You are on page 1of 13

Applied Catalysis B: Environmental 286 (2021) 119884

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

Boosting hydrogen production from steam reforming of ethanol on nickel


by lanthanum doped ceria
Zhourong Xiao a, 1, Chan Wu a, 1, Li Wang a, b, Jisheng Xu a, Qiancheng Zheng a, Lun Pan a, b,
Jijun Zou a, b, Xiangwen Zhang a, b, *, Guozhu Li a, b, *
a
Key Laboratory for Green Chemical Technology of Ministry of Education, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
b
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China

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

Keywords: Stable and efficient hydrogen production has always been the focus of industry, and steam reforming of ethanol
La doped ceria (SRE) is one of the most robust, renewable and cheap technologies. To develop excellent SRE catalysts highly
Metal-support interaction resisting coke formation and metal sintering, La, Tb, and Zr have been doped as a third metal into Ni-CeO2
Ethanol steam reforming
catalyst using a sol-gel method. The size and dispersion of Ni nanoparticles, surface oxygen vacancies and metal-
Hydrogen production
supported interaction have been finely tuned, with which the catalytic performance is closely associated. La-
doped ceria supported Ni (Ni-CeLa0.20) exhibited superior catalytic performance, which is ascribed to its large
Ni active surface area, abundant oxygen vacancies and suitable metal-supported interaction. DFT calculations
showed that the doping of La into Ni-CeO2 can promote the generation of oxygen vacancies and facilitate the
dissociation of water. On Ni-CeLa0.20, complete conversion of ethanol and stable hydrogen production have been
maintained in the 3000-min test. The coke deposition is only 0.48 mgc/gcat h at 600 ◦ C under the GHSV of 55920
mL/gcat∙h.

1. Introduction
2CO → C + CO2 (4)
Nowadays most of the energy demands are provided by fossil fuels CH4 → C + 2H2 (5)
which causes serious resource and environmental problems [1–3]. In
this scenario, exploring cheap, sustainable, and environmentally benign C2H4 → Coke (6)
energy sources has captured much attention to replace fossil fuels [3].
C + H2O→ CO + H2 (7)
Hydrogen is an ideal alternative fuel because it is efficient and clean
with the harmless water as the only by-product. Production of H2 from To gain insight into the reaction mechanism of SRE and to achieve a
steam reforming of renewable oxygenated hydrocarbons, such as high and stable hydrogen yield, various catalytic systems have been
ethanol (Eq. (1)), is a promising technology. Because the feedstock can extensively investigated [2]. Recently, many catalysts based on Pt, Ru,
be obtained from the fermentation of biomass, avoiding fossil source Co, or Ni were developed for efficient SRE reactions [5–11]. Compared
problems [4]. However, steam reforming of ethanol is a very complex to the expensive noble metals, the Ni-based catalysts with low cost are
process that involves many parallel reactions like dehydration (Eq. (2)), considered as good candidates due to their high activities towards the
decomposition (Eq. (3)) and carbon formation reaction (Eqs. (4)–(6)). rupture of C–C and C–H bonds. However, the main challenges are their
low hydrogen selectivity and bad stability due to serious deposition of
C2H5OH + 3H2O → 6H2 + 2CO2 (1)
carbonaceous species and easy sintering of Ni particles [12–17].
C2H5OH → C2H4 + H2O (2) Designing new Ni-based catalysts to effectively prevent deactivation
is an urgent issue [18]. Previous studies showed that the synergy of
C2H5OH → CH4 + CO + H2 (3) metal and support is the key to guarantee high activity and stability in

* Corresponding authors at: Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China.
E-mail addresses: zhangxiangwen@tju.edu.cn (X. Zhang), gzli@tju.edu.cn (G. Li).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.apcatb.2021.119884
Received 30 July 2020; Received in revised form 6 December 2020; Accepted 29 December 2020
Available online 11 January 2021
0926-3373/© 2021 Elsevier B.V. All rights reserved.
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

ethanol steam reforming [18–20]. For this reason, the interaction be­ 2.2. Characterization
tween metal and support was intensively investigated for an active and
stable SRE process [19,20]. Previous studies showed that Ni-metal XRD were tested on Rigaku D8-Focus diffractometer equipped with
nanoparticles could suppress the formation of filamentous carbon [6, Cu Kα radiation (λ =1.54056 Å) at the 2θ values range of 20◦ - 85◦ with a
13]. The Ni nanoparticles were confined by SBA-15 or ordered meso­ scan rate of 8◦ /min. The N2 adsorption-desorption was collected on
porous alumina [6,13,21–23]. Some deactivations were observed on Micromeritics Tristar 3000 analyzer at -196 ◦ C. 0.12 g of the sample was
these catalysts due to coke formation because of the inert support and first heated at 300 ◦ C under vacuum for 4 h. The specific surface area of
weak metal-support interaction. Ceria-based materials received great the sample was calculated by BET method, and pore volume and average
attentions as the support of nickel for hydrocarbon reforming owing to pore diameters were obtained from BJH method using desorption data.
their oxygen storage capacity (OSC) and oxygen mobility (OM) [24–26]. H2-TPR and O2-TPO measurements were performed on an AMI-300 in­
Zhou et al. showed that strong interaction between Ni and ceria in the strument. In a H2-TPR typical test, 0.1 g of the sample was first degassed
Ce1-xNixOy system could effectively turn the electronic structure of at 400 ◦ C for 1 h, then cooled to 50 ◦ C, and finally a H2-Ar (10 vol%, 30
nickel and produce rich active sites, leading to the efficient cleavage of mL/min) was introduced. Then the sample was raised to 800 ◦ C at a
C–C and C–H bonds during SRE [20]. In addition, the OSC and OM heating rate of 10 ◦ C/min. For the test of O2-TPO, 0.03 g of the spent
properties of ceria could further be enhanced by introducing a dopant catalyst was degassed at 150 ◦ C under Ar atmosphere, then a O2-Ar (5
[5,7,19,27–29]. Liu et al. showed that Ni and W co-doped ceria vol%, 30 mL/min) was introduced after cooled to 50 ◦ C. Then the
possessed abundant oxygen vacancies that helped partial dissociation of sample was heated to 800 ◦ C at a rate of 10 ◦ C/min. H2 chemisorption
water [19]. Our group [7] showed that good catalytic performance for (H2-TPD) was recorded in an Autochem 2920 (Micromeritics). The as-
ethanol steam reforming could be achieved by introducing praseodym­ prepared sample was reduced with H2/Ar (10 vol%, 30 mL/min) at
ium (Pr) into ceria. 500 ◦ C for 1 h, then cooled to 50 ◦ C under Ar flow. H2 was introduced to
Previous results spurred us to develop new solid oxides of doped the as-reduced catalysts until H2 saturation. Then, the temperature was
ceria supported Ni to finely tune the structure of the catalyst and raised to 800 ◦ C at a rate of 10 ◦ C/min. According to the amount of
effectively realize high-performance SRE. Herein, the elements of La, Tb, chemisorbed H2, the Ni dispersion was calculated, in which we assumed
and Zr were doped into ceria to prepare new Ni-based catalysts for that one surface Ni atom adsorbed one hydrogen atom. EPR was con­
hydrogen production by ethanol steam reforming. In order to under­ ducted on Bruker E500 spectrometer at -196 ◦ C. ICP-OES was measured
stand the structure-performance relationship, physicochemical proper­ on an Optima 5300DV (Perkin Elmer). UV–vis DRS were collected on a
ties of the catalysts were finely tuned and their catalytic performances Hitachi U-3010 spectrometer. Raman spectra were recorded on a DXR
were evaluated. The structures of the catalysts have been well charac­ Microscope system with 532 nm as laser source. SEM images were
terized in an attempt to determine the key factors affecting the catalytic collected on FEI Nanosem 430. The TEM images of the catalysts were
activity and stability for SRE. In addition, DFT + U simulations were collected on a JEM-2100 F microscope under a working voltage of 200
conducted to deeply explore the mechanisms of oxygen vacancy for­ kV. TG was recorded on TQ-500. The experiments of TG were performed
mation and water dissociation on the catalysts. in air flow (50 mL/min) from room temperature to 800 ◦ C at a heating
rate of 10 ◦ C/min. The coke deposition rate was calculated based on the
2. Experimental data of TG and the reaction time. The coke amount in the spent catalyst
was calculated using the mass loss percentage from the TG data. In order
2.1. Preparation of the catalysts to eliminate the effect of mass increase by nickel oxidation, the data that
after complete oxidization of Ni species were used. Then, coke deposi­
In this work, the catalysts were prepared by a sol-gel (SG) method tion rate was calculated via dividing the coke amount by the reaction
using citric acid (AR) as the chelating agent [7]. The as-prepared cata­ time. The XPS were measured on an ESCALAB 250Xi spectrometer with
lysts were named as Ni-CeO2, Ni-La2O3 and Ni-CeLn0.2 (Ln = La, Tb, Zr), Al Kα radiation (hν = 1486.6 eV). The binding energy of the C 1s peak
respectively. The Ni loading was fixed to 10 wt% and the mole ratio of from adventitious carbon was set at 284.8 eV to correct the charge
Ce/Ln was 4 for all the catalysts. The metal precursors were Ni effects.
(NO3)2∙6H2O, La(NO3)3∙6H2O, Ce(NO3)3∙6H2O, Zr(NO3)4∙2H2O, Tb
(NO3)3∙6H2O with AR purity, which were purchased from Beijing 2.3. Catalyst evaluation
HWRK Chem Co. The detailed procedure for the synthesis of Ni-CeLn0.2
is as follows: the metal precursors of Ni(NO3)2∙6H2O, Ce(NO3)3∙6H2O The catalytic steam reforming of ethanol was carried out in a stain­
and Ln(NO3)3∙xH2O were simultaneously added in a 500 ml beaker, and less steel fixed-bed tubular reactor (6 mm ID) at atmospheric pressure
then citric acid solution was introduced. The amount of citric acid is loaded with a mixture of 0.25 g catalyst (20–40 mesh) and 1.2 g silicon
twice that of metal ions. In particular, for the synthesis of Ni-CeLa0.2, carbide (20–40 mesh). Prior to the reaction, the catalyst was activated
5.453 g of Ce(NO3)3∙6H2O, 1.359 g of La(NO3)3∙6H2O and 1.486 g of Ni by H2 for 1 h at 500 ◦ C. Then N2 (60 mL/min) was used as a carrier gas
(NO3)2∙6H2O were mixed with citric acid solution (20 mL). The amount and purged the pipeline. The ethanol solution (0.255 mL/min) was
of citric acid in the solution was 7.991 g. The resulting solution was introduced to the reactor by high performance liquid chromatography
stirred for 12 h at 60 ◦ C, and then heated to evaporate water and then pump. The molar ratio of water to ethanol was fixed to 4 and gas hourly
dried at 120 ◦ C to form spongy solid. The solid was grinded, placed in a space velocity (GHSV) of ethanol solution was set to 55920 mL/gcat∙h.
quartz boat and calcined at 600 ◦ C in air for 4 h. In addition, a control The mixed liquid was heated and evaporated at 300 ◦ C in an evaporator
catalyst of CeLa0.2 supported Ni was synthesized using the traditional before entering the catalytic bed. A gas chromatograph (Varian Micro-
impregnation (IMP) method, which is similar to those reported previ­ GC 490), which was equipped with three columns of activated
ously [1,11], and the catalysts were denoted as Ni-CeLa0.2-IMP. Firstly, alumina, PPU and 5 Å molecular sieve and with the detectors of TCD,
CeLa0.2 was synthesized using a similar procedure of Ni-CeLa0.2 prepa­ was used to analyze the gas products online. Activated alumina column
ration. Then, a solution of Ni(NO3)2∙6H2O (2.04 mol/L, 2.5 mL) was was used for detecting light hydrocarbons such as ethane and ethylene.
added in 2.700 g of the as-prepared CeLa0.2. The mixture was kept 12 h The PPU column was adopted to analyze CO2. The 5 Å molecular sieve
at room temperature. The obtained solid was dried at 120 ◦ C for 24 h and column was employed for the analyses of H2, N2, CH4, CO. The con­
calcined at 600 ◦ C for 4 h. In order to characterize the structure of the version of ethanol and gas product rate were calculated using Eqs. (8)–
reduced catalyst, in-situ reduction by H2 at 500 ◦ C for 1 h was carried (9) [29].
out.

2
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Fout out out


CO + FCO2 + FCH4
Ethanol conversion (wt%) = in
× 100 (8) ∆E = EH∗+OH∗ − EH2 O∗ (12)
Fethanol
EH2 O∗ and EH∗+OH∗ represent the energies of optimized initial and final
/
AH2 ⋅fH2 states for water dissociation, respectively.
Gas product rate (μmol min) R(H2 ) = × R(N2 ) (9)
AN2 ⋅fN2

Fethonal + Fgas 2.5. Description of the relationship between structure and activity
Carbon balance (wt%) = cout cout
× 100 (10)
Fethanol
cin
The relative oxygen vacancy concentration was calculated based on
where AH2 represents GC peak area of hydrogen, fH2 represents the the intensity of EPR peak. The peak intensity of Ni-CeO2 possessing the
response factor of hydrogen, and RN2 is the nitrogen flow rate (60 mL/ lowest oxygen vacancy concentration was set as the standard and
min). defined as 1. Exceptionally, the relative oxygen vacancy concentration
of Ni-CeLa0.2-IMP was calculated based on Raman spectra. The relative
metal-support interaction was calculated based on the reduction peak
temperature. Ni-CeLa0.2-IMP possessing the lowest reduction peak
2.4. DFT simulation
temperature was set as 1. The relative activity was calculated based on
the reaction data in Fig. 6a. The ethanol conversion on Ni-CeLa0.2-IMP
In terms of electronic structure of pure ceria and metal doped ceria,
was defined as 1 due to its lowest value at 550 ◦ C.
DFT + U method was performed. According to previous studies, the U
values for Ce is 4.5 eV [30–35].
The ceria surface was modeled by (3 × 3) unit cells as a nine atomic 3. Results and discussion
layer (three O-Ce-O tri-layers) slab with 15 Å vacuum space. The energy
cutoff of 450 eV was adopted, and the Monkhorst Pack scheme was 3.1. Characterization of the as-prepared fresh catalysts
sampled with (2 × 2 × 1) k-point. Structure optimization were allowed
to relax until the forces of all atoms were less than 0.02 eV/Å. Fig. 1 displays the XRD patterns of the reduced catalyst. The fluorite
Formation energy of the oxygen vacancy (Evac structure of ceria could be clearly observed in all the samples including
f ) is formulated and
pure ceria, La doped ceria and the catalysts of metal (La, Tb, or Zr) doped
defined as Eq. (11) which could be seen [7].
ceria supported Ni [7]. As shown in Fig. 1a, the peaks corresponding to
Efvac = E(CeO2− x ) + 1 2 E(O2 ) − E(CeO2 )
/
(11) La2O3, TbO2, and ZrO2 did not appear in the as-prepared catalysts,
indicating the formation of solid solution. Fig. 1b displays the XRD
The dissociation energy of water (∆E) was defined by Eq. (12) [36,37]. patterns of La2O3 supported Ni (Ni-La2O3) and La doped ceria supported

Fig. 1. (a–b) Wide-angle XRD patterns of the reduced catalysts. The peaks labeled by ▴, ◆, ♣ and Δ are the characteristic peaks of CeO2, Ni, La2O3 and LaNiO3,
respectively. (c) The enlarged XRD patterns in the 2θ range of 43˚-46˚. (d) The enlarged XRD patterns in the 2θ range of 27˚-30˚.

3
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Ni prepared by the IMP method (Ni-CeLa0.2-IMP). For the sample of elemental distributions of O, Ce, La and Ni are presented in Fig. 2b–e.
Ni-La2O3, the crystalline of La2O3 as well as some LaNiO3 was detected, The elements of O, Ce and La were uniformly distributed, which in good
which corresponds to the perovskite phase with rhombohedral structure agreement with the SEM images (Fig. S2). The uniformly loaded small Ni
(JCPDF-330711) [27]. Meanwhile, the absence of Ni peak at 44.5◦ metal particles possessed enhanced interfacial contacting with the
suggests (Fig. 1c) that Ni nanoparticles were amorphous or the particle CeLa0.2 support. For Ni-CeLa0.2-IMP, the elements of O, Ce and La were
size was too small to be detected [29]. Because all the catalysts were also uniformly distributed, which benefited from the preparation of the
prepared by H2 reduction at high temperature, amorphous Ni nano­ CeLa0.2 support by the SG method. However, the Ni metal species in Ni-
particles hardly existed. Therefore, it can be deduced that small Ni CeLa0.2-IMP exhibit some agglomeration and uneven distribution, which
nanoparticles were highly dispersed on corresponding support for the Ln is consistent with our XRD results (Fig. 1c) and the previous reports [1,
(La, Tb, Zr)-doped catalysts prepared by the SG method. In comparison, 11]. This may be caused by weak interaction between Ni metal and the
the Ni peak at 44.4◦ is strong for Ni-CeLa0.2-IMP as shown in the Fig. 1b support, which will be further discussed in latter section based on the
and c. It indicates the presence of large Ni particles in Ni-CeLa0.2-IMP. TPR analysis.
The results illustrate that the SG method is superior to the IMP method in ICP was employed to accurately quantify the nickel amount in the
decreasing Ni particle size and improving their dispersion on the sup­ catalyst (Table 2). All the catalysts loaded reasonable amount of nickel
port. Moreover, we speculate that Ni particles and the support (CeO2, metal, which was very close to the theoretical value of 10 wt %. The
metal doped CeO2, or La2O3) strongly interacted in the catalyst prepared amount of dopant was also similar to the theoretical value, which
by the SG method. The aggregation of Ni nanoparticles during matches well with the SEM-EDX analysis.
high-temperature reduction will be effectively suppressed by the strong The reduction behaviors of the as-prepared catalysts were studied by
metal-support interaction. H2-TPR, and the results are presented in Fig. 3. For the freestanding NiO,
Compared with the peak 2θ value of pure CeO2, corresponding a small reduction peak at 317.8 ◦ C was detected followed by a sharp
diffraction 2θ value of La doped ceria shifted to lower angle. It is peak at 382.6 ◦ C due to the big particle size. Two reduction peaks rep­
attributed to the fact that the radius of La3+ (1.10 Å) is larger than that of resented the reduction behavior of NiO undergoing the reaction of
Ce4+ (0.97 Å), which leaded to the crystal lattice expansion [5]. The NiO→Niσ+→Ni◦ [39]. For the Ni catalysts supported by ceria or metal
peaks of La doped ceria were weaker and broader than those of pure (La, Tb, Zr) doped ceria, three reduction peaks (α, β, γ) were presented.
ceria, which indicates that La species were homogeneously doped into The first peak (α) located at low reduction temperature is responsible for
ceria lattice. The same trend was observed for Ni-CeO2 and Ni-CeLa0.2. It surface adsorbed oxygen species, which is easily to be reduced under H2
is expected that the anti-sintering ability of the catalyst is improved by atmosphere [40]. The second peak (β) was ascribed to the reduction of
the lattice doping of lanthanum. In comparison, the sample of Zr doped bulk NiO that weakly interacts with the support. The third peak (γ) was
ceria supported Ni showed that its diffraction peaks slightly shifted to attributable to the reduction of relative small NiO particles strongly
higher angle. Similarly, it is ascribed to the smaller radius of Zr4+ (0.84 interacting with ceria [7,39]. Compared with the peaks of Ni-CeO2, the
Å) than that of Ce4+ (0.97 Å), leading to the contraction of the crystal reduction peaks of β and γ shifted to higher temperature for Ni-CeLa0.2
lattice. The diffraction peaks of Tb doped ceria supported Ni also slightly and Ni-CeZr0.2. It illustrates the enhanced metal-support interaction by
shifted to lower angle. The ionic radius of Tb3+ is 1.04 Å, while that of the doping of La or Zr into ceria. When Tb doped ceria was used as the
Tb4+ is 0.88 Å [38]. Therefore, both valance states of Tb were present in support, the peaks of β and γ moved to lower reduction temperature. The
the sample, and Tb3+ was dominant to lead lattice expansion. All the cell areas of γ peak were all increased by the introduction of La, Tb, and Zr
lattice parameters were calculated and presented in Table 1. The into ceria using the SG method. For Ni-CeLa0.2-IMP prepared by the
changes of cell lattice parameter show that La or Tb doping expanded impregnation method, the main reduction peak was the β peak. It il­
the crystal lattice of ceria, while Zr doping caused lattice contraction. lustrates the nickel oxides species were mainly large NiO particles which
Table 1 summarizes the textural properties of the samples. All the possessed relative weak metal-supported interaction. These results are
samples show the type IV isotherm with the characteristic hysteresis in accord with the XRD (Fig. 1) and TEM (Fig. 2) data. For Ni-La2O3, the
loop (Fig. S1), suggesting the presence of mesoporous structure. Their first peak (I) appeared at 416.9 ◦ C and can be ascribed to the reduction
BET surface areas range from 18 m2/g to 30 m2/g. of surface Ni2+ species. The second peak (II) appeared at temperature of
SEM images of Ni-CeLa0.2 are presented in Fig. S2. The EDX results 609.3 ◦ C, indicating the reduction of LaNiO3 to La2Ni2O5. The third peak
show that the elements of O, La, Ce and Ni were all present (Figs. S2b, (III) at 767.7 ◦ C corresponded to the reduction of Ni2+ in LaNiO3 to
2c, 2d and 2e). The elements were uniformly distributed, which metallic Ni [41]. According to the XRD results (Fig. 1), the LaNiO3
benefited from the SG method. As shown in Fig. S2g, the amount of Ni perovskite was formed which needs high reduction temperature.
was 10.11 wt% and the mole ratio of Ce/La was 3.7, which are consis­ H2-TPD was measured to determine Ni dispersion and active Ni
tent with the theoretical values. Similar results were obtained for Ni- surface area. The data are shown in Table 2. Both the active Ni surface
CeO2, Ni-CeZr0.2 and Ni-CeTb0.2 (Figs. S2, S3, S4 and S5), demonstrating area and the Ni dispersion were improved by the introduction of La, Tb,
that Ni nanoparticle were highly dispersed in the Ce-Ln (La, Tb and Zr) or Zr into ceria. It is ascribed to their high proportion of Ni species that
solid oxides, which well matched with the XRD results. strong interact with the support. The catalyst of Ni-CeLa0.2 possessed the
HAADF images of Ni-CeLa0.2 are shown in Fig. 2a, and corresponding highest Ni active surface area (26.5 m2/g) and Ni dispersion (17.7 %). In
comparison, Ni-CeLa0.2-IMP showed the active Ni surface area of only
8.4 m2/g and the Ni dispersion of 5.6 %, which are less than one third of
Table 1
Ni-CeLa0.2. Therefore, the SG method can effectively improve the
Textural properties of the samples.
dispersion of Ni and increase Ni active surface area [5], as demonstrated
Catalysts BET Surface Pore Volume Average Pore Cell lattice by XRD (Fig. 1), TEM (Fig. 2) and TPR (Fig. 3). Previous study showed
Area (m2/g) (cm3/g) Size (nm) Parameter (Å)a
that the dispersion and anti-sintering performance of Ni particles could
Ni-CeO2 29.37 0.056 3.98 5.412 be effectively promoted by their strong and intimate interacts with the
Ni-CeLa0.2 20.71 0.042 8.71 5.442 support [19]. Moreover, the Ni particles in close contact with ceria could
Ni-La2O3 23.30 0.041 8.76 –
Ni-CeLa0.2- 16.63 0.076 16.44 5.440
guarantee the sufficient supply of OH from the surrounding ceria, which
IMP can enhance carbon gasification [20]. Therefore, it is expected that the
Ni-CeTb0.2 16.06 0.052 9.81 5.423 highly dispersed Ni particles supported on metal doped ceria show good
Ni-CeZr0.2 18.23 0.048 8.62 5.395 performance for SRE.
a
The cell lattice parameter was calculated by (111) facets of ceria using the The nature of catalyst support directly influences the catalytic per­
Scherrer equation. formance in the SRE process. For the redox supports of ceria and metal-

4
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Fig. 2. a) HAADF image of Ni-CeLa0.2, b-e) corresponding elemental distributions of O, Ce, La, and Ni, and f) the overlapping of Ce, La, Ni elements. g) HAADF of Ni-
CeLa0.2-IMP, h-k) corresponding elemental distributions of O, Ce, La, Ni, and l) the overlapping of Ce, La, Ni elements.

doped ceria, their oxygen vacancies help to dissociate water and form mode). After metal doping (La, Tb, or Zr) (Figs. S6 and 4 a), the shift of
surface hydroxyl groups [20,42]. In order to gain insight into the oxygen the F2g peak suggests the formation of homogeneous solid oxide, which
vacancies and oxygen storage-release capability (OSC) in the catalysts, is also confirmed by the XRD data (Fig. 1). The weak peaks of F2g mode
Raman spectroscopy was adopt due to its high sensitivity to Ce-O bond for the Tb and Zr doped samples are ascribed to their stronger fluores­
vibration and lattice defects [28]. For pure ceria (Figure S6), the sharp cence effects which reduced the detectable of Raman signal [43]. No
peak at 465 cm− 1 is ascribed to the fluorite structure vibration (F2g Raman response was observed for Ni-La2O3, which is ascribed to the non

5
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Table 2
Bulk phase composition, Ni dispersion and active Ni surface area of the reduced catalysts.
Catalysts Ni (wt%)a Ce (wt%)a La (wt%)a Tb (wt%)a Zr (wt%)a Ni dispersion (%)b Ni active surface areas (m2/gNi)c

Ni-CeO2 9.92 90.08 – – – 12.5 18.7


Ni-CeLa2 9.99 71.95 18.06 – – 17.7 26.5
Ni-La2O3 9.96 – 90.04 – – 16.3 24.4
Ni-CeLa2-IMP 9.97 71.86 18.17 – – 5.6 8.4
Ni-CeTb2 9.98 71.46 – 18.56 – 16.9 25.3
Ni-CeZr2 9.94 72.03 – – 17.03 15.2 22.7
a
Calculated based on ICP results.
b
The dispersion of Ni was estimated by the data of H2-chemisorption, which was equal to the ratio of Ni atoms on the surface and total Ni atoms.
c Ni atoms on the surface SNi ∗ M 2
D= = p . Where SNi is the represents the Ni metal surface area (m /gNi), N is Avogadro’s number, A is the metal surface area per
Total Ni atoms N∗A∗
100
atom (for Ni, A = 6.494*10− 20 m2/atom).

Fig. 3. TPR profiles of the as-prepared catalysts.

cubic fluorite-type crystal structure. Besides the F2g mode, two other changes. When Ce4+ was substituted by aliovalent ions (Tb3+), more
Raman peaks located at 550 cm− 1 and 610 cm− 1 were detected, which oxygen vacancies were generated to ensure the charge neutrality in the
can be differentiated from that of NiO (513 cm− 1). The two peaks are ceria lattice [44]. In particular, the highly intensive signal of Ni-CeLa0.2
assigned to oxygen vacancies in the lattice [7,44]. Both Raman modes suggest the presence of abundant oxygen vacancies. It is explained by
come from the substitution of Ce4+ by other dopant (La, Zr, Tb or Ni) or the large crystal lattice expansion caused by the extensive substitution of
the lattice changes originating from the ionic radii disparity between Ce4+ by aliovalent ions (La3+) with different ionic radius and valence.
Ce4+ and the dopant (La, Zr, Tb or Ni) [1,7,43]. Compared with pure The XPS spectra of O-1 s and Ce-3d are shown in Figs. 4c and S8a.
ceria, the relative content of oxygen vacancies was sharply incerased The shifting of O-1 s and Ce-3d peaks to lower binding energy demon­
after the doping by La, Zr, Tb or Ni. strates the formation of some oxygen defects on the surface [37]. In the
UV–vis diffuse reflectance spectroscopy was used to illustrate the O-1 s core level region (Fig. S8c), the sharp peak centered at 529 ± 0.5
electronic state of the metal ions, and the data are shown in Fig. S7. Two eV (OI) is attributed to the lattice oxygen, and the peak at 531 ± 1 eV
peaks located at 233 nm and 270 nm could be obviously observed, which (OII) is associated with chemisorbed oxygen species (Oα) and hydroxyl
can be assigned to O2− →Ce3+ and O2− →Ce4+ charge-transfer transi­ species (Oβ) on the surface [40]. The OII peak was observed in most
tions, respectively [44,45]. Compared with Ni-CeO2, the metal (La, Tb, synthesized materials, and generally regarded as the presence of oxygen
Zr) doped Ni-CeO2 exhibited higher peak intensities of the (O2− →Ce4+) vacancies. With the doping of La, Tb, or Zr, the peak intensity of OII was
and (O2− →Ce3+) bands, which is explained by the increasing number of increased, indicating the existence of abundant oxygen vacancies. The
oxygen vacancies [45]. The peak intensities for (O2− →Ce4+) and relative concentration of oxygen vacancies is as follows: Ni-CeLa0.2 >
(O2− →Ce3+) bands in Ni-CeLa0.2-IMP was low, indicating that many Ni Ni-CeTb0.2 > Ni-CeZr0.2 > Ni-CeO2. These results match well with the
atoms were not inserted into ceria lattice. EPR analysis. Moreover, similar results could be obtained in terms of O-1
To further characterize the oxygen vacancies, the measurements of s and Ce-3d XPS spectra for the catalysts treated in hydrogen at 500 ◦ C
EPR and XPS were conducted. As shown in Fig. 4b, Ni-CeO2 and metal for 1 h (Figs. S8c and S8d). Compared with Ni-CeO2, Ni-CeLa0.2 doped
(La, Tb, Zr) doped Ni-CeO2 showed symmetrical EPR signals (g = 1.996) with La possessed increased contents of both oxygen vacancies and Ce3+.
due to the existence of oxygen vacancies [37]. Obviously, compared the The concentration of Ce3+ was calculated based on the results of XPS
un-doped sample (Ni-CeO2), metal (La, Tb, Zr) doped samples possessed curve fitting (Fig. S8d) following the assignment reported by Trudeau
stronger EPR signals. The doping of La, Tb, or Zr into ceria could et al. [46]. Specifically, the ratio of Ce3+/(Ce3++Ce4+) was increased to
effectively increase the concentration of oxygen vacnacies. The 0.23 in Ni-CeLa0.2 compared with that in Ni-CeO2 (0.16).
increased EPR singal by Zr doping is attributed to the contraction of the In order to know more about the information of oxygen vacancies,
crystal lattice due to the different ionic radii of Ce and Zr. Because Zr4+ DFT + U method was adopted to simulate the redox behaviors of various
and Ce4+ have the same valence state. The enhanced oxygen vacancy ceria surfaces (bare CeO2, La-doped CeO2, Ni-doped CeO2 and Ni-La-co-
concentation by Tb doping is ascribed to the existing of two valence doped CeO2) [33,47]. Previous studies showed that surface oxygen va­
states of Tb (Tb4+, Tb3+), which caused more complex crytal lattice cancy can help to dissociate water and oxidize carbon [34,48]. With this

6
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Fig. 4. a) Raman profiles, b) EPR spectra, and c) O 1s XPS spectra of the as-prepared catalysts before H2 treatment.

consideration, oxygen vacancies are demanded on the ceria supported breaking C-C at low temperature. With the increase of reaction tem­
Ni catalyst to enhance catalytic performance for ethanol steam reform­ perature, both ethanol conversion and hydrogen formation rate were
ing [49]. After the optimization of different ceria (111) surfaces, the remarkably increased, which is ascribed to the endothermic property of
formation energies of oxygen vacancies were calculated, and the results SRE reaction. The catalysts prepared by the SG method exhibited higher
are summarized in Fig. S9. CeO2(111) possessed higher formation en­ activity and HPR than that synthesized by the IMP method. Moreover,
ergy of oxygen vacancy than La-CeO2(111) and Ni-CeO2(111). The the ethanol conversion and HPR were further improved by the doping of
formation energy of oxygen vacancies on Ni-La-CeO2(111) is only − 0.96 La, Tb, or Zr into the catalyst. Ni-CeLa0.2, Ni-CeTb0.2, and Ni-CeZr0.2
eV, which is much lower than that of Ni-CeO2(111) (− 0.24 eV). It in­ exhibited superior performance in comparison to Ni-CeO2 without any
dicates that La doping can promote the generation of oxygen vacancies. metal doping. In particular, the conversion of ethanol at the temperature
Thus, the conclusions drawn based on the experimental data of Raman of 550 ◦ C is in the order of Ni-CeLa0.2 (87.6 %) > Ni-CeTb0.2 (82.9 %) >
(Fig. 4a), EPR (Fig. 4b) and XPS (Figs. 4c, d and S8c, d) can be properly Ni-La2O3 (80.7 %) > Ni-CeZr0.2 (75.4 %) > Ni-CeO2 (72.7 %) > Ni-
explained. CeLa0.2-IMP (63.9 %). Similar trend was obtained for the generation rate
Moreover, the dissociation energies of water on CeO2(111), Ni-doped of H2 (Table S1). The temperatures for complete conversion of ethanol
CeO2(111) and Ni-La-co-doped CeO2(111) surface were calculated, and catalyzed by metal (La, Tb, Zr) doped ceria supported Ni are all lower
the results are plotted in Fig. 5. The dissociation of H2O to OH and H on than that over Ni-CeO2 as shown in Fig. 6a.
Ni doped CeO2(111) and Ni-La co-doped CeO2(111) is easy to occur. As The catalytic activities are related with the concentration of oxygen
shown in Fig. 5, the relative dissociation energies (ΔE) of water on vacancies and the state of metal-support interaction as shown in Fig. 6b.
CeO2(111), Ni-CeO2(111) and Ni-La co-doped CeO2(111) surface are As determined by Raman (Fig. 4a), EPR (Fig. 4b), XPS (Fig. 4c), DFT
-0.01 eV, -0.21 eV, and -0.25 eV, respectively. Therefore, the presence of calculations (Figs. 5 and S9) and TPR (Fig. 3), Ni-CeLa0.20 and Ni-
doped Ni and La in ceria plays a crucial role in dissociating water and CeZr0.20 possessed the high Ce3+ ratio, abundant oxygen vacancies as
generating abundant surface hydroxyl groups [19], which will be well as enhanced metal-support interaction compared with Ni-CeO2.
beneficial to SRE reaction. Although the metal-support interaction in Ni-CeTb0.2 was weaker than
that in Ni-CeO2, its higher concentration of oxygen vacancies and rela­
3.2. Catalytic activity and stability of the catalysts tive larger active Ni surface area endowed Ni-CeTb0.2 with high per­
formance for ethanol conversion and hydrogen production (Fig. 6b).
In order to figure out the effect of dopant on catalytic performance, Though abundant oxygen vacancies existed in Ni-La2O3 (Fig. S8c), its
the initial catalytic activities of all the catalysts for SRE were evaluated. excessively strong metal-support interaction (Fig. 3) caused that most of
The reaction temperature was set in the range of 500–650 ◦ C, and the the Ni active sites could not to be utilized. In fact, the oxygen storage
experimental data are shown in Fig. 6a and Table S1. At low reaction release capability and oxygen mobility of La2O3 are weaker than those of
temperature, relative low ethanol conversion and hydrogen production ceria as reported by Cristina et al., [1]. Thus, Ni-La2O3 may exhibit
rate (HPR) were observed on the catalysts prepared by the SG method. relatively high initial activity but bad long-time stability. This deduction
C2 was detected in the product, indicating the weak activity of Ni for will be confirmed by the stability test in latter section (Fig. 7).

7
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Fig. 5. The relative dissociation energies of H2O (△E) on CeO2 (111), Ni-CeO2 (111) and Ni-La co-doped CeO2 (111) surface.

Fig. 6. a) Conversion of ethanol under different reaction temperatures over different catalysts. Reaction conditions: H2O/C = 2, GHSV of C2H5OH solution = 55920
mL/gcat⋅h, P =1 atm, N2 = 60 mL/min. b) The relationship of relative catalytic activity, oxygen vacancy concentration and metal-support interaction.

8
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Fig. 7. The conversion of ethanol and hydrogen production rate with time on stream over (a) Ni-CeO2, (b) Ni-La2O3, (c) Ni-CeLa0.2-IMP, (d) Ni-CeLa0.2, (e) Ni-
CeTb0.2 and (f) Ni-CeZr0.2 at the GHSV of 55920 mL/gcat⋅h for C2H5OH solution, reaction conditions: T =600 ◦ C, H2O/C2H5OH = 4, P =1 atm, N2 = 60 mL/min.

The stability test was carried out at 600 ◦ C and atmospheric pressure caused by some coking, which can be attributed to its relative small
under H2O/C2H5OH ratio of 4 and ethanol solution GHSV of 55920 mL/ amount of oxygen vacancies (Fig. 4b). Oxygen vacancies are necessary
gcat∙h. As shown in Figs. 7 and S10, catalysts prepared by the SG method to maintain stable reaction. On Ni-CeTb0.2, ethanol conversion was
possessed higher activity and stability than that synthesized via the IMP decreased from 100 % initially to 83.6 % finally. The decreased catalytic
method (Ni-CeLa0.2-IMP). For Ni-CeLa0.2-IMP, the conversion of ethanol performance may be ascribed to the sintering of Ni due to its relative
declined quickly from 76.6 % to only 40.7 % after 120 min (Fig. 7c). weak metal-support interaction (Fig. 3). In particular, on the La doped
Similarly, HPR was decreased from 5677.2 μmol/min to 2900.7 μmol/ catalyst (Ni-CeLa0.2), complete conversion of ethanol without any active
min finally. The rapid deactivation of Ni-CeLa0.20-IMP in SRE may be loss has been achieved throughout (3000 min) and the HPR was still as
explained by the serious metal-sintering and severe coking on the high as 8057.3 μmol/min in the final. The highly enhanced ethanol
catalyst. On Ni-La2O3 (Fig. 7b), the conversion of ethanol was decreased steam reforming performance was attributed to the effectively sup­
from 99.2% to 60.5% in the first 300 min, and then gradually declined to pressed coke deposition by the highly dispersed Ni nanoparticles (Fig. 2
57.4 %. The HPR was decreased from 6692.3 μmol/min to 3925.7 μmol/ and Table 2) with suitable metal-support interaction and rich oxygen
min after 1500 min. As catalyzed by Ni-CeO2, the conversion of ethanol vacancies. Previously, Ma et al. [6] pointed out that the catalytic activity
was reduced gradually from 88.8% to 70.1% with the HPR dropped from can be closely related to Ni active surface area. Cristina et al. [1] showed
6245.9 μmol/min to 5179.9 μmol/min in 3000 min. With the intro­ that carbonaceous species could be easily oxidized by the oxygen va­
ducing of dopants (Ni-CeLa0.2, Ni-CeTb0.2, Ni-CeZr0.2), both the con­ cancies on catalyst surface. Moreover, our catalysts were tested in higher
version of ethanol and HPR were increased. For Ni-CeZr0.2, the slightly ethanol solution GHSV of 93200 mL/gcat∙h, and the data are displayed
decreased ethanol conversion from initial 95.8 % to final 72.5 % was in Figs. S11 and S12. The La doped sample (Ni-CeLa0.2) showed much

9
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

higher ethanol conversion, larger hydrogen production rate as well as SRE [7]. In addition, some large Ni particles with the size of up to 20 nm
better stability compared with Ni-CeO2. These results further prove that were detected, which verifies the sintering of nickel under the reaction
La doping into ceria could enhance the catalytic performance of sup­ condition. On the spent Ni-CeLa0.2-IMP, many large Ni particles were
ported Ni catalyst which benefit from highly dispersed of Ni and rich divorced from the support and encapsulated by many layers of graphitic
oxygen vacancies. carbon. Xiao et al. [7] showed that reactant could not access to Ni metal
As reported in previous studies, many novel Ni or promoted Ni-based encapsulated in carbon, which hence ceased the reaction. Thus, the
catalysts with various structures were developed and used in SRE under sharply decreased H2 production and the forced stopped reaction in
different reaction conditions. As summarized in Table S2, the Ni short time on Ni-CeLa0.2-IMP can be explained. For the spent catalysts of
component, water to carbon ratio, reaction temperature, initial activity, Ni-La2O3 and metal doped samples (Ni-CeLa0.2, Ni-CeTb0.2 and
activity loss percentage and carbon deposition are presented and Ni-CeZr0.2), no obvious cover of Ni particles by carbon was observed,
compared with those of the La doped ceria supported Ni (Ni-CeLa0.2) only some Ni particles on the tip of filamentous carbon were detected.
developed in this work. Relatively higher activity, better stability and These results suggest that coking was greatly suppressed by the intro­
less carbon deposition were achieved on Ni-CeLa0.2, which is a duction of La, Tb, or Zr element. Moreover, the small size of Ni particles
competitive candidate for efficient industrial SRE. was still maintained.
Filamentous carbon with the shape of nanotube could be observed in
3.3. Characterization of the spent catalysts Fig. S14. The spent Ni-CeO2 and Ni-CeLa0.2-IMP possessed many carbon
nanotubes (Fig. S14). In comparison, the carbon nanotubes on the spent
To reveal the key factors affecting catalytic performance, the struc­ Ni-CeLa0.2 were much less. The results demonstrate that the catalyst
tures of the spent catalysts were characterized by various technologies. doped by La and prepared by the SG method could effectively suppress
As shown in Fig. 8, any peak that can be assigned to Ni species was not carbon deposition, which is confirmed by TEM (Fig. S13) and XRD
observed on the spent metal-doped catalysts (Ni-CeLn0.2) and Ni-La2O3. (Fig. 8) analyses.
It illustrates that the high dispersion of Ni nanoparticles were main­ TG was adopted to identify the type and amount of carbon in the
tained under the harsh reaction conditions for long time. For the used spent catalysts. The TG curves (Fig. 9a) of the spent catalysts showed a
catalyst of Ni-CeO2, the diffraction peaks of both Ni and CeO2 were slightly rise in the temperature range of 300 ◦ C–400 ◦ C, which was
observed with high intensity, which indicates the sintering of metal Ni caused by the oxidation of the metallic Ni and Ce3+ [44]. The initial
species and ceria support to some extent. These results demonstrate that oxidation of carbonaceous deposits happened at around 350 ◦ C, and no
the metal doping into ceria could effectively improve the stability of the further mass change was detected under the temperature of above 750
catalysts by tuning metal-support interaction [50]. In the TPR profiles,

C. The rate of carbon deposition was calculated, and the results are
the reduction temperature for Ni-CeTb0.2 was lower than that of Ni-CeO2 listed in Fig. 9a. The rates of carbon deposition on different catalysts are
(Fig. 3b), resulting in its higher catalytic performance for SRE. For the in the following order: Ni-CeLa0.2 < Ni-CeZr0.2 < Ni-CeTb0.2 < Ni-CeO2
catalyst prepared by the IMP method (Ni-CeLa0.2-IMP), La doping could < Ni-La2O3 << Ni-CeLa0.2-IMP. Compared with Ni-CeO2, the intro­
greatly increase the concentration of oxygen vacancies (Fig. 4a), but its duction of another metal (La, Tb, Zr) dopant effectively decreased the
weak metal-support interaction leaded to serious sintering of Ni species. coke deposition rate. In particular, the rate of carbon deposition on
The diffraction peak of Ni species with high intensity could be obviously Ni-CeLa0.2 was only 0.48 mgc/gcat h, which is less than one fifth of that
observed for the spent Ni-CeLa0.2-IMP. In addition, new diffraction peak on Ni-CeO2 (2.60 mgc/gcat h). The coke deposition rate on Ni-Ce­
located at 26.4◦ was detected which is assigned to graphitic carbon [29]. La0.2-IMP (212 mgc/gcat h) was 441 times that of Ni-CeLa0.2 (0.48
Previous studies reported that the growth of nickel particles by Ostwald mgc/gcat h), which can be explained by the bad dispersion of Ni in
ripening was one of the major reasons for catalytic deactivation [40]. In Ni-CeLa0.2-IMP that accelerated coke formation. Previous study showed
this work, the sharply decreased catalytic performance of Ni-Ce­ that the performance of the catalysts for SRE deteriorated due to serious
La0.2-IMP for SRE is attributed to the serious sintering of Ni species. carbon deposition [13,51]. Similar situation happened on Ni-Ce­
Apart from the sintering of Ni species, coking is another important La0.2-IMP, which leaded to forced stopped reaction in short time and
factor causing the deactivation of the catalyst. Coke deposition was sharply decreased the reaction activity. These results demonstrate that
characterized by TEM for all the spent catalysts (Fig. S13). On the spent coke could be effectively inhibited by the introduction of anther metal
Ni-CeO2, both filamentous carbon and amorphous carbon were (La, Tb, Zr) into Ni-Ce solid oxide using the SG method, which attributes
observed. In some case, filamentous carbon with Ni particle on its tip to the promoted dispersion of Ni, suitable metal-support interaction,
was generated. These Ni species could still provide some active sites for abundant oxygen vacancies and facilitated water dissociation.
To differentiate the carbonaceous deposits, the derivative of the TG
curves was calculated, and the obtained DTG profiles are shown in
Fig. 9b. The two peaks could be assigned to the oxidization of two types
of carbon species [13]. The lower oxidation temperature of coke (α) is
due to its deposition on metal surface or its amorphous state [6,13]. The
higher oxidation temperature of coke (β) is ascribed to the presence of
crystalline carbon such as graphite, carbon nanotube [6]. On Ni-CeO2,
the α peak was mainly detected. Both the peak intensity and peak area
were larger than those of the metal (La, Tb, Zr) doped samples. It in­
dicates substantial coke depositions on metal Ni surface, which caused
severe deactivation. For the catalysts of Ni-CeLa0.2, Ni-CeZr0.2 and
Ni-La2O3, almost no peak appeared at low oxidation temperature.
Moreover, the peak at higher oxidation temperature for Ni-CeLa0.2 was
hardly to be observed, suggesting the small amount of carbon deposited
on the spent catalyst. In comparison, Ni-CeLa0.2-IMP possessed strong
peaks of the two coke species, especially at high temperature. Therefore,
its severe deactivation can be explained by the serious carbon deposition
on catalyst surface.
Fig. 8. XRD patterns of the spent catalysts. The main peaks of CeO2, Ni, La2O3, Moreover, Raman spectra of the spent catalysts were collected, and
LaNiO3 and carbon are labeled by ♠, ◆, ♣, Δ respectively. the results are shown in Fig. S15. According to Fig. S15a, two obvious

10
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

Fig. 9. TG/DTG profiles of the spent catalysts.

peaks located at 1337 cm− 1 (D band) and 1590 cm− 1 (G band) are was maintained throughout the stability test to achieve high ethanol
observed for all the spent catalysts, indicating the coke deposition on conversion and H2 production.
catalyst surface [40]. Compared with Ni-CeLa0.2-IMP, the catalysts
prepared by the SG method possessed weak peak intensities of D band 4. Discussion
and G band. Meanwhile, compared with Ni-CeO2, the doping of another
metal (La, Tb, or Zr) effectively decreased the peak intensities of D band Based on the results of catalyst structure, DFT simulation and cata­
and G band. These results further confirm that metal dopant could lytic performance, the mechanisms of long-term SRE reaction on Ni-
effectively inhibit coke deposition, which is consistent with the previous CeO2, Ni-CeLa0.2 and Ni-CeLa0.2-IMP are proposed in Scheme 1. The
TG (Fig. 9) analysis. The Raman spectra relating with oxygen vacancies incorporation of La into Ni-CeO2 solid oxide increased oxygen vacancies
were also measured, and the data are depicted in Fig. S15b. Compared in the catalyst (Figs. 4 and S8c and d, and S9) [1,28]. The oxygen va­
with its fresh state, the spent Ni-CeLa0.2-IMP showed sharply reduced cancies helped to activate water for the generation of many OH groups
peak of oxygen vacancies. The oxygen vacancies were maintained well and inhibit the dehydrogenation of CH4 (Eq. (5)) and the accumulation
in the spent catalysts of Ni-CeLn0.2. of carbon deposition [1,13,36]. Therefore, the existence of abundant
Additionally, TPO was also employed to identify the carbon oxi­ oxygen vacancies in the reforming catalyst endowed it high capacity of
dization on the spent catalysts (Fig. S16). Two oxidation peaks at low H2 production (Figs. 6,7, S11, and Table S1) and superior stability with
temperature and high temperature appeared, which are assigned to two less coke deposition (Fig. 9, Figs. S15 and S16). The existence of La in the
different coke species [6,7]. Compared with Ni-CeO2, the samples doped ceria support also helped to improve the dispersion of metallic Ni par­
by La, Tb, and Zr possessed sharply decreased peak intensity, suggesting ticles (Fig. 2 and Table 2). The small Ni nanoparticles interacted strongly
less coke deposited on their surfaces. Meanwhile, the peaks at high with the La-doped ceria support, which were active and stable (Figs. 2
temperature shifted to lower temperature, indicating that the coke was and 3). The highly dispersed Ni nanoparticles provided more active
more easily oxidized by oxygen. Limited carbon deposition on Ni surface metallic sites (Table 2) and abundant metal-support interfacial sites,

Scheme 1. Schematic illustration of ethanol steam reforming over the catalysts of Ni-CeLa0.20, Ni-CeO2 and Ni-CeLa0.20-IMP.

11
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

which are available for activating reactants [42,50]. Moreover, the References
sintering of metal Ni was effectively suppressed by their strong inter­
action with the support (Figs. 1 and 3). Therefore, high ethanol con­ [1] Cristina Pizzolitto, Federica Menegazzo, Elena Ghedini, Giada Innocenti,
Alessandro Di Michele, Giuseppe Cruciani, Fabrizio Cavani, Michela Signoretto,
version and hydrogen production rate were achieved on Ni-CeLa0.2. DFT Increase of ceria redox ability by lanthanum addition on Ni based catalysts for
calculations showed that the presence of doped Ni and La plays a crucial hydrogen production, ACS Sustain. Chem. Eng. 6 (2018) 13867–13876.
role in dissociating water and generating abundant surface hydroxyl [2] D. Li, X. Li, J. Gong, Catalytic reforming of oxygenates: state of the art and future
prospects, Chem. Rev. 116 (2016) 11529–11653.
groups, which should be one of the origins of high activity and stability. [3] J. Sun, Y. Wang, Recent advances in catalytic conversion of ethanol to chemicals,
In comparison, Ni-CeLa0.2-IMP possessing relatively low Ni dispersion ACS Catal. 4 (2014) 1078–1090.
and weak metal-support interaction showed bad SRE performance, [4] L. Li, D. Tang, Y. Song, B. Jiang, Q. Zhang, Hydrogen production from ethanol
steam reforming on Ni-Ce/MMT catalysts, Energy 149 (2018) 937–943.
though rich oxygen vacancies were present. The large Ni particles [5] F. Liu, L. Zhao, H. Wang, X. Bai, Y. Liu, Study on the preparation of Ni–La–Ce oxide
(Table 2 and Fig. 2) were less active and easily coked (Fig. 9, Figure S13, catalyst for steam reforming of ethanol, Int. J. Hydrogen Energy 39 (2014)
S15 and S16). Moreover, the weak metal-support interaction (Fig. 3) 10454–10466.
[6] H. Ma, L. Zeng, H. Tian, D. Li, X. Wang, X. Li, J. Gong, Efficient hydrogen
leaded to serious Ni sintering, and the Ni particles were easily departed
production from ethanol steam reforming over La-modified ordered mesoporous
from the support (Fig. S13). Then, serious deterioration of Ni-Ce­ Ni-based catalysts, Appl. Catal. B: Environ. 181 (2016) 321–331.
La0.2-IMP was observed in a short period of time (Fig. 7). [7] Z. Xiao, Y. Li, F. Hou, C. Wu, L. Pan, J. Zou, L. Wang, X. Zhang, G. Liu, G. Li,
Engineering oxygen vacancies and nickel dispersion on CeO2 by Pr doping for
highly stable ethanol steam reforming, Appl. Catal. B: Environ. 258 (2019),
5. Conclusions 117940.
[8] R. Dai, Z. Zheng, C. Lian, X. Li, X. Wu, X. An, X. Xie, A high performance CeO2@Pt
To summarize, metal (La, Tb, Zr)-doped ceria supported Ni catalysts Beta yolk-shell catalyst used in low temperature ethanol steam reforming for high
purity hydrogen production, Int. J. Energy Res. 43 (2019) 2075–2085.
(Ni-CeLn0.2) have been successfully synthesized by the SG method and [9] Cristian H. Camposa, Gina Pecchi, JoséLuis G. Fierro, Paula Osorio-Vargas,
applied for hydrogen production by SRE. Various characterization Enhanced bimetallic Rh-Ni supported catalysts on alumina doped with mixed
techniques were employed to explore the relationship between catalyst lanthanum-cerium oxides for ethanol steam reforming, Mol. Catal. 469 (2019)
87–97.
structure and its catalytic performance. High Ni surface area and [10] M. Chen, C. Wang, Y. Wang, Z. Tang, Z. Yang, H. Zhang, J. Wang, Hydrogen
enhanced metal-support interaction were achieved on the catalysts production from ethanol steam reforming: effect of Ce content on catalytic
prepared by the SG method compared with the IMP method. Moreover, performance of Co/Sepiolite catalyst, Fuel 247 (2019) 344–355.
[11] M. Greluk, M. Rotko, S. Turczyniak-Surdacka, Enhanced catalytic performance of
the doping of another metal (La, Tb, Zr) into ceria could increase the La2O3 promoted Co/CeO2 and Ni/CeO2 catalysts for effective hydrogen production
concentration of oxygen vacancies and maintain the high dispersion of by ethanol steam reforming, Renew. Energy 155 (2020) 378–395.
Ni by the facilitated metal-support interaction. La-doped ceria supported [12] S.M. de Lima, A.M. da Silva, L.O.O. da Costa, J.M. Assaf, L.V. Mattos, R. Sarkari,
A. Venugopal, F.B. Noronha, Hydrogen production through oxidative steam
Ni prepared by the SG method exhibited high performance for SRE. The
reforming of ethanol over Ni-based catalysts derived from La1-xCexNiO3 perovskite-
high SRE performance for La-doped ceria supported Ni was explained by type oxides, Appl. Catal. B: Environ. 121-122 (2012) 1–9.
DFT simulation. Remarkably reduced formation energy of oxygen va­ [13] D. Li, L. Zeng, X. Li, X. Wang, H. Ma, S. Assabumrungrat, J. Gong, Ceria-promoted
cancies was achieved on Ni and La co-doped ceria (-0.96 eV). In addi­ Ni/SBA-15 catalysts for ethanol steam reforming with enhanced activity and
resistance to deactivation, Appl. Catal. B: Environ. 176-177 (2015) 532–541.
tion, the dissociation of water is facilitated on Ni and La co-doped ceria. [14] C.J. Liu, J. Ye, J. Jiang, Y. Pan, Progresses in the preparation of coke resistant Ni-
The water dissociation energy on Ni and La co-doped ceria is -0.25 eV based catalyst for steam and CO2 reforming of methane, ChemCatChem 42 (2011)
which is much lower than that on bare ceria (-0.01 eV). Consequently, 529–541.
[15] D. Li, M. Tamura, Y. Nakagawa, K. Tomishige, Metal catalysts for steam reforming
the catalyst of Ni-CeLa0.2 provided high hydrogen production rate of tar derived from the gasification of lignocellulosic biomass, Bioresour. Technol.
(8057.3 μmol/min) without any activity loss during the 3000 min test 178 (2015) 53–64.
under the water to carbon mole ratio of 2 and GHSV of 55920 mL/gcat∙h. [16] D. Li, M. Koike, J. Chen, Y. Nakagawa, K. Tomishige, Preparation of Ni–Cu/Mg/Al
catalysts from hydrotalcite-like compounds for hydrogen production by steam
In comparison, Ni-CeLa0.2-IMP seriously deteriorated in 120 min. In reforming of biomass tar, Int. J. Hydrogen Energy 39 (2014) 10959–10970.
addition, the rate of carbon deposition on Ni-CeLa0.20 was sharply [17] Y. Li, L. Zhang, Z. Zhang, Q. Liu, S. Zhang, Q. Liu, G. Hu, Y. Wang, X. Hu, Steam
reduced to only 0.48 mgc/gcat h compared with that on Ni-CeLa0.20-IMP reforming of the alcohols with varied structures: impacts of acidic sites of Ni
catalysts on coking, Appl. Catal. A Gen. 584 (2019) 117162.
(212 mgc/gcat h) and Ni-CeO2 (1.62 mgc/gcat h). [18] S. Li, J. Gong, Strategies for improving the performance and stability of Ni-based
catalysts for reforming reactions, Chem. Soc. Rev. 43 (2014) 7245–7256.
CRediT authorship contribution statement [19] Z. Liu, W. Xu, S. Yao, A.C. Johnson-Peck, F. Zhao, P. Michorczyk, A. Kubacka, E.
A. Stach, M. Fernández-García, S.D. Senanayake, J.A. Rodriguez, Superior
performance of Ni–W–Ce mixed-metal oxide catalysts for ethanol steam reforming:
Zhourong Xiao: Conceptualization, Data curation, Formal analysis, synergistic effects of W- and Ni-dopants, J. Catal. 321 (2015) 90–99.
Investigation, Methodology, Resources, Software, Visualization, Writing [20] G. Zhou, L. Barrio, S. Agnoli, S.D. Senanayake, J. Evans, A. Kubacka, M. Estrella, J.
C. Hanson, A. Martinez-Arias, M. Fernandez-Garcia, J.A. Rodriguez, High activity
- original draft, Writing - review & editing. Chan Wu: Data curation,
of Ce1-xNixO2-y for H2 production through ethanol steam reforming: tuning
Formal analysis, Investigation, Methodology. Li Wang: Methodology, catalytic performance through metal-oxide interactions, Angew. Chemie 49 (2010)
Resources, Writing - review & editing, Supervision, Funding acquisition. 9680–9684.
Jisheng Xu: Methodology, Software, Formal analysis. Qiancheng [21] N. Wang, K. Shen, L. Huang, X. Yu, W. Qian, W. Chu, Facile route for synthesizing
ordered mesoporous Ni-Ce-Al oxide materials and their catalytic performance for
Zheng: Methodology, Software, Formal analysis. Lun Pan: Methodol­ methane dry reforming to hydrogen and syngas, ACS Catal. 3 (2013) 1638–1651.
ogy, Formal analysis, Supervision. Jijun Zou: Methodology, Formal [22] Z. Xiao, C. Wu, L. Li, G. Li, G. Liu, L. Wang, Pursuing complete and stable steam
analysis, Supervision. Xiangwen Zhang: Methodology, Resources, reforming of n-dodecane over nickel catalysts at low temperature and high LHSV,
Int. J. Hydrogen Energy 42 (2017) 5606–5618.
Writing - review & editing, Supervision, Funding acquisition. Guozhu [23] Z. Zhang, X. Hu, L. Zhang, Y. Yang, Q. Li, H. Fan, Q. Liu, T. Wei, C.-Z. Li, Steam
Li: Methodology, Validation, Formal analysis, Investigation, Data reforming of guaiacol over Ni/Al2O3 and Ni/SBA-15: impacts of support on
curation, Writing - review & editing, Supervision, Funding acquisition. catalytic behaviors of nickel and properties of coke, Fuel Process. Technol. 191
(2019) 138–151.
[24] T.S. Moraes, R.C. Rabelo Neto, M.C. Ribeiro, L.V. Mattos, M. Kourtelesis, S. Ladas,
Declaration of Competing Interest X. Verykios, F.B. Noronha, Ethanol conversion at low temperature over CeO2-
Supported Ni-based catalysts. Effect of Pt addition to Ni catalyst, Appl. Catal. B:
Environ. 181 (2016) 754–768.
The authors report no declarations of interest. [25] H. Tian, X. Li, S. Chen, L. Zeng, J. Gong, Role of Sn in Ni-Sn/CeO2 catalysts for
ethanol steam reforming, Chin. J. Chem. 35 (2017) 651–658.
Appendix A. Supplementary data [26] S. Xu, X. Yan, X. Wang, Catalytic performances of NiO-CeO2 for the reforming of
methane with CO2 and O2, Fuel 85 (2006) 2243–2247.
[27] C.A. Franchini, W. Aranzaez, A.M. Duarte de Farias, G. Pecchi, M.A. Fraga, Ce-
Supplementary material related to this article can be found, in the substituted LaNiO3 mixed oxides as catalyst precursors for glycerol steam
online version, at doi:https://doi.org/10.1016/j.apcatb.2021.119884. reforming, Appl. Catal. B: Environ. 147 (2014) 193–202.

12
Z. Xiao et al. Applied Catalysis B: Environmental 286 (2021) 119884

[28] D. Harshini, D.H. Lee, J. Jeong, Y. Kim, S.W. Nam, H.C. Ham, J.H. Han, T.-H. Lim, [40] B. Wang, Y. Xiong, Y. Han, J. Hong, Y. Zhang, J. Li, F. Jing, W. Chu, Preparation of
C.W. Yoon, Enhanced oxygen storage capacity of Ce0.65Hf0.25M0.1O2-δ (M=rare stable and highly active Ni/CeO2 catalysts by glow discharge plasma technique for
earth elements): applications to methane steam reforming with high coking glycerol steam reforming, Appl. Catal. B: Environ. 249 (2019) 257–265.
resistance, Appl. Catal. B: Environ. 148-149 (2014) 415–423. [41] K. Li, X. Chang, C. Pei, X. Li, S. Chen, X. Zhang, S. Assabumrungrat, Z.-J. Zhao,
[29] Z. Xiao, S. Ji, F. Hou, Y. Li, H. Zhang, L. Wang, X. Zhang, G. Liu, J. Zou, G. Li, n- L. Zeng, J. Gong, Ordered mesoporous Ni/La2O3 catalysts with interfacial
Dodecane steam reforming catalyzed by Ni-Ce-Pr catalysts. Part 1: Catalyst synergism towards CO2 activation in dry reforming of methane, Appl. Catal. B:
preparation and Pr doping, Catal. Today 316 (2018) 78–90. Environ. 259 (2019), 118092.
[30] J. Carrasco, L. Barrio, P. Liu, J.A. Rodriguez, M.V. Ganduglia-Pirovano, Theoretical [42] S. Chen, C. Pei, J. Gong, Insights into interface engineering in steam reforming
studies of the adsorption of CO and C on Ni(111) and Ni/CeO2(111): evidence of a reactions for hydrogen production, Energy Environ. Sci. 12 (2019) 3473.
strong metal-support interaction, J. Phys. Chem. C 117 (2013) 8241–8250. [43] G. Niu, E. Hildebrandt, M.A. Schubert, F. Boscherini, M.H. Zoellner, L. Alff,
[31] S. Fabris, S.D. Gironcoli, S. Baroni, G. Vicario, G. Balducci, Taming multiple D. Walczyk, P. Zaumseil, I. Costina, H. Wilkens, Oxygen vacancy induced room
valency with density functionals: a case study of defective ceria, Phys. Rev. B 71 temperature ferromagnetism in Pr-Doped CeO2 thin films on silicon, ACS Appl.
(041101-041104) (2005) 041102. Mater. Interfaces 6 (2014) 17496–17505.
[32] R.G.-C, Norge C. Herná ndez, Nora H. de Leeuw, Javier Fdez. Sanz, Electronic [44] S.T. Igor Luisetto, Claudia Romano, Marta Boaro, Elisabetta Di Bartolomeo, S.S.
charge transfer between ceria surfaces and gold adatoms: a GGA+U investigation, K. Jagadesh Kopula Kesavan, Karuppiah Selvakumar, Dry reforming of methane
Phys. Chem. Chem. Phys. 11 (2009) 5246–5252. over Ni supported on doped CeO2: new insight on the role of dopants for CO2
[33] Y. Chen, J. Cheng, P. Hu, H. Wang, Examining the redox and formate mechanisms activation, J. CO2 Util. 30 (2019) 63–78.
for water-gas shift reaction on Au/CeO2 using density functional theory, Surf. Sci. [45] L. Pino, C. Italiano, A. Vita, M. Laganà, V. Recupero, Ce0.70La0.20Ni0.10O2-δ catalyst
602 (2008) 2828–2834. for methane dry reforming: Influence of reduction temperature on the catalytic
[34] M.Sa.T. Ziegler, The electronic structure and chemical properties of a Ni/CeO2 activity and stability, Appl. Catal. B: Environ. 218 (2017) 779–792.
anode in a solid oxide fuel cell: a DFT + U study, J. Phys. Chem. C 114 (2010) [46] A.T.M.L. Trudeau, J.Y. Ying, XPS investigation of surface oxidation and reduction
21411–21416. in nanocrystalline CexLa1-xO2-y, Surf. Interface Anal. 23 (1995) 219–226.
[35] Bo Li, Horia Metiu, DFT studies of oxygen vacancies on undoped and doped La2O3 [47] Delfina García Pintos, Alfredo Juan, Beatriz Irigoyen, Oxygen vacancy formation
surfaces, J. Phys. Chem. C 114 (2010) 12234–12244. on the Ni/Ce0.75Zr0.25O2(111) surface. A DFT+U study, Int. J. Hydrogen Energy 37
[36] D.L.D. Javier Carrasco, Zongyuan Liu, Tomas Duchon, Jaime Evans, Sanjaya (2012) 14937–14944.
D. Senanayake, Ethan J. Crumlin, Vladimir Matolim, Jose A. Rodriguez, [48] Z. Yang, T.K. Woo, K. Hermansson, Effects of Zr doping on stoichiometric and
M. Veronica Ganduglia-Pirovano, In situ and theoretical studies for the dissociation reduced ceria: a first-principles study, J. Chem. Phys. 124 (2006), 224704.
of water on an active Ni/CeO2 catalyst:importance of strong metal-support [49] A.H. Eranda Nikolla, Johannes Schwank, Suljo Linic, Controlling carbon surface
interactions for the cleavage of O–H bonds, Angew. Chem. 127 (2015) 3989–3993. chemistry by alloying: carbon tolerant reforming catalyst, J. Am. Chem. Soc. 128
[37] Y. Zhang, L. Zhang, L. Pan, X. Zhang, L. Wang, Fazal-e-Aleem, J. Zou, Role of (2006) 11354–11355.
oxygen vacancies in photocatalytic water oxidation on ceria oxide: experiment and [50] Feng Zhang, Zongyuan Liu, Xiaobo Chen, Ning Rui, Luis E. Betancourt, Lili Lin,
DFT studies, Appl. Catal. B: Environ. 224 (2018) 101–108. Wenqian Xu, Cheng-Jun Sun, A.M. Milinda Abeykoon, Jose A. Rodriguez, J.
[38] Z. Xiao, X. Zhang, F. Hou, C. Wu, L. Wang, G. Li, Tuning metal-support interaction T. Kristijan Lorber, Petar Djinovic, Sanjaya D. Senanayake, The effects of Zr-doping
and oxygen vacancies of ceria supported nickel catalysts by Tb doping for n- into ceria for the dry reforming of methane over Ni/CeZrO2 catalysts: in-situ
dodecane steam reforming, Appl. Surf. Sci. 503 (2020), 144319. studies with XRD, XAFS and AP-XPS, ACS Catal. 10 (2020) 3274–3284.
[39] L. Pino, A. Vita, F. Cipitì, M. Laganà, V. Recupero, Catalytic performance of [51] X. Yan, T. Hu, P. Liu, S. Li, B. Zhao, Q. Zhang, W. Jiao, S. Chen, P. Wang, J. Lu,
Ce1− xNixO2 catalysts for propane oxidative steam reforming, Catal. Lett. 122 L. Fan, X. Deng, Y.-X. Pan, Highly efficient and stable Ni/CeO2-SiO2 catalyst for dry
(2007) 121–130. reforming of methane: effect of interfacial structure of Ni/CeO2 on SiO2, Appl.
Catal. B: Environ. 246 (2019) 221–231.

13

You might also like