Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Catalysis Letters (2022) 152:1170–1181

https://doi.org/10.1007/s10562-021-03709-2

Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen


Production from Ammonia Decomposition
Xilun Zhang1,2 · Lin Liu1 · Ji Feng1,2 · Xiaohua Ju1 · Jiemin Wang1,3 · Teng He1 · Ping Chen1

Received: 19 February 2021 / Accepted: 11 June 2021 / Published online: 29 June 2021
© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2021

Abstract
Development of efficient catalysts for ammonia decomposition is crucial for hydrogen production from ammonia. Herein,
Ru nanoparticles (NPs) highly dispersed on P ­ r2O3 (Ru/Pr2O3) were prepared and investigated as catalyst for ammonia
decomposition reaction. Under relatively high weight hourly space velocity (WHSV = 30,000 mL ­gcat−1 ­h−1), the Ru/Pr2O3
achieves a ­H2 production rate of 20.9 mmol ­gcat−1 ­min−1 at 450 °C, which is superior to Ru/Pr6O11 and many other typical
oxides supported Ru catalysts. Detailed characterization results demonstrate that P ­ r2O3 has high density of basic sites, abun-
dant oxygen vacancies, and strong interaction with Ru, which can modulate electronic properties of Ru NPs and enhance
the intrinsic activity of the catalyst, making the catalyst highly active for ammonia decomposition. Moreover, the Ru/Pr2O3
catalyst also exhibits excellent stability and can maintain its high activity for over 100 h. The use of P
­ r2O3 as support offers
opportunity to improve the catalytic performance of Ru-based catalysts for ammonia decomposition.
Graphic Abstract
The high density of basic sites, abundant oxygen vacancies of ­Pr2O3, and the strong metal–support interaction effectively
promote the electron donation to Ru nanoparticles, enabling the superior catalytic performance of Ru/Pr2O3 for ammonia
decomposition.

Keywords Ru nanoparticles · Pr2O3 · Ammonia decomposition · Hydrogen production · Metal–support interaction ·


Electron transfer

* Lin Liu
liulin@dicp.ac.cn
* Ping Chen
pchen@dicp.ac.cn
Extended author information available on the last page of the article

13
Vol:.(1234567890)
Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen Production from Ammonia… 1171

1 Introduction (n = 4, 7, 9, 10, 11, 12, ∞) structure, and the extreme cases


are ­Pr2O3 and ­PrO2 [33, 34]. Among these different stoi-
Hydrogen is one of the most promising energy carriers chiometric praseodymium oxides, ­Pr6O11 is the most stable
for building a more sustainable future energy system. Up phase at ambient temperature in air. Due to the multiple
to now, storage of ­H2 through economic and sustainable valence states and oxygen-deficient structures, praseo-
way is still a major challenge in the scientific community dymium oxides have been widely used as catalysts, sup-
[1–3]. Ammonia ­(NH3) plays an important role in the earth ports, and dopants in a variety of catalytic reactions (e.g.,
ecosystem, providing a nitrogen source for manmade ferti- CO oxidation, ­CO2 methanation, synthesis of ethane and
lizers and serving as a precursor of various fibers and dyes ethylene, automotive emission control, ammonia synthesis
[4]. In recent years, ammonia attracts increasing attention and decomposition) [35–40]. Over the past few decades,
as a carbon-free H­ 2 carrier for its high content of hydrogen many studies have shown that the surface property and
(17.7 wt%), high density of energy (ca. 3 kWh ­kg−1), and structure of oxide support can profoundly affect catalytic
the ease of storage and transportation [5, 6]. ­NH3 decom- performance of supported Ru catalysts in N ­ H3 decomposi-
position reaction, a critical step for utilization of N ­ H3 as tion. Therefore, tuning the phases of praseodymium oxides
­H 2 feedstock for proton-exchange membrane fuel cells may offer new opportunity for improving the catalytic per-
(PEMFC), can only be achieved in the presence of efficient formance of supported Ru catalysts.
catalysts [7–9]. Because of high activity at relatively low In this work, highly dispersed Ru nanoparticles anchored
temperatures (400–500 °C), high degree of metal disper- on ­Pr2O3 and ­Pr6O11 were prepared and investigated in ­NH3
sion and excellent stability, supported Ru-based catalysts decomposition reaction. The obtained Ru/Pr 2O3 shows
have been widely used as efficient catalysts in ­NH3 decom- remarkably enhanced catalytic activity in comparison with
position reaction during the past 30 years [10, 11]. that of the Ru/Pr6O11. The structures of the Ru nanoparticles
NH3 decomposition on Ru-based catalysts has been rec- in Ru/Pr2O3 and Ru/Pr6O11 catalysts were characterized by
ognized to be a structure sensitive reaction. In Ru-based a variety of complementary techniques including X-ray dif-
catalytic systems, the size and structure of catalytically fraction (XRD), transmission electron microscopy (TEM),
active Ru particles greatly affect the catalytic perfor- electron paramagnetic resonance (EPR), X-ray photoelec-
mances of catalysts. Meanwhile, the surface property and tron spectroscopy (XPS) and X-ray absorption spectroscopy
structure of support can also influence the catalytic per- (XAS). Structure analyses and characterizations revealed
formances of supported Ru catalysts significantly [12–15]. that the synergy between the surface basic sites, oxygen
Thus, discovering new kinds of oxide supports that can vacancies of ­Pr2O3, and the strong interaction between Ru
effectively disperse and anchor Ru particles is one of NPs and P ­ r2O3 remarkably enhanced the intrinsic activity of
the major concerns for design of efficient Ru-based N ­ H3 Ru NPs in Ru/Pr2O3 for ­NH3 decomposition.
decomposition catalysts. In recent years, a variety of metal
oxides including MgO, ­SiO2, ­Al2O3, ­TiO2, ­ZrO2, ­CeO2,
Mg–Al mixed oxide, barium hexaaluminate, and perovs- 2 Experimental
kite material have been investigated as supports of Ru-
based catalysts for ­NH3 decomposition [16–23]. The role 2.1 Materials
of oxide supports is mainly to immobilize the active com-
ponent and improve the metal dispersion. Additionally, Praseodymium nitrate hexahydrate (Pr(NO3)3·6H2O) was
the metal–support interaction has significant impact on the purchased from Aladdin Co., Ltd. Ruthenium chloride
geometric structures and electric properties of metal cata- hydrate ­RuCl3·xH2O was purchased from Shanghai Tuosi
lysts and plays a vital role in tuning the catalytic behavior chemical Co., Ltd. Potassium bicarbonate (­ KHCO3), urea
of supported metal catalysts [24]. (CO(NH2)2) and nitric acid (­ HNO3) were all obtained from
Owing to the interactions between the localized 4f elec- Sinopharm Chemical Reagent Co. Ltd. All chemical rea-
trons and delocalized conduction bands, rare earth oxides gents in this work were analytical grade and used as received
such as ­La2O3, ­CeO2 and Y ­ 2O3 have attracted increasing without further purification.
attention as supports of catalysts for N ­ H3 decomposition
[20, 25–28]. As one of the important rare earth materials, 2.2 Preparation of ­PrCO3OH Support Precursor
praseodymium oxides with unique optical and electrical
properties have been studied for application in materials The ­PrCO3OH support precursor was prepared as following
with higher electrical conductivity, oxygen-storage com- steps. First, 4.35 g of Pr(NO3)3·6H2O and 3.0 g of K
­ HCO3
ponents, pigments, and catalysis [29–32]. Praseodymium were dissolved in 100 ml deionized water, respectively.
oxides have a series of stoichiometrically defined P ­ r nO2n−2 Then the Pr(NO3)3·6H2O solution was added dropwise into
the ­KHCO3 solution under stirring at room temperature,

13
1172 X. Zhang et al.

followed by an additional stirring for 1 h with the formation ANH3,in − ANH3,out


of a milky slurry. The product was filtered and washed with CNH3 (%) =
ANH3,in
deionized water for three times at least and dried at 80 °C
for 8 h to obtain ­PrCO3OH support precursor. where ANH3,in and ANH3,out is the total amount of ­NH3 in inlet
and outlet, respectively.
The ­H2 production rate was calculated form the N­ H3 con-
2.3 Preparation of the Catalysts version (CNH3) by the equation below:
WHSV × CNH3 × 1.5
Ru/Pr2O3 catalyst with different Ru contents of 1–5 wt% H2 production rate (mmol g−1 min−1 ) =
cat Vm × 60
were prepared by deposition–precipitation method. In a typi-
cal synthesis of 5% Ru/Pr2O3 sample, 1.356 g of ­PrCO3OH where WHSV is the weight hourly space velocity (30,000 mL
support precursor was added into 2 mL of R ­ uCl3·xH2O aque- ­gcat−1 ­h−1), CNH3 is the conversion of N
­ H3, Vm is molar vol-
ous solution (molar ratio of R­ uCl3: ­HNO3 = 1:1, 51.36 g/L), ume of gas at 25 °C and 1 atm (24.45 mL ­mmol−1).
and then dispersed into 60 ml of deionized water with stir- TOF H2 was calculated from the H ­ 2 production rate
ring. Subsequently, 5.94 g of urea (molar ratio of ­RuCl3: divided by the exposed surface of the Ru atoms:
urea = 1:200) was added into the suspension. The mixture
was refluxed at 80 °C for 8 h and aged at room tempera- H2 production rate × 10−3
ture for 12 h with stirring. The dark precipitate was sepa- TOFH2 (s−1 ) =
60 × Ru content × DRu ∕Ruat
rated by filtration and washed by deionized water to neu-
tral. The obtained sample were dried at 80 °C overnight. where D Ru is Ru dispersion of the catalyst deter-
The Ru/Pr2O3 catalyst was obtained after the dried sample mined by TEM statistical analysis, H ­ 2 production
was reduced at 550 °C in an ­NH3 flow (25 mL/min) for rate (mmol ­g cat −1 ­m in −1 ), Ru at is molar mass of Ru
2 h. Unless noted, Ru/Pr2O3 refers to the sample with a Ru (101.07 g ­mol−1).
content of 5 wt%.
For the 5% Ru/Pr6O11 catalysts, prior to the preparation
of the catalyst, the P­ rCO3OH support precursor was calci- 2.5 Catalyst Characterization
nated in air at 500 °C and the P ­ rCO3OH was decomposed
into ­Pr6O11. Then the obtained P ­ r6O11 was used as support. The actual Ru contents of the sample were determined by
The Ru/Pr6O11 catalysts was prepared by incipient wetness inductively coupled plasma optical emission spectroscopy
impregnation method with acetone as solvent. Typically, (ICP-OES, Optima 7300DV, Perkin-Elmer, USA). X-ray
0.1351 g of R ­ uCl3·xH2O was dissolved in 5 mL of acetone. diffraction (XRD) patterns were recorded on a PANalytical
1.0 g of obtained P ­ r6O11 was added into the solution and X’Pert with Cu-Kα radiation (λ = 1.5418 Å) at 40 kV and
stirred for 1 h. Then the solvent was removed and the sample 40 mA. Nitrogen physisorption was undertaken at − 196 °C
was dried at 80 °C overnight. The Ru/Pr6O11 catalyst was with a QUADRASORB SI (Quantachrome, USA) ana-
obtained after the dried sample was reduced at 550 °C in an lyzer. The samples were degassed at 300 °C for 6 h before
­NH3 flow (25 mL/min) for 2 h. Normally, Ru/Pr6O11 also the measurements. The surface areas were determined by
refers to the sample with a Ru content of 5 wt%. Brunauer–Emmett–Teller (BET) method. Scanning elec-
tron microscopy (SEM) was performed on a JEOL JSM-
7800F instrument. Transmission electron microscopy (TEM)
2.4 Catalytic Performance Evaluation and high-resolution transmission electron microscopic
(HRTEM) images were obtained on a JEOL JEM-2100X at
Catalytic ­NH3 decomposition reaction was performed in a 200 kV. The dispersion of Ru particles (DRu) was calculated
continuous fixed-bed flow quartz reactor under atmospheric using the equation by Borodziński and Bonarowska [41]:
pressure. 50 mg of the reduced catalyst (40–60 mesh) was √
dat × 3.32
placed in the central section of reactor. Catalytic perfor- DRu = 1.23
for 0.2 ≤ DRu ≤ 0.92
mance evaluation was carried out in pure N­ H3 with a flow dTEM
rate of 25 mL ­min−1 (WHSV = 30,000 mL ­gcat−1 ­h−1) in the
range of 300–550 °C. Gas product analysis was conducted dat × 5.01
on a gas chromatograph (Agilent, 7890B) equipped with a DRu =
dTEM
for DRu < 0.2
TCD detector and Poropak N column by using ­H2 as carrier
gas. where dat is the atomic diameter of Ru (0.269 nm).
The conversion of ­NH3 (CNH3) was calculated as the given Temperature-programmed desorption of ­CO2 ­(CO2-TPD)
equation below: was performed on an AutoChem 2920 instrument. Prior to

13
Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen Production from Ammonia… 1173

the experiment, 50 mg of sample was reduced in a pure


­ 2 flow at 500 °C for 2 h, purged in a He flow for another
H
1 h, and cooled to 50 °C. Then pure C­ O2 was adsorbed for
30 min. After purging the physisorbed C­ O2 in a He flow for
30 min, the sample was heated from 50 to 800 °C with a
ramping rate of 10 °C/min.
The electron paramagnetic resonance (EPR) spectra
were recorded on a Bruker A200 spectrometer operated at
the X-band frequency. The X-ray absorption fine structure
(XAFS) spectra at Ru K edge (22,117 eV) were collected at
the BL14W1 beamline of Shanghai Synchrotron Radiation
Facility (SSRF) using a Si (311) double-crystal monochro-
mator. The samples were sealed with KAPTON film to avoid
air contamination before test. The spectra were recorded at
room temperature under transmission mode and the energy
was calibrated by using Ru foil. k2-weighted Fourier trans-
form EXAFS (FT-EXAFS) were analyzed using carried out
Athena and Artemis Software. Wavelet transform EXAFS
(WT-EXAFS) was analyzed by Morlet wavelets with opti-
mum resolution. Surface chemical states of catalysts were
characterized by X-ray photoelectron spectroscopy (XPS)
using an Escalab 250 Xi X-ray photoelectron spectrometer
(Thermo Scientific) with nonmonochromatic Al Ka radia-
tion. The binding energy was calibrated according to the
XPS peak at carbon 1s at 284.8 eV.

3 Results and Discussion

Praseodymium carbonate hydroxide ­(PrCO3OH) and ­Pr6O11


derived from P ­ rCO3OH were used as supports to prepare
Ru/Pr2O3 and Ru/Pr6O11 catalysts, respectively. The crystal
structure of the supports and catalysts after N­ H3 decomposi-
tion reaction were analyzed by XRD (Fig. 1). The diffraction
peaks of Ru/Pr2O3 can be indexed into a hexagonal P ­ r 2O 3 Fig. 1  XRD pattens of Ru/Pr2O3 and Ru/Pr6O11 catalysts. Standard
patterns of Ru, P
­ r2O3 and ­Pr6O11 were included as references
phase (JCPDS Card No. 06-0410, Fig. 1a). No feature peaks
corresponding to the ruthenium species can be found in the
XRD patterns, indicating the high dispersion of Ru NPs on
the surface of P ­ r2O3 support. For the Ru/Pr6O11, relatively is only 0.5°–1.0°, the bulk phase of cubic ­Pr6O11 structure in
weak peaks at 41.6° and 44.5° associated with the (002) and Ru/Pr6O11 were still maintained, indicating its highly crys-
(101) reflections of Ru metal (JCPDS Card No. 01-1256) talline structure. There are some additional weak peaks not
can be observed in addition to the characteristic diffraction assigned to P­ r6O11 or Ru, including the peaks at 25.6°, 31.1°,
peaks associated with cubic ­Pr6O11 phase (JCPDS Card No. and 34.5°. These additional peaks should be associated with
42-1121, Fig. 1b). The diffraction peaks of metallic Ru sug- other phases of praseodymium oxides with a mixture of P ­ r3+
4+
gest the relatively larger Ru NPs in Ru/Pr6O11. It is also and ­Pr [34]. Due to the weak intensity of these peaks and
worth to be noted that the characteristic diffraction peaks the complex phases of praseodymium oxides, these peaks
of ­Pr6O11 support in Ru/Pr6O11 shift towards lower angle in are difficult to be attributed to the individual phase. The
compared with that of pure P ­ r6O11 support, which could be physical properties of the catalysts are shown in Table 1, the
attributed to the lattice expansion of ­Pr6O11 under reducing ­Pr2O3 exhibits higher BET surface area than that of ­Pr6O11,
atmosphere. Because the ionic radius of ­Pr3+ (0.1126 nm) which may be beneficial for the dispersion of Ru NPs. The
is larger than that of ­Pr4+ (0.096 nm), some part of ­Pr4+ was actual Ru content was determined with ICP-OES analysis.
reduced to P­ r3+ during the reduction, resulting in the lattice Ru contents in these two samples are close to the desired 5
expansion of ­Pr6O11 [42]. However, due to the shift of peaks wt% Ru mass loading.

13
1174 X. Zhang et al.

Table 1  Physical properties of the Ru/Pr2O3 and Ru/Pr6O11 catalysts mildly endothermic nature of ­NH3 decomposition reaction.
Samples Ru content BET Mean Dispersion (%) a For the catalysts with similar Ru content, the 5% Ru/Pr2O3
(wt%) surface area particle size catalyst shows significantly higher catalytic activity than
­(m2/g) (nm)a 5% Ru/Pr6O11 in the whole temperature range. The activ-
ity of 5% Ru/Pr6O11 could only be detected at the tempera-
Ru/Pr2O3 4.8 30.1 2.0 52.0
tures above 350 °C. By contrast, 5% Ru/Pr2O3 exhibits an
Ru/Pr6O11 4.7 18.1 6.6 20.4
obvious ­NH3 conversion of 6.13% corresponding to a ­H2
a
Mean particle size and dispersion of Ru NPs were determined by production rate of 1.9 mmol ­gcat−1 ­min−1 at the tempera-
TEM study using the equation of Borodziński and Bonarowska [41] ture as low as 300 °C. Furthermore, the N ­ H3 conversion
of 5% Ru/Pr6O11 at 450 °C is only 16.6% with a ­H2 pro-
The activities of praseodymium oxide supported Ru NPs duction rate of 5.1 mmol ­gcat−1 ­min−1, which is at a com-
catalysts in N
­ H3 decomposition reaction were evaluated, parable level with the activity of previously reported Ru/
and the temperature-dependence of activities of Ru/Pr2O3 Pr6O11 (3.1 mmol ­gcat−1 ­min−1) in the literature at the same
and Ru/Pr6O11 catalysts results are presented in Fig. 2a. temperature [39]. However, 5% Ru/Pr2O3 gives a ­NH3 con-
For all the samples, the ­NH3 conversion increases with version as high as 68.2% at 450 °C, 4.1 times that of the 5%
increase of reaction temperature, agreeing well with the Ru/Pr6O11 at the same temperature. The activities of Ru/

Fig. 2  a Temperature-dependence of ­NH3 decomposition activities ture. c Long-term stability tests of 5% Ru/Pr2O3 catalyst at 450 and
of Ru/Pr2O3 with different Ru contents and 5% Ru/Pr6O11 catalysts. 500 °C, respectively. d Arrhenius plots and apparent activation ener-
b Influence of WHSV on the ­NH3 conversion and ­H2 production rate gies of Ru/Pr2O3 with different Ru contents and 5% Ru/Pr6O11 cata-
of 5% Ru/Pr2O3 catalyst for ­NH3 decomposition at different tempera- lysts for N
­ H3 decomposition

13
Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen Production from Ammonia… 1175

Pr2O3 catalysts with different Ru contents were also evalu- Additionally, the similar Ea values elucidate the similar
ated. As shown in Fig. 2a, ­NH3 conversion increases with structure of active sites in Ru/Pr2O3 catalysts.
the increase of Ru content from 1 to 5 wt%. Specifically, the The morphology and structure of the catalysts were ana-
­H2 production rate at 450 °C shows linear increase from 3.6 lyzed by SEM and TEM. As shown in Fig. 3a, ­Pr2O3 displays
to 20.9 mmol ­gcat−1 ­min−1 with the increase of Ru content a well-defined petal-like structure with a diameter of ca.
from 1 to 5% (Fig. S1). Remarkably, 2% Ru/Pr2O3 even gives 2.0 μm, in which many exposed defects could be observed
higher activity than that of the 5% Ru/Pr6O11, manifesting on the surface. The TEM image demonstrated that the ­Pr2O3
the superior catalytic activity of Ru/Pr2O3 catalyst. A com- support in Ru/Pr2O3 possess good crystallinity. Interplanar
parison of activity of Ru/Pr2O3 catalyst with those of the typ- spacings of 0.30 nm, corresponding to the lattice fringes
ical supported Ru catalysts for N ­ H3 decomposition reported of the (002) facet of P ­ r2O3, can be clearly identified in the
in literatures is listed in Table S1. The activity of 5% Ru/ TEM image of Ru/Pr2O3 (Fig. 3b). Meanwhile, the P ­ r2O 3
Pr2O3 is not only significantly higher than that of 5% Ru/ exhibits higher BET surface area than that of ­Pr6O11, which
Pr6O11 but also outperforms most of the promoter-free effi- also should be beneficial for dispersion of Ru NPs (Table 1).
cient Ru catalysts under similar reaction conditions. Under To confirm this, TEM was applied to investigate the parti-
identical reaction conditions, the 5% Ru/Pr2O3 at 450 °C cle size distribution of Ru NPs of the catalysts. As the low
achieves a ­H2 production rate of 20.9 mmol ­gcat−1 ­min−1, contrast of Ru and Pr elements in the TEM images, it is chal-
comparable with that of the highly efficient Ru/c–MgO lenging to distinguish the Ru NPs from the more condensed
catalyst [43]. praseodymium oxide support. Thus, the Ru NPs can only
The weight hourly space velocity (WHSV) is a significant be determined from the NPs distributed on the edge of the
factor for practical application [44]. The influence of WHSV support. By counting ca. 100 Ru NPs from different TEM
on the ­NH3 decomposition activity was also investigated. images in various regions (Fig. S2 and S3), the particle size
As shown in Fig. 2b, ­NH3 conversion decreases with the distribution and average diameter of Ru NPs of Ru/Pr2O3
increase of WHSV over 5% Ru/Pr2O3 at the temperature of and Ru/Pr6O11 samples can be preliminarily estimated. The
450, 475 and 500 °C as a result of decrease in the residence average particle size of Ru NPs in Ru/Pr2O3 is ca. 2.0 nm
time of the N­ H3 over the catalyst bed [45]. By contrast, the (Fig. 3c), which is much smaller than that of the Ru/Pr6O11
increase of WHSV results in a remarkable increase of H ­ 2 (6.6 nm, Fig. 3f). The dispersion of Ru NPs are shown
production rate. For example, at 500 °C, the ­H2 production in Table 1, the Ru dispersion of Ru/P2O3 is about 52.0%,
rate increases from 5.1 to 50 mmol ­gcat−1 ­min−1 with WHSV which is much higher than that of the Ru/Pr6O11 (20.4%).
increasing from 5000 to 100,000 mL ­gcat−1 ­h−1. Further- It reveals that the use of P­ r2O3 support as support leads to
more, the long-term stability is critical for evaluating the much enhanced dispersion of Ru, which is in line with the
catalytic performance of the catalyst in ­NH3 decomposition. superior catalytic activity of Ru/Pr2O3 catalyst. Furthermore,
Here, two sets of stability test of 5% Ru/Pr2O3 catalyst were HRTEM image of Ru/Pr2O3 reveals that highly dispersed
carried out at 450 and 500 °C with relatively high space Ru NPs are embedded in support and partially covered by
velocity of 30,000 mL ­gcat−1 ­h−1, respectively. As shown in ­Pr2O3 (inset of Fig. 3b), indicating the strong metal–support
Fig. 2c, both of the ­NH3 conversions at 450 and 500 °C over interaction between Ru NPs and ­Pr2O3.
5% Ru/Pr2O3 remain almost constant during the test period Currently, the B5 sites on the surface of Ru particles have
of 100 h, evidencing the high stability of 5% Ru/Pr2O3 cata- been widely accepted as the active sites of Ru-based cata-
lyst, which is crucial for the potential application of 5% Ru/ lysts for ­NH3 decomposition [14, 46, 47]. Many theoretical
Pr2O3 for application in ­COx-free hydrogen production from and experimental results demonstrate that the amount of
­NH3 decomposition. B5 sites on the Ru NPs is highly dependent on the particle
To clarify the origin of high activity of Ru/Pr2O3 catalyst, size of Ru NPs. The maximum probability for B5 sites is of
the apparent activation energies (Ea) of different catalysts 1.8–2.5 nm in Ru particle size. For the particles larger than
were calculated based on H ­ 2 production rates far away from the optimal size region, the probability for B5 sites decrease
the equilibrium value. Figure 2d shows Arrhenius plots of monotonically [46]. As a result, the Ru NPs on P ­ r2O3 with
5% Ru/Pr6O11 and Ru/Pr2O3 catalysts with different Ru con- an optimal particle size of 2.0 nm are expected to show much
tents. Correlating well with the catalytic activity results, a enhanced concentration of B5 sites on the surface, agreeing
remarkable decrease in the Ea value can be observed for well the superior catalytic performance.
5% Ru/Pr2O3 (61.9 kJ ­mol−1) in comparison with that of Ru catalyzed ­NH3 synthesis is recognized to be a structure
the 5% Ru/Pr6O11 (87.2 kJ ­mol−1). For the Ru/Pr2O3 with sensitive reaction, and the ­TOFH2 value usually increases
different Ru contents, the apparent activation energies are with the increase of size of Ru nanoparticles. A rise in TOF
in the range of 57.7–66.3 kJ ­mol−1. The Ea of 3% Ru/Pr2O3 over promoted Ru/Al2O3 and AC-supported Ru catalysts has
(57.7 kJ ­mol−1) is one of the lowest Ea values for Ru-based been observed with the increase of size of Ru particles [48,
­NH3 decomposition catalysts in the literatures (Table S1). 49]. Additionally, it has been found that ­NH3 decomposition

13
1176 X. Zhang et al.

Fig. 3  Representative SEM, TEM images and Ru particle size distributions of Ru/Pr2O3 (a–c) and Ru/Pr6O11 (d–f) catalysts after N
­ H3 decompo-
sition reaction

on Ru/γ-Al2O3 catalysts is also highly structure sensitive,


with ­TOFH2 values increasing by almost 2 orders of magni-
tude as the particle size increases from 0.8 to 7.5 nm [15].
Different from the Ru-based catalyst systems reported in
the previous studies, 5% Ru/Pr2O3 with a relatively small
particle size of Ru NPs shows much higher ­TOFH2 value
than that of the 5% Ru/Pr6O11 catalyst under identical reac-
tion conditions. For example, 5% Ru/Pr6O11 with an average
Ru particle size of 6.6 nm gives a ­TOFH2 value of 0.89 ­s−1
at 450 °C. For 5% Ru/Pr2O3, the average Ru particle size
is only 2.1 nm. However, a remarkably increased ­TOFH2
value of 1.41 ­s−1 can be observed, which is 1.6 times that
of the 5% Ru/Pr6O11 (Fig. 4). In comparison with that of
the 5% Ru/Pr6O11, the Ru NPs in 5% Ru/Pr2O3 sample not
only show much enhanced catalytic activity but also show
much enhanced intrinsic activity for ­NH3 decomposition.
The remarkably increased ­TOFH2 value over 5% Ru/Pr2O3
catalyst indicates that another factor is highly related with
Fig. 4  TOFH2 values of 5% Ru/Pr2O3 and 5% Ru/Pr6O11 catalysts as a
the intrinsic activity instead of particle size effect.
function of reaction temperature in ­NH3 decomposition
As both of the P ­ r2O3 and ­Pr6O11 are derived from same
­PrCO3OH precursor, the superior intrinsic activity of Ru/
Pr2O3 catalyst may arise from the difference of surface
properties of supports. The C ­ O2-TPD measurement was

13
Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen Production from Ammonia… 1177

Pr2O3 sample possesses significantly higher density of sur-


face basic sites than that of the Ru/Pr6O11. The density of
medium and strong basic sites of Ru/Pr2O3 sample is about
21.66 μmol ­m−2, which is much higher than that of Ru/
Pr6O11 (12.57 μmol ­m−2). The higher amount of strong basic
sites of Ru/Pr2O3 is beneficial for promoting the electron
donation to Ru surface, which may account for the superior
activity of Ru/Pr2O3 catalyst [52, 53].
Metal oxides are frequently used as catalysts and cata-
lyst supports in the field of heterogeneous catalysis. The
oxygen vacancies and F centers, surface hydroxyl groups
and reduction behavior of oxides strongly affect the sur-
face properties of metal oxides [54]. As one of the most
interesting catalytic structures, the oxygen vacancies on
the surface of ceria play a critical role in regulating the
catalytic performances of ceria nanostructured materials in
a variety of reactions (e.g., CO oxidation, water gas shift
reaction, hydrogenation of CO/CO2) [55–58]. Besides the
Fig. 5  CO2-TPD profiles of Ru/Pr2O3 and Ru/Pr6O11 catalysts ceria, praseodymia with multiple valence states also has
an oxygen-deficient structure [59]. To further understand
the surface properties of Ru/Pr2O3, EPR characterization
conducted to study the basic properties of the Ru/Pr2O3 and was performed to probe the density of oxygen vacancies
Ru/Pr6O11 samples. Figure 5 displays the C ­ O2 desorption of Ru/Pr2O3 and Ru/Pr6O11 samples. Here, EPR spectrum
profiles of Ru/Pr2O3 and Ru/Pr6O11. The Ru/Pr6O11 sam- of Ru/Pr2O3 shows a symmetrical and sharp EPR signal at
ple shows a weak ­CO2 desorption peaks at ca. 80 °C and g = 2.003 (Fig. 6). According to the literature, the signal
another peak at ca. 621 °C. However, three ­CO2 desorp- at g = 2.003 can be ascribed to the electron trapped on the
tion peaks at ca. 80, 300, and 711 °C can be observed for oxygen vacancy [60]. The strong signal at g = 2.003 for
Ru/Pr2O3, which are corresponding to the weak, medium, Ru/Pr2O3 demonstrates the presence of a large amount of
and strong basic sites of the ­Pr2O3 support. Generally, the oxygen vacancies in the Ru/Pr2O 3 sample. By contrast,
amount of C ­ O2 desorption indicates the number of basic only a very weak EPR signal corresponding to the oxygen
sites and the temperature of the desorption is related to the vacancies can be observed in the EPR spectrum of Ru/
strength of the basic sites [50]. The desorption temperature
of strong basic sites over Ru/Pr2O3 (711 °C) is higher than
that of Ru/Pr6O11 (621 °C), and the amount of C ­ O2 desorp-
tion over Ru/Pr2O3 (671.9 μmol ­g−1) is 2.4 times that of Ru/
Pr6O11 (227.5 μmol ­g−1). The density of basic sites (per unit
surface area) was calculated by the integral method based
on the area of the corresponding desorption peaks [51]. As
shown in Table 2, the quantified results suggest that the Ru/

Table 2  The amount and density of basic sites in Ru/Pr2O3 and Ru/
Pr6O11 catalysts determined by C
­ O2-TPD
Sample Peak temperature Adsorption of Density of
(°C) ­CO2 basic sites
(μmol ­g−1) (μmol ­m−2)

Ru/Pr2O3 80 19.9 0.66


300 337.3 11.21
711.4 314.7 10.45
Total 617.9 22.32
Ru/Pr6O11 80 – –
621 227.6 12.57
Total 227.6 12.57
Fig. 6  X‐band EPR spectra of Ru/Pr2O3 and Ru/Pr6O11 catalysts

13
1178 X. Zhang et al.

Pr6O11 at the same g value, implying that only a very small interaction between the ­Pr2O3 support and Ru NPs, causing
amount of oxygen vacancies presents in Ru/Pr6O11 sample. the higher electron density of Ru NPs in Ru/Pr2O3.
The EPR results confirm the existence of abundant oxygen The surface composition of the catalysts and the chemical
vacancies in the Ru/Pr2O3, which could promote the elec- states of the elements were investigated by XPS technique.
tron transfer from ­Pr2O3 support to Ru surface [50, 61]. Figure 8a shows the Pr 3d XPS spectra of Ru/Pr2O3 and Ru/
Here, EXAFS and XANES spectroscopies were per- Pr6O11 catalysts after N­ H3 decomposition reaction. The Pr
formed to gain a deep insight of the structure of the Ru ­3d5/2 and ­3d3/2 peaks of Ru/Pr6O11 shift to higher binding
NPs in Ru/Pr2O3 and Ru/Pr6O11 samples. Similar peaks at energies than that of Ru/Pr2O3, centered at ca. 935.1 and
∼ 2.4 Å, associated with Ru–Ru scattering in the first coor- 955.1 eV, respectively, which can be assigned to ­Pr4+ [65].
dination shell of metallic Ru, can be observed in the EXAFS The satellite peaks at lower binding energy indicates the
spectra of Ru/Pr2O3 and Ru/Pr6O11 (Fig. 7a). Analogous to existence of P­ r3+, suggesting the coexistence of P ­ r3+ and
4+
FT, wavelet transform (WT) is a mathematical complete ­Pr in Ru/Pr6O11 [66]. For Ru/Pr2O3 sample, the Pr ­3d5/2
transformation of EXAFS signals. It not only provides radial and ­3d3/2 peaks are located at ca. 933.6 and 954.1 eV respec-
distance resolution but also resolve in k space [62]. As shown tively, in agreement with the binding energies of P ­ r3+ in
3+
in Fig. S4, the maximum intensity of WTs for Ru/Pr2O3 ­Pr2O3, which suggests the Pr mainly exists as P ­ r in Ru/
and Ru/Pr6O11 samples are located at ca. 8.8 Å−1, which Pr2O3 [67, 68]. The XPS analyses of the valence states of
is assigned to Ru–Ru scattering, demonstrating the metal- Pr are consistent with the XRD results, in which the bulk
lic state of Ru species in the two samples. In comparison phases of Ru/Pr2O3 and Ru/Pr6O11 are ­Pr2O3 and P ­ r6O11,
with that of the Ru foil, the intensities of peak from the first respectively.
shell Ru–Ru scattering show obvious decrease for the Ru/ Understanding the origin of the high intrinsic activity
Pr2O3 and Ru/Pr6O11 samples, demonstrating the nanopar- of Ru NPs on P ­ r2O3 and ­Pr6O11 is particularly important.
ticle form of the Ru species in the catalysts. The Ru K-edge On the different substrates, the active chemical state of the
XANES spectra of Ru/Pr2O3 and Ru/Pr6O11 are shown in Ru species may vary greatly. The electron structure of Ru
Fig. 7b, with Ru and ­RuO2 as references. The XANES spec- NPs has a significant impact on the catalytic activity of N ­ H3
tra of Ru/Pr2O3 and Ru/Pr6O11 are located between the Ru decomposition. We suppose that the strong interaction at the
foil and R­ uO2, but closer to that of the Ru foil, indicating Ru–support interface can tune the electronic structure of Ru
the metallic Ru are dominant species in these two samples. NPs and thus make the catalyst superior active in the ­NH3
Compared with Ru/Pr6O11, a shift of edge energy towards Ru decomposition reaction. To check this, XPS was employed
foil could be observed in Ru/Pr2O3, revealing the Ru/Pr2O3 to investigate the chemical state of Ru species in Ru/Pr2O3
possesses relatively higher electron density than that of the and Ru/Pr6O11 catalysts. Due to the overlap of Ru 3­ d3/2 and
Ru/Pr6O11 [63, 64]. The EPR results confirm the presence C 1s peak at c.a. 284.8 eV, Ru chemical state was deter-
of oxygen vacancies in Ru/Pr2O3, which can promote the mined by XPS analyses of Ru 3p core level spectra. Fig-
electron transfer form ­Pr2O3 support to Ru NPs through the ure 8b illustrates the Ru ­3p3/2 XPS spectra of Ru/Pr2O3 and

Fig. 7  a Fourier transforms of k2-weighted Ru K-edge EXAFS spectra and b Ru K-edge XANES spectra for Ru foil, R
­ uO2, Ru/Pr2O3 and Ru/
Pr6O11 catalysts

13
Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen Production from Ammonia… 1179

Fig. 8  XPS analyses of Ru/Pr2O3 and Ru/Pr6O11 catalysts: a Pr 3d core level spectra; b Ru 3­ p3/2 core level spectra; c relative abundance (%) of
the Ru species in Ru 3­ p3/2 spectra

Ru/Pr6O11 samples. It can be found that the binding energy structure makes the metallic Ru species in Ru/Pr2O3 highly
of Ru 3­ p3/2 of Ru/Pr2O3 is obviously lower than that of Ru/ active in N
­ H3 decomposition, which is responsible for its
Pr6O11. The Ru ­3p3/2 region could be deconvoluted into two superior intrinsic activity in ­NH3 decomposition.
peaks of metallic ­Ru0 (ca. 462.2 eV) [48, 69] and R ­ un+ (ca.
463.7–464.4 eV) [70, 71]. The partially positive charge ­Run+
species can be attributed to the Ru species located at the 4 Conclusion
interface between Ru NPs and support. The relative contents
of different Ru species are shown in Fig. 8c. The relative In summary, this work demonstrates that P ­ r2O3 is a superior
content of R­ u0 in Ru/Pr2O3 (63.5%) is much higher than that support for preparation of highly efficient and stable Ru/
of Ru/Pr6O11 (41.7%), demonstrating the significant increase Pr2O3 catalysts for N­ H3 decomposition. The high density
of electron density of Ru NPs in Ru/Pr2O3, which is consist- of basic sites, abundant oxygen vacancies of P ­ r2O3, and the
ent with the result of XANES. strong metal–support interaction can not only render the
Generally, many different factors such as the dispersion, high dispersion of Ru NPs on the ­Pr2O3 surface, but also
size and morphology of Ru NPs, interaction between Ru NPs benefit for the electron donation to the Ru NPs, leading to
and support, and surface property of support can influence the increase of electron density of Ru NPs in Ru/Pr2O3. As
the intrinsic catalytic activity of supported Ru catalysts. It a result, Ru/Pr2O3 exhibits superior catalytic activity and
is well-known that the B5 site on the surface of metallic stability in ­NH3 decomposition reaction. We anticipate that
Ru species is recognized as the active site for decomposi- these finding could provide an insightful understanding of
tion of N­ H3 [14, 46, 47]. Hence, the percentage of metallic the role of the support on the catalytic behavior and open
Ru species is vital to the catalytic decomposition of ­NH3. opportunities to develop highly efficient catalysts for N­ H3
High percentages of metallic Ru species could lead to much decomposition.
enhanced catalytic activity. For the Ru/Pr2O3 with high per-
centage of metallic Ru species, much enhanced T ­ OFH2 can Supplementary Information The online version contains supplemen-
be observed even the sample has a relatively small particle tary material available at https://d​ oi.o​ rg/1​ 0.1​ 007/s​ 10562-0​ 21-0​ 3709-2.
size, revealing the superior intrinsic activity of Ru/Pr2O3
Acknowledgements This work was supported by the Strategic
sample (Fig. 4). The C ­ O2-TPD and EPR results show the Priority Research Program of the Chinese Academy of Sciences
presence of high density of basic sites and oxygen vacancies (XDA21010208), the Light Source Fund of Dalian Institute of Chemi-
in Ru/Pr2O3 sample respectively, which can facilitate the cal Physics, Chinese Academy of Sciences (DICP DCLS201701), and
electron transfer from P ­ r2O3 support to Ru NPs, leading to the K. C. Wong Education Foundation (GJTD-2018-06).
the increase of the electron density of Ru NPs in Ru/Pr2O3.
The electron-rich Ru surface could enhance the recombina-
tive desorption of adsorbed nitrogen and improve the ­NH3 References
decomposition activity. Thus, it is worthwhile to empha-
1. Züttel A (2003) Mater Today 6:24–33
size the critical role of surface basic properties and oxygen
2. Mori D, Hirose K (2009) Int J Hydrog Energy 34:4569–4574
vacancies of ­Pr2O3 support in modulating the electronic 3. He T, Pachfule P, Wu H, Xu Q, Chen P (2016) Nat Rev Mater
structure of Ru/Pr2O3 catalyst. The modulated electronic 1:16059

13
1180 X. Zhang et al.

4. Erisman JW, Sutton MA, Galloway J, Klimont Z, Winiwarter W 38. Shi J, Li H, Zhao W, Qi P, Wang H (2020) Catal Sci Technol
(2008) Nat Geosci 1:636–639 10:7291–7301
5. Klerke A, Christensen CH, Nørskov JK, Vegge T (2008) J Mater 39. Nagaoka K, Honda K, Ibuki M, Sato K, Takita Y (2010) Chem
Chem 18:2304 Lett 39:918–919
6. Lan R, Irvine JTS, Tao S (2012) Int J Hydrog Energy 40. Nagaoka K, Eboshi T, Abe N, Miyahara S-i, Honda K, Sato K
37:1482–1494 (2014) Int J Hydrog Energy 39:20731–20735
7. Choudhary T, Sivadinarayana C, Goodman D (2001) Catal Lett 41. Borodziński A, Bonarowska M (1997) Langmuir 13:5613–5620
72:197–201 42. Matović B, Pantić J, Prekajski M, Stanković N, Bučevac D,
8. Schüth F, Palkovits R, Schlögl R, Su DS (2012) Energy Environ Minović T, Čebela M (2013) Ceram Int 39:3151–3155
Sci 5:6278–6289 43. Ju X, Liu L, Zhang X, Feng J, He T, Chen P (2019) Chem-
9. Yao Q, Ding Y, Lu Z-H (2020) Inorg Chem Front 7:3837–3874 CatChem 11:4161–4170
10. Yin SF, Xu BQ, Zhou XP, Au CT (2004) Appl Catal A 277:1–9 44. Lamb KE, Dolan MD, Kennedy DF (2019) Int J Hydrog Energy
11. Mukherjee S, Devaguptapu SV, Sviripa A, Lund CRF, Wu G 44:3580–3593
(2018) Appl Catal B 226:162–181 45. Wang W, Herreros JM, Tsolakis A, York APE (2013) Int J
12. Raróg-Pilecka W, Szmigiel D, Komornicki A, Zieliński J, Kow- Hydrog Energy 38:9907–9917
alczyk Z (2003) Carbon 41:589–591 46. Jacobsen CJH, Dahl S, Hansen PL, Tornqvist E, Jensen L, Top-
13. Zheng W, Zhang J, Xu H, Li W (2007) Catal Lett 119:311–318 soe H, Prip DV, Moenshaug PB, Chorkendorff I (2000) J Mol
14. García-García FR, Guerrero-Ruiz A, Rodríguez-Ramos I (2009) Catal A Chem 163:19–26
Top Catal 52:758–764 47. Rarogpilecka W, Miskiewicz E, Szmigiel D, Kowalczyk Z
15. Karim AM, Prasad V, Mpourmpakis G, Lonergan WW, (2005) J Catal 231:11–19
Frenkel AI, Chen JG, Vlachos DG (2009) J Am Chem Soc 48. Murata S, Aika K-I (1992) J Catal 136:110–117
131:12230–12239 49. Liang C, Wei Z, Xin Q, Li C (2001) Appl Catal A 208:193–201
16. Ju X, Liu L, Yu P, Guo J, Zhang X, He T, Wu G, Chen P (2017) 50. Ogura Y, Tsujimaru K, Sato K, Miyahara S-i, Toriyama T,
Appl Catal B 211:167–175 Yamamoto T, Matsumura S, Nagaoka K (2018) ACS Sustain
17. Li X, Ji W, Zhao J, Wang S, Au C (2005) J Catal 236:181–189 Chem Eng 6:17258–17266
18. Yin SF, Xu BQ, Zhu WX, Ng CF, Zhou XP, Au CT (2004) Catal 51. Kikhtyanin O, Čapek L, Smoláková L, Tišler Z, Kadlec D,
Today 93–95:27–38 Lhotka M, Diblíková P, Kubička D (2017) Ind Eng Chem Res
19. Yin SF, Xu BQ (2003) ChemPhysChem 4:277–281 56:13411–13422
20. Hu X-C, Fu X-P, Wang W-W, Wang X, Wu K, Si R, Ma C, Jia C-J, 52. Ni J, Jing B, Lin J, Lin B, Zhao Z, Jiang L (2018) J Rare Earths
Yan C-H (2020) Appl Catal B 268:118424 36:135–141
21. Su Q, Gu LL, Zhong AH, Yao Y, Ji WJ, Ding WP, Au CT (2018) 53. Wang Z, Qu Y, Shen X, Cai Z (2019) Int J Hydrog Energy
Catal Lett 148:894–903 44:7300–7307
22. Wang Z, Cai Z, Wei Z (2019) ACS Sustain Chem Eng 54. Vedrine JC (2019) ChemSusChem 12:577–588
7:8226–8235 55. Zhu Q, Peng Y, Lin L, Fan C-M, Gao G-Q, Wang R-X, Xu A-W
23. Doh H, Kim HY, Kim GS, Cha J, Park HS, Ham HC, Yoon SP, (2014) J Mater Chem A 2:4429–4437
Han J, Nam SW, Song KH, Yoon CW (2017) ACS Sustain Chem 56. Fu Q, Saltsburg H, Flytzani-Stephanopoulos M (2003) Science
Eng 5:9370–9379 301:935–938
24. van Deelen TW, Hernández Mejía C, de Jong KP (2019) Nat Catal 57. Zhu J, Su Y, Chai J, Muravev V, Kosinov N, Hensen EJM (2020)
2:955–970 ACS Catal 10:11532–11544
25. Huang C, Yu Y, Yang J, Yan Y, Wang D, Hu F, Wang X, Zhang 58. Montini T, Melchionna M, Monai M, Fornasiero P (2016) Chem
R, Feng G (2019) Appl Surf Sci 476:928–936 Rev 116:5987–6041
26. Furusawa T, Kuribara H, Kimura K, Sato T, Itoh N (2020) Ind 59. Sonström P, Birkenstock J, Borchert Y, Schilinsky L, Behrend
Eng Chem Res 59:18460–18470 P, Gries K, Müller K, Rosenauer A, Bäumer M (2010) Chem-
27. Yu Y, Gan Y-M, Huang C, Lu Z-H, Wang X, Zhang R, Feng G CatChem 2:694–704
(2020) Int J Hydrog Energy 45:16528–16539 60. Nakamura I, Negishi N, Kutsuna S, Ihara T, Sugihara S,
28. Feng J, Liu L, Ju X, Wang J, Zhang X, He T, Chen P (2021) Takeuchi K (2000) J Mol Catal A Chem 161:205–212
ChemCatChem 13:1552–1558 61. Ogura Y, Sato K, Miyahara SI, Kawano Y, Toriyama T, Yama-
29. Shrestha S, Yeung CMY, Nunnerley C, Tsang SC (2007) Sens moto T, Matsumura S, Hosokawa S, Nagaoka K (2018) Chem
Actuator A Phys 136:191–198 Sci 9:2230–2237
30. Kawabe M, Ono H, Sano T, Tsuji M, Tamaura Y (1997) Energy 62. Funke H, Scheinost AC, Chukalina M (2005) Phys Rev B
22:1041–1049 71:094110
31. Šulcová P (2005) J Therm Anal Calorim 82:51–54 63. Ribeiro J, Tremiliosi-Filho G, Olivi P, de Andrade AR (2011)
32. Bernal S, Botana FJ, Cifredo G, Calvino JJ, Jobacho A, Rodri- Mater Chem Phys 125:449–460
guez-Izquierdo JM (1992) J Alloys Compd 180:271–279 64. Ogawa T, Kobayashi Y, Mizoguchi H, Kitano M, Abe H,
33. Thangadurai V, Huggins RA, Weppner W (2001) J Solid State Tada T, Toda Y, Niwa Y, Hosono H (2018) J Phys Chem C
Electrochem 5:531–537 122:10468–10475
34. Borchert Y, Sonström P, Wilhelm M, Borchert H, Bäumer M 65. Sarma DD, Rao CNR (1980) J Electron Spectrosc Relat Phenom
(2008) J Phys Chem C 112:3054–3063 20:25–45
35. Huang PX, Wu F, Zhu BL, Li GR, Wang YL, Gao XP, Zhu HY, 66. Siakavelas GI, Charisiou ND, AlKhoori S, AlKhoori AA,
Yan TY, Huang WP, Zhang SM, Song DY (2006) J Phys Chem B Sebastian V, Hinder SJ, Baker MA, Yentekakis IV, Polychro-
110:1614–1620 nopoulou K, Goula MA (2021) Appl Catal B 282:119562
36. Ayodele BV, Khan MR, Cheng CK (2016) Clean Technol Environ 67. Mekki A, Ziq KA, Holland D, McConville CF (2003) J Magn
Policy 19:795–807 Magn Mater 260:60–69
37. Jiang N, Zhou X, Jiang Y-F, Zhao Z-W, Ma L-B, Shen C-C, 68. Ogasawara H, Kotani A, Potze R, Sawatzky GA, Thole BT
Liu Y-N, Yuan C-Z, Sahar S, Xu A-W (2018) RSC Adv (1991) Phys Rev B 44:5465–5469
8:17504–17510

13
Ru Nanoparticles on ­Pr2O3 as an Efficient Catalyst for Hydrogen Production from Ammonia… 1181

69. Furusawa T, Shirasu M, Sugiyama K, Sato T, Itoh N, Suzuki N Publisher’s Note Springer Nature remains neutral with regard to
(2016) Ind Eng Chem Res 55:12742–12749 jurisdictional claims in published maps and institutional affiliations.
70. Marco Y, Roldán L, Armenise S, García-Bordejé E (2013) Chem-
CatChem 5:3829–3834
71. Morgan DJ (2015) Surf Interface Anal 47:1072–1079

Authors and Affiliations

Xilun Zhang1,2 · Lin Liu1 · Ji Feng1,2 · Xiaohua Ju1 · Jiemin Wang1,3 · Teng He1 · Ping Chen1

1 3
Dalian National Laboratory for Clean Energy, State Key Zhang Dayu School of Chemistry, Dalian University
Laboratory of Catalysis, Dalian Institute of Chemical of Technology, Dalian 116024, China
Physics, Chinese Academy of Sciences, Dalian 116023,
China
2
University of Chinese Academy of Sciences, Beijing 100049,
China

13

You might also like