolite-encaged Pd-Mn nanocatalysts for
hydrogenation and dehydrogenation of N-ethylcarbazole over bimetallic Pd-Rh CO2 hydrogenation and formic acid dehydrogenation, Angew. Chem. Int. Ed. 59
catalyst for hydrogen storage, Chem. Eng. J. (2020). (2020) 20183–20191.
[15] M. Jang, Y.S. Jo, W.J. Lee, B.S. Shin, H. Sohn, H. Jeong, S.C. Jang, S.K. Kwak, J. [36] J. Shan, M. Li, L.F. Allard, S. Lee, M. Flytzani-Stephanopoulos, Mild oxidation of
W. Kang, C.W. Yoon, A. High-Capacity, Reversible liquid organic hydrogen carrier: methane to methanol or acetic acid on supported isolated rhodium catalysts,
H2-release properties and an application to a fuel cell, ACS Sustain. Chem. Eng. 7 Nature 551 (2017) 605–608.
(2019) 1185–1194. [37] F. Sun, Y. An, L. Lei, F. Wu, J. Zhu, X. Zhang, Identification of the starting reaction
[16] Y. Qin, J. Shi, X. Bai, Preparing ultra-stable Ru nanocatalysts supported on partially position in the hydrogenation of (N-ethyl)carbazole over Raney-Ni, J. Energy
graphitized biochar via carbothermal reduction for hydrogen storage of N- Chem. 24 (2015) 219–224.
ethylcarbazole, Int. J. Hydrog. Energy 46 (2021) 25543–25554. [38] L. Ge, G. Yu, X. Chen, W. Li, W. Xue, M. Qiu, Y. Sun, Effects of particle size on
[17] R. Qin, L. Zhou, P. Liu, Y. Gong, K. Liu, C. Xu, Y. Zhao, L. Gu, G. Fu, N. Zheng, bifunctional Pt/SAPO-11 catalysts in the hydroisomerization of n-dodecane, New J.
Alkali ions secure hydrides for catalytic hydrogenation, Nat. Catal. 3 (2020) Chem. 44 (2020) 2996–3003.
703–709. [39] C. Zhang, J. Sha, H. Fei, M. Liu, S. Yazdi, J. Zhang, Q. Zhong, X. Zou, N. Zhao,
[18] M. Babucci, A. Guntida, B.C. Gates, Atomically dispersed metals on well-defined H. Yu, Z. Jiang, E. Ringe, B.I. Yakobson, J. Dong, D. Chen, J.M. Tour, Single-atomic
supports including zeolites and metal-organic frameworks: structure, bonding, ruthenium catalytic sites on nitrogen-doped graphene for oxygen reduction
reactivity, and catalysis, Chem. Rev. 120 (2020) 11956–11985. reaction in acidic medium, ACS Nano 11 (2017) 6930–6941.
[19] S. Zhang, Z. Xia, M. Zhang, Y. Zou, H. Shen, J. Li, X. Chen, Y. Qu, Boosting selective [40] B. Yu, H. Li, J. White, S. Donne, J. Yi, S. Xi, Y. Fu, G. Henkelman, H. Yu, Z. Chen,
hydrogenation through hydrogen spillover on supported-metal catalysts at room T. Ma, Tuning the catalytic preference of ruthenium catalysts for nitrogen
temperature, Appl. Catal. B-Environ. 297 (2021). reduction by atomic dispersion, Adv. Funct. Mater. 30 (2019).
[20] M. Xiong, Z. Gao, Y. Qin, Spillover in heterogeneous catalysis: new insights and [41] Z. Liu, F. Huang, M. Peng, Y. Chen, X. Cai, L. Wang, Z. Hu, X. Wen, N. Wang,
opportunities, ACS Catal. 11 (2021) 3159–3172. D. Xiao, H. Jiang, H. Sun, H. Liu, D. Ma, Tuning the selectivity of catalytic nitriles
[21] Z. Gao, G. Wang, T. Lei, Z. Lv, M. Xiong, L. Wang, S. Xing, J. Ma, Z. Jiang, Y. Qin, hydrogenation by structure regulation in atomically dispersed Pd catalysts, Nat.
Enhanced hydrogen generation by reverse spillover effects over bicomponent Commun. 12 (2021) 6194.
catalysts, Nat. Commun. 13 (2022) 118. [42] X. Lu, C. Guo, M. Zhang, L. Leng, J.H. Horton, W. Wu, Z. Li, Rational design of
[22] W. Karim, C. Spreafico, A. Kleibert, J. Gobrecht, J. VandeVondele, Y. Ekinci, J. palladium single-atoms and clusters supported on silicoaluminophosphate-31 by a
A. van Bokhoven, Catalyst support effects on hydrogen spillover, Nature 541 photochemical route for chemoselective hydrodeoxygenation of vanillin, Nano
(2017) 68–71. Res. 14 (2021) 4347–4355.
[23] S. Wang, Z.J. Zhao, X. Chang, J. Zhao, H. Tian, C. Yang, M. Li, Q. Fu, R. Mu, [43] H. Xu, Z. Zhang, J. Liu, C.L. Do-Thanh, H. Chen, S. Xu, Q. Lin, Y. Jiao, J. Wang,
J. Gong, Activation and spillover of hydrogen on Sub-1 nm palladium nanoclusters Y. Wang, Y. Chen, S. Dai, Entropy-stabilized single-atom Pd catalysts via high-
confined within sodalite zeolite for the semi-hydrogenation of alkynes, Angew. entropy fluorite oxide supports, Nat. Commun. 11 (2020) 3908.
Chem. Int. Ed. 58 (2019) 7668–7672. [44] B. Wu, T. Lin, R. Yang, M. Huang, H. Zhang, J. Li, F. Sun, F. Song, Z. Jiang,
[24] H. Lee, Y.N. Choi, D.W. Lim, M.M. Rahman, Y.I. Kim, I.H. Cho, H.W. Kang, J. L. Zhong, Y. Sun, Ru single atoms for efficient chemoselective hydrogenation of
H. Seo, C. Jeon, K.B. Yoon, Formation of frustrated lewis pairs in Ptx -loaded nitrobenzene to azoxybenzene, Green Chem. 23 (2021) 4753–4761.
zeolite NaY, Angew. Chem. Int. Ed. 54 (2015) 13080–13084. [45] T.W. Kim, M. Kim, S.K. Kim, Y.N. Choi, M. Jung, H. Oh, Y.-W. Suh, Remarkably fast
[25] K. Chernichenko, Á. Madarász, I. Pápai, M. Nieger, M. Leskelä, T. Repo, low-temperature hydrogen storage into aromatic benzyltoluenes over MgO-
A. frustrated-Lewis-pair, approach to catalytic reduction of alkynes to cis-alkenes, supported Ru nanoparticles with homolytic and heterolytic H2 adsorption, Appl.
Nat. Chem. 5 (2013) 718–723. Catal. B-Environ. 286 (2021).
[26] A. Berkefeld, W.E. Piers, M. Parvez, Tandem frustrated lewis pair/tris [46] X. Deng, B. Qin, R. Liu, X. Qin, W. Dai, G. Wu, N. Guan, D. Ma, L. Li, Zeolite-
(pentafluorophenyl)borane-catalyzed deoxygenative hydrosilylation of carbon encaged isolated platinum ions enable heterolytic dihydrogen activation and
dioxide, J. Am. Chem. Soc. 132 (2010) 10660–10661. selective hydrogenations, J. Am. Chem. Soc. 143 (2021) 20898–20906.
[27] W. Chen, J. Han, Y. Wei, A. Zheng, Frustrated lewis pair in zeolite cages for alkane [47] M. Carosso, E. Vottero, A. Lazzarini, S. Morandi, M. Manzoli, K.A. Lomachenko, M.
activations, Angew. Chem. Int. Ed. 61 (2022) e202116269. J. Ruiz, R. Pellegrini, C. Lamberti, A. Piovano, E. Groppo, Dynamics of reactive
[28] S. Liu, M. Dong, Y. Wu, S. Luan, Y. Xin, J. Du, S. Li, H. Liu, B. Han, Solid surface species and reactant-induced reconstruction of Pt clusters in Pt/Al2O3 catalysts,
frustrated Lewis pair constructed on layered AlOOH for hydrogenation reaction, ACS Catal. 9 (2019) 7124–7136.
Nat. Commun. 13 (2022) 2320. [48] H. Ishikawa, J.N. Kondo, K. Domen, Hydrogen adsorption on Ru/ZrO2 studied by
[29] H. Jorschick, M. Vogl, P. Preuster, A. Bösmann, P. Wasserscheid, Hydrogenation of FT-IR, J. Phys. Chem. B 103 (1999) 3229–3234.
liquid organic hydrogen carrier systems using multicomponent gas mixtures, Int. J. [49] K.M. Eblagon, D. Rentsch, O. Friedrichs, A. Remhof, A. Zuettel, A.J. Ramirez-
Hydrog. Energy 44 (2019) 31172–31182. Cuesta, S.C. Tsang, Hydrogenation of 9-ethylcarbazole as a prototype of a liquid
[30] Y. Tian, H. Duan, B. Zhang, S. Gong, Z. Lu, L. Dai, C. Qiao, G. Liu, Y. Zhao, hydrogen carrier, Int. J. Hydrog. Energy 35 (2010) 11609–11621.
Template guiding for the encapsulation of uniformly subnanometric platinum [50] Y. Ding, Y. Dong, H. Zhang, Y. Zhao, M. Yang, H. Cheng, A highly adaptable Ni
clusters in beta-zeolites enabling high catalytic activity and stability, Angew. catalyst for liquid organic hydrogen carriers hydrogenation, Int. J. Hydrog. Energy
Chem. Int. Ed. 60 (2021) 21713–21717. 46 (2021) 27026–27036.
[31] X. Chen, M. Qiu, S. Li, C. Yang, L. Shi, S. Zhou, G. Yu, L. Ge, X. Yu, Z. Liu, N. Sun, [51] Y.D. Ming Yanga, Hansong Chengc, Hydrogenation kinetics of N-ethylcarbaozle as
K. Zhang, H. Wang, M. Wang, L. Zhong, Y. Sun, Gamma-ray irradiation to a heteroaromatic liquid organic hydrogen carrier, Adv. Mater. Res. 953–954
accelerate crystallization of mesoporous zeolites, Angew. Chem. Int. Ed. 59 (2020) (2014) 981–984.
11325–11329. [52] J. Im, H. Shin, H. Jang, H. Kim, M. Choi, Maximizing the catalytic function of
[32] Q. Sun, N. Wang, T. Zhang, R. Bai, A. Mayoral, P. Zhang, Q. Zhang, O. Terasaki, hydrogen spillover in platinum-encapsulated aluminosilicates with controlled
J. Yu, Zeolite-encaged single-atom rhodium catalysts: highly-efficient hydrogen nanostructures, Nat. Commun. 5 (2014) 3370.
generation and shape-selective tandem hydrogenation of nitroarenes, Angew.
Chem. Int. Ed. 58 (2019) 18570–18576.

11

You might also like