You are on page 1of 10

Applied Catalysis B: Environmental 286 (2021) 119896

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

Catalytic ammonia decomposition over Ni-Ru supported on CeO2 for


hydrogen production: Effect of metal loading and kinetic analysis
Ilaria Lucentini a, Germán García Colli b, c, Carlos D. Luzi b, c, Isabel Serrano a,
Osvaldo M. Martínez b, c, **, Jordi Llorca a, *
a
Institute of Energy Technologies, Department of Chemical Engineering and Barcelona Research Center in Multiscale Science and Engineering, Universitat Politècnica de
Catalunya, EEBE, Eduard Maristany 10-14, 08019 Barcelona, Spain
b
Department of Chemical Engineering, Faculty of Engineering, Universidad Nacional de La Plata, Calle 115 e/48 y 49, La Plata, Argentina
c
Development and Investigation Centre in Applied Science ‘Dr. J. J. Ronco’ (CINDECA), CONICET - Universidad Nacional de La Plata, Calle 47 n◦ 257, La Plata,
Argentina

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

Keywords: Ceria-supported Ni-Ru bimetallic catalysts with different metal loadings have been prepared by co-impregnation,
Ammonia decomposition characterized and tested in the production of hydrogen from the catalytic decomposition of ammonia. The
Bimetallic catalysts bimetallic catalysts showed an excellent catalytic performance in long-term stability tests with respect to
Nickel-ruthenium
monometallic Ru/CeO2 and Ni/CeO2 and in multicycle tests under pure ammonia. The best catalytic perfor­
Ceria catalysts
Kinetics
mance has been obtained over catalysts with 2.4− 5 wt.% Ni, 0.4− 0.6 wt.% Ru, and a Ni/Ru wt.% ratio of ca. 7.
TOFH2 values exceeding 2 s− 1 have been obtained, which are among the highest reported for ammonia
decomposition at 400 ◦ C. Raman spectroscopy, XRD, HRTEM, XPS, TPR and H2 chemisorption have revealed the
existence of an intimate contact between Ni and Ru and CeO2, which is considered the reason of the excellent
catalytic activity and stability observed. A kinetic model has been developed using the Langmuir-Hinshelwood-
Hougen-Watson approach for the decomposition of ammonia in a fixed bed reactor. The reaction rate expression
of the ammonia decomposition on Ni-Ru bimetallics supported on ceria suggests that the dehydrogenation of the
ammonia adsorbed on the surface of the catalyst is the limiting step of the reaction and that ammonia decom­
position is inhibited by the presence of H2.

been studied, such as Ni, Ir, Mo, Co, Pt, Pd or Rh [4], and their combi­
nations Co-Mo [5], Fe-Co [6], Fe-Mo [7], Ni-Fe [8] and Ni-Ru [9]. The
1. Introduction
catalytic activity in the decomposition of ammonia does not depend only
on the metal phase, but also on the type of support and electron donor
The on-site catalytic production of hydrogen from ammonia is
promoters [10], such as K [11], Cs [12] or Ba [13]. Among the catalysts
considered as an interesting approach to feed Proton Exchange Mem­
reported so far, currently those based on ruthenium have shown the
brane Fuel Cells (PEMFC), which require a high purity of hydrogen [1].
highest catalytic activity for the ammonia decomposition reaction [14].
Among numerous advantages, the hydrogen content of ammonia is high
However, they experience deactivation by sintering (i.e. Ru/Al2O3) or
(17.8 wt.% and 108 kg H2 m− 3 at 20 ◦ C and 8.6 bar) and can be trans­
by volatilization (i.e. Ru/CeO2) [9]. In a previous work, we have
ported easily as liquid at low pressure [2]. In addition, it can be
recently demonstrated the benefits of the bimetallic Ni-Ru system sup­
decomposed into hydrogen and nitrogen at moderate temperature, NH3
− 1 ported on ceria for the decomposition of ammonia [9]. Here we extend
(g) ⇆ 1/2 N2(g) + 3/2 H2(g), ΔH = 46.22 kJ mol . This reaction is the

these results by carefully studying the effect of the metal loading and the
inverse of the Haber-Bosh process for the synthesis of ammonia; for this
relative Ni:Ru atomic ratio of Ni-Ru/CeO2 on catalytic activity as well as
reason and taking into account the microreversiblity principle in het­
their effect on long-time stability and multicycle tests. In addition, a
erogeneous catalysts, catalysts used for the synthesis of ammonia have
mathematical model based on the analysis of the kinetics is developed
been widely studied for the decomposition of ammonia, such as those
over the catalysts showing the best catalytic performance.
based on Ru and Fe [3]. In more recent times other types of metals have

* Corresponding author at: Universitat Politècnica de Catalunya, Eduard Maristany 10-14, 08019 Barcelona, Spain.
** Corresponding author at: Universidad Nacional de La Plata, Calle 47 n◦ 257, 1900AJK, La Plata, Argentina.
E-mail addresses: ommartin@ing.unlp.edu.ar (O.M. Martínez), jordi.llorca@upc.edu (J. Llorca).

https://doi.org/10.1016/j.apcatb.2021.119896
Received 2 November 2020; Received in revised form 22 December 2020; Accepted 28 December 2020
Available online 11 January 2021
0926-3373/© 2021 Elsevier B.V. All rights reserved.
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

Nomenclature TOF Turnover frequency [s− 1]


V Volume [m3]
Symbol VAR Variance [-]
C Molar concentration [mol m− 3] wt. Weight [%]
d Diameter [m] x Conversion [%]
Ek Activation energy [kJ mol− 1] ε Porosity [-]
F Molar flow [mol s− 1] ρ Density [kg m− 3]
F/W Flow to weight ratio [L g− 1 h− 1]
fv Catalyst volumetric fraction [-] Subscript
GHSV Gas hourly space velocity [h− 1] (a) Adsorbed
ΔHº Standard enthalpy of reaction [kJ mol− 1] (g) Gas
k Specific constant of reaction rate [mol g− 1 s− 1] ap Apparent
K Constant [m3 mol− 1] av Average
k0 Frequency or pre-exponential factor [mol g− 1 s− 1] cat Catalyst
MSC Model Selection Criterion [-] H2 Hydrogen
N Total number [-] i Inner
n Value [-] in Inlet
p Number of parameters [-] L Catalytic bed
R Gas constant [kJ mol− 1 K− 1] n n value
r Reaction rate [mol g− 1 s− 1] NH3 Ammonia
s Vacant site [-] out Outlet
SSR Sum of the square of the residuals [-] P Inert particle
SST Sum of total square [-] pred Predicted
T Temperature [K]

In the past few years a large number of studies have been done on the the incipient wetness impregnation (IWI) method by co-impregnation
kinetics of ammonia decomposition using different catalyst composi­ with an aqueous solution (0.3 M total metal) of Ni(CH3COO)2⋅4H2O
tions to determine the expression of the reaction rate and to understand (Probus) and RuCl3 (Tokyo Chemical Industries). The samples were
the mechanism involved. Two rate-determining steps are usually pro­ calcined at 550 ◦ C (4 h, 5 ◦ C min− 1). The amount of nickel and ruthe­
posed for ammonia decomposition: (i) the recombinative desorption of nium in these catalysts were varied between 1.5− 10 wt.% Ni and 0.3− 2
nitrogen atoms [15], or (ii) the scission of the N–H bond [14,16], wt.% Ru. They are labelled as 10Ni2Ru, 5Ni1Ru, 2.5Ni0.5Ru,
although some studies claim that the limiting step of the reaction 2.5Ni0.3Ru, 2.0Ni0.4Ru and 1.5Ni0.5Ru, where the number preceding
changes according to the reaction conditions, in particular temperature the metal indicates the respective nominal wt.%. For comparative pur­
[17]. The first kinetic expressions considered that the decomposition of poses, two monometallic samples, 2 wt.% Ru/CeO2 and 5 wt.% Ni/
ammonia was a first-order reaction [18], or a combination of zero-order CeO2, were also prepared following the same synthesis method.
and first-order depending on temperature [19]. Recent studies have
shown that the reaction is inhibited by the presence of H2, i.e., therefore 2.2. Catalyst characterization
the reaction rate depends on both, the ammonia and hydrogen con­
centration [20]. Different methods have been used to determine the Raman spectroscopy was performed with a Renishaw inVia Qontor
expression of the reaction rate, such as the confocal Raman microscope (532.1 nm laser) and a Leica DM2700 M
Langmuir-Hinshelwood-Hougen-Watson (LHHW) approach, which microscope (x50). An exposure time of 1 s and 12 repetitions were used
considers a single-binding state with coverage-independent parameters to acquire the spectra. X-ray diffraction (XRD) was performed with a
[15], the Temkin-Pyzhev model, which considers that the associative Siemens D5000 diffractometer (Cu Kα, 45 kV, 35 mA). The crystallite
desorption of nitrogen is the rate-limiting step and that if the influence size was calculated with the Debye-Scherrer. X-ray photoelectron spec­
of the reverse reaction can be ignored the reaction rate is given as a troscopy (XPS) was performed with a SPECS XR50 source (250 W) and a
power-law expression [21], or the Tamaru model, which considers that Phoibos 150 MCD-9 detector. Binding energy values were corrected for
according to the reaction conditions different models apply [16,22]. In the position of the C 1s peak at 284.8 eV. Data processing was performed
this work, the LHHW approach has been employed and the kinetic pa­ with the CasaXPS program (Casa Software Ltd., UK). The spectra of Ce 3d
rameters have been obtained by testing different kinetic expressions spectra was analyzed using three pairs of spin-orbit doublets for Ce4+
obtained by alternatively considering each elementary step of the re­ and two pairs of spin-orbit doublets for Ce3+ [24]. The spectra of Ni 2p
action as the rate-limiting to find the one that best fit the experimental was analyzed using two pairs of spin-orbit doublets with their corre­
results, considering all the other steps of the reaction in equilibrium sponding satellite lines for Ni2+ and one spin-orbit doublet with the
(Quasi-Equilibrium Approximation) [23]. Finally, the kinetic expression corresponding satellite lines for Ni3+ [25]. The spectra of Ru 3p was
has been used to determine the best operational conditions for the analyzed using two pairs of spin-orbit doublets for Ru4+ and one
ammonia decomposition on Ni-Ru/CeO2. spin-orbit doublet for Ru◦ [26]. High resolution transmission electron
microscopy (HRTEM) was performed with a FEI TECNAI F20 micro­
2. Materials and methods scope equipped with a field emission electron gun (200 kV). Tempera­
ture programmed reduction (TPR) was performed with a Chemstar
2.1. Preparation of catalysts apparatus with TCD detector. Samples were first treated at 450 ◦ C in Ar
(50 mL min− 1, 10 ◦ C min− 1), and then TPR was carried out from 30 to
CeO2 was prepared by adding NH4OH to a magnetically stirred so­ 850 ◦ C (10 ◦ C min− 1) under 10 % H2 and kept at this temperature for 10
lution of Ce(NO3)3⋅6H2O in water, until pH~9. The precipitate was min. H2 chemisorption experiments were performed with a BELCAT
calcined at 500 ◦ C (4 h, 5 ◦ C min− 1). The catalysts were prepared using instrument equipped with TCD. Samples were treated at 450 ◦ C in 10 %

2
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

H2 (50 mL min− 1), cooled to 20 ◦ C under Ar, and subjected to pulses of calculated as the catalyst volume (0.606⋅10-7 m3) divided by the solid
10 % H2 in Ar until no further H2 was consumed. The metal dispersion phase total volume (3.095⋅10-7 m3), VL is the total volume of the cata­
was calculated assuming an adsorption stoichiometry factor of 2. The lytic bed (4.095⋅10-7 m3), x is the ammonia conversion and r is the re­
loading of Ni and Ru loading was measured by inductively coupled action rate (mol NH3 g-1 -1
cat s ). εL is the bed porosity (0.2451), calculated
plasma-mass spectrometry (ICP-MS) using an Agilent Technologies 7800 through the model proposed by Dias et al. (2004) for beds composed of a
ICP-MS instrument. Porosity and apparent and bulk density were mixture of spherical particles of two materials with different diameters
measured by mercury intrusion porosimetry (MIP) with an AutoPore IV [27] (Eq. 5):
9500 Micromeritics apparatus. ⎛ ⎞
⎜ / ⎟
⎜ √̅̅̅̅̅ ⎟
2.3. Catalytic tests
1
εav fv ⎜ dcat

exp⎜ ⎟ (5)
dP
εL = ⎜ 1.2264(1 − fv ) ⎟
1 − εav (1 − fv ) ⎜ ⎟
The decomposition of ammonia was conducted at atmospheric ⎝ ⎠
pressure in a stainless-steel reactor. The mass of the catalyst used was
0.1 g, which was mixed with SiC to obtain a constant bed volume of ca.
0.41 cm3. Temperature-dependent reaction tests were carried out in the where εav (0.3855) is the average porosity of the catalytic bed if it was
range 350− 600 ◦ C at steps of 50 ◦ C (30 min at each temperature) using a composed only of the catalyst particles with average diameter dcat
Ar:NH3 ratio of 1.2:1 on a molar basis and a total gas flow of 25 mL (0.00010 m), calculated through the model proposed by Mariani et al.
min− 1. The gases were provided by Bronkhorst mass flow controllers. (2002) [28] reported in Eq. 6, and dP is the SiC average particle diameter
Prior to the catalytic tests, samples were first activated at 300 ◦ C for 1 h (0.00037 m).
with 10 % H2. The reaction was monitored by mass spectrometry with a dcat
OmniStar instrument (measurement error of ±5 %). Stability tests were εav = 0.375 + 0.48 (6)
di
carried out at 450 ◦ C for at least 100 h using a Ar:NH3 = 1.2:1 mixture on
a molar basis. The multicycle test was carried out with a pure ammonia where di is the inner reactor diameter (0.004572 m).
flow of 50 mL min− 1 repeating the same temperature cycle (350− 550 The kinetic model was fitted to the experimental data of ammonia

C, steps of 50 ◦ C) ten times. Catalytic tests under different flow-to- conversion through a parameter estimation problem using MATLAB
weight (F/W) values were carried out from 350 to 550 ◦ C at steps of software by minimizing the objective function for the sum of the square
50 ◦ C using 25, 50 and 75 mL min-1 total gas flow values. Catalytic tests of the residuals, SSR (Eq. 7). To this end, an unweighted nonlinear
under different ammonia concentration in the inlet flow (50, 60 and 70 regression was performed using the MATLAB function nlinfit.
%) were carried out from 350 to 550 ◦ C at steps of 50 ◦ C maintaining a
constant F/W value of 30.0 L g−cat1 h− 1 (50 mL min-1 total flow). The

N
( )2
SSR = xn − xpred,n (7)
evaluation of H2 and N2 inhibition was done between 350 to 550 ◦ C at n=1
steps of 50 ◦ C maintaining a constant flow of NH3 (5 mL min− 1) and
changing the concentration of H2 (0, 22.5, 45.0 or 67.5 %) or N2 (0, 22.5 where xpred,n is the conversion estimated by the model, xn corresponds to
or 67.5 %) (50 mL min-1 total flow, Ar balance). Ammonia conversion x the n value of the conversion measured experimentally according to Eq.
has been obtained from Eq. 1, where FinNH3 and FoutNH3 are the inlet and 1 and N is the total number of experiments. The kinetic analysis was
outlet molar flow of ammonia, respectively (mol NH3 s− 1). completed with an estimation of the confidence intervals of the reaction
rate parameters, with the complementary function nlpredci.
FinNH3 − FoutNH3
x= (1)
FinNH3 3. Results and discussion

2.4. Kinetic analysis 3.1. Catalytic tests

A continuous-type model was used to determine the kinetics of the Table 1 compiles the ammonia conversion values at 400, 450 and
ammonia decomposition reaction over the powder catalysts in a fixed- 500 ◦ C attained by the Ni-Ru/CeO2 catalysts at F/W ratio of 15.0 L g−cat1
bed reactor. The simplest model is the pseudo-homogeneous one- h− 1. The Ni and Ru metal loadings obtained by ICP-MS and the metal
dimensional model without axial dispersion, where interfacial gradients dispersion obtained by H2 chemisorption are also listed in Table 1,
are neglected. The flow was considered in a plug flow regime. Prior to which are used to calculate the catalytic turnover frequency values at
the kinetic analysis, an estimation of the catalytic bed outlet tempera­ 400 ◦ C. The only products of the reaction were, in all cases, hydrogen
ture was carried out, demonstrating that the heat transfer properties of and nitrogen in a molar ratio of H2:N2 = 3:1, as expected from the
the fixed-bed reactor used were suitable to consider that the system was stoichiometry of the reaction. Blank tests carried out with the empty
isothermal under all the reaction conditions. On the other hand, it was reactor, with SiC and with bare CeO2 yielded much lower ammonia
verified that heat and mass resistances inside the particle were negli­ conversion values and similar between them, pointing out to a neglibible
gible. The reaction equilibrium constant was calculated in order to work ammonia conversion activity of the support (NH3 conversion values of 4,
out the ammonia concentration at equilibrium for each temperature 9 and 24 % at 400, 450 and 500 ◦ C, respectively).
tested. At steady state, the mass balance can be expressed through Eqs. The difference in catalytic activity exhibited by the different samples
2–4: allows to discuss about the effect of the metal loading, the Ni:Ru metal
ratio and the dispersion of the metals on CeO2. On one hand, there is a
dFNH3
− rρap (1 − εL )fv = (2) positive trend between the total metal dispersion (Ni + Ru) and the TOF
dV
values; on the other hand, the metal dispersion strongly decreases with
dFNH3 = − FinNH3 dx (3) the total metal loading following a power trend (Fig. 1a). These two
observations reveal that the metal dispersion is strongly affected by the
∫ ∫
VL x
FinNH3 dx metal loading, which, in turn, has a direct influence on catalytic activity.
ρap (1 − εL )fv dV = (4) This is likely a direct consequence of the metal particle size, i.e. the
0 0 r
number of metal atoms exposed at the surface. Another interesting trend
where ρap is the apparent density of the catalyst determined with MIP is that exhibited by the metal ratio and the catalytic activity. Fig. 1b
technique (1651.8 kg m− 3), fv is the catalyst volumetric fraction

3
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

Table 1
Properties of the Ni-Ru/CeO2 catalysts and catalytic results. GHSV = 3660 h− 1, 0.1 gcat, Ar:NH3 = 1.2:1 and F/W = 15.0 L g-1 -1
cat h .

Catalyst (supported on CeO2) 10Ni2Ru 5Ni1Ru 2.5Ni0.5Ru 2.0Ni0.4Ru 1.5Ni0.5Ru 2.5Ni0.3Ru


a
Ni wt.% 12.73 4.98 2.39 1.40 0.85 1.77
Ru wt.%a 1.22 0.64 0.35 0.25 0.29 0.17
Dispersion (%)b 2.0 4.6 6.6 10.2 11.5 7.7
x(%)NH3 400 ◦ C 35.1 36.5 49.7 36.2 41.9 23.2
x(%)NH3 450 ◦ C 84.5 86.1 88.7 81.2 86.3 73.4
x(%)NH3 500 ◦ C 99.2 99.6 99.2 98.4 97.8 98.1
TOFH2 (s− 1)c 0.88 1.01 2.00 1.57 2.44 1.11
a
Metal loading determined by ICP-MS.
b
Metal dispersion determined by H2 chemisorption.
c
TOFH2 calculated at 400 ◦ C.

Fig. 1. a) Correlation between the total metal content, TOFH2 values and metal dispersion. b) Correlation between TOFH2 values and the Ru/(Ru + Ni) (wt. %). GHSV
= 3660 h− 1, 0.1 gcat, Ar:NH3 = 1.2:1, F/W = 15.0 L g-1 -1
cat h .

shows the relationship between the Ru/(Ru + Ni) ratio (wt. %) and the Ni-based catalysts are usually lower than 1 s− 1 [8,9], and those reported
TOF values. It is observed that the higher the Ru/(Ru + Ni) content, the for pure Ru supported on different types of oxides and carbon nanotubes
higher the catalytic activity. Nevertheless, the TOF values do not follow found in literature are rarely superior than 1 s− 1 [11] and, in a few cases,
a trend with the Ru content of the catalyst, since an increase of the metal between 1 and 4 s− 1 [9,29].
content lead to a poorer dispersion. It can be concluded that the catalyst Stability tests were carried out over 10Ni2Ru/CeO2, 5Ni1Ru/CeO2
2.5Ni0.5Ru/CeO2 presents the best trade-off between Ru loading, Ni/Ru and 2.5Ni0.5Ru/CeO2 at 450 ◦ C (Fig. 2a). The results obtained on 10Ni/
metal ratio and dispersion. By comparing the TOFH2 values obtained at CeO2 and 2Ru/CeO2 under exactly the same reaction conditions were
400 ◦ C with those presented in the literature under similar reaction reported in a previous study [9] and have been included in Fig. 2a for
conditions, it is concluded that the values of the Ni-Ru/CeO2 bimetallic comparative purposes (dashed lines). In contrast to the monometallic
catalysts prepared in this study are among the best TOFH2 values for counterparts, the good stability exhibited by the bimetallic samples after
ammonia decomposition reported so far. The TOFH2 values obtained on 100 h of continuous operation merits to be highlighted. In particular, the

Fig. 2. a) Stability tests in the decomposition of ammonia at 450 ◦ C under GHSV = 3660 h− 1, F/W = 15.0 L g−cat1 h− 1
and Ar:NH3 = 1.2:1. b) Multicycle tests on
5Ni1Ru/CeO2 and 2.5Ni0.5Ru/CeO2 using 100 % NH3 under GHSV = 7320 h− 1 and F/W = 30.0 L g−cat1 h− 1.

4
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

5Ni1Ru/CeO2 catalyst was the most stable, presenting a constant


ammonia conversion, whereas the 10Ni2Ru/CeO2 catalyst showed a
progressive but almost insignificant deactivation. The 2.5Ni0.5Ru/CeO2
catalyst presented a higher initial conversion, but it deactivated more
rapidly (2 % over 100 h). For a further assessment of the robustness of
the bimetallic samples under realistic operation conditions, the
2.5Ni0.5Ru/CeO2 and 5Ni1Ru/CeO2 catalysts were tested in 10
consecutive temperature cycles from 350 to 550 ◦ C (duration of each
cycle 2.5 h) under a continuous flow of 50 mL min− 1 of pure ammonia
(Fig. 2b). During the first cycles, there is a slight improvement of the
catalytic activity at low temperatures (350− 400 ◦ C) due to the reduction
of the metal phase during reaction at high temperatures. Metallic Ni and
Ru are known to be the active phases for the reaction [9]. From the third
cycle onwards the catalysts show a perfectly constant catalytic activity,
which is a clear indication of the robustness of the Ni-Ru/CeO2 bime­
tallic catalysts taking into account that the feed is composed by pure
ammonia. The 2.5Ni0.5Ru/CeO2 catalyst, despite having a higher initial
conversion, it follows the same behaviour as 5Ni1Ru/CeO2 starting from
the second cycle.
An analysis of the effects of hydrogen and nitrogen (the products of
the reaction) on the conversion of ammonia was carried out with the
5Ni1Ru/CeO2 catalyst by introducing different concentrations of H2 and Fig. 4. Raman spectra of A) 10Ni2Ru/CeO2, B) 5Ni1Ru/CeO2, C) 2.5Ni0.5Ru/
N2 in the inlet gas flow (Fig. 3). According to the results, the addition of CeO2, D) 1.5Ni0.5Ru/CeO2, E) 2.0Ni0.4Ru/CeO2, F) 2.5Ni0.3Ru/CeO2, G) 5Ni/
hydrogen to the inlet gas flow led to an important decrease in the con­ CeO2, H) 2Ru/CeO2, I) CeO2, J) RuO2, K) NiO.
version of ammonia (Fig. 3a). On the contrary, after the addition of
nitrogen the ammonia conversion remained constant (Fig. 3b). This cm− 1, and a weak band at 1050 cm− 1, which were ascribed to TO and
shows not only that N2 is kinetically inert, but also that there is no LO-one phonon and LO-two phonon scattering modes, respectively [33].
equilibrium limitation on the studied conditions and, consequently, the Except for the reference samples NiO and RuO2, the intensity of the
behaviour shown by the hydrogen is a truly inhibition effect. The H2 spectra was normalized to that of the F2g band of ceria. The F2g bands of
inhibition effect has been found in various works reported in the liter­ all catalysts appeared red-shifted and broader with respect to that of
ature [30]. bare CeO2, which is attributed to a distortion of the ceria lattice due to
the formation of Ni–OCe and RuOC–––e bonds [34], which in turn is
3.2. Characterization an indication of the existence of a strong interaction between ceria and
nickel oxide and ruthenium oxide. The possibility of solid solution be­
Fig. 4 shows the Raman spectra recorded for all catalysts as well as tween ceria and the metal oxides can be ruled out because there was no
those of monometallic Ni/CeO2 and Ru/CeO2. The reference samples change in the lattice parameter of ceria in the catalysts as measured by
CeO2, NiO and RuO2 are also included for comparison. The spectrum of XRD (Fig. 5). Interestingly, the presence of several bands in the range
CeO2 showed the main peak at 462 cm− 1 related to the symmetrical 520− 700 cm− 1 in the Raman spectra of the catalysts (also in the range
stretching mode of the ceria fluorite phase (F2g mode) and three weaker 250− 360 cm− 1 for Ru/CeO2) is a strong indication of the presence of
bands at 258, 600 and 1167 cm− 1, corresponding to second order defects. Different bands related to extrinsic oxygen vacancies and
transverse acoustic (2TA), Frenkel defect-induced (D), and second order intrinsic Frenkel-type vacancies in ceria have been identified when a
longitude optical (2LO) modes, respectively [31]. The characteristic dopant is added [35]. Again, this is an indication that an intimate con­
bands of RuO2 were identified at 500, 615 and 684 cm− 1 and they were tact exists between the ceria support and the metals [36]. Regarding the
attributed to the Eg, A1g and B2g modes, respectively [32]. The spectrum other bands detected in the spectra of the Ru/CeO2 and of the bimetallic
of NiO presented a broad band at 506 cm− 1, with a shoulder at 425 samples at about 700 and 980 cm− 1, they were assigned to the

Fig. 3. Ammonia decomposition catalytic tests carried out on 5Ni1Ru/CeO2 at different inlet hydrogen (a) and nitrogen (b) concentrations at GHSV = 7320 h− 1, F/W
= 30.0 L g−cat1 h− 1 and 10 % of NH3 balanced with Ar.

5
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

profiles recorded over 10Ni2Ru/CeO2 and 5Ni1Ru/CeO2 (Fig. 5). In


addition to the ceria peaks at ca. 29, 33, 48, 56, 60 and 70◦ [39], peaks of
NiO were identified at ca. 37, 43 and 63◦ [40], and in the sample with
the higher metal content, 10Ni2Ru/CeO2, a small peak at ca. 35◦ cor­
responding to RuO2 was tentatively identified [41]. By using the
Debye-Scherrer formula, the mean crystallite size for CeO2 was esti­
mated to be 13− 15 nm. This is in agreement with TEM results (Fig. 6a).
The HRTEM analysis of the 5Ni1Ru/CeO2 catalyst (Fig. 6) allowed
the study of the microstructure of the bimetallic catalysts. The CeO2
crystallites are polyhedral and the common facets of ceria at 3.12 and
2.71 Å corresponding to the (111) and (200) crystallographic planes,
respectively, are clearly identified. In Fig. 6b, nickel and ruthenium
were identified by the presence of spots at 2.09 and 2.21 Å in the Fourier
Transform (FT) image, which correspond to the (200) crystallographic
planes of NiO and RuO2, respectively. The average particle size of NiO is
ca. 5− 6 nm and that of RuO2 is about 3− 4 nm. It is important to note
that EDX spectra (a representative spectrum is shown in Fig. 6c) recor­
ded over different parts of the sample revealed, in all cases, the simul­
taneous presence of Ni and Ru peaks, which is an additional indication
of a close contact between the two metals in the bimetallic catalysts.
The intimate interaction between the ceria support and the metals
was also corroborated by TPR measurements (Fig. 7). The profile of the
Fig. 5. XRD patterns of 10Ni2Ru/CeO2, 5Ni1Ru/CeO2 and CeO2.
CeO2 support (spectrum I) presented two peaks; the one at around 750

C corresponds to the reduction of bulk Ce(IV) to Ce(III) and that close to
Ru–OC–e bond [37]. A particular feature appearing in the Raman
500 ◦ C is ascribed to the surface reduction of ceria involving bridging
spectra of the bimetallic catalysts was the existence of a band at ca. 230
OH group formation [42]. These two peaks were detected in all the other
cm− 1. This band cannot be attributed to any nickel or ruthenium oxide
samples analysed; the one corresponding to the reduction of bulk ceria
species [37] and we ascribe them to the vibrations of Ce in a distorted
was located at the same temperature than bare CeO2, whereas the one
oxygen lattice environment [38].
related to the reduction of surface ceria was shifted to lower tempera­
As pointed out above, no differences in the lattice parameter of CeO2
tures (around 400 ◦ C). This is a well-known phenomenon that has been
upon incorporation of nickel and ruthenium was measured in the XRD

Fig. 6. HRTEM images and EDX of 5Ni1Ru/CeO2.

6
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

Fig. 7. a) TPR profiles of A) 10Ni2Ru/CeO2, B) 5Ni1Ru/CeO2, C) 2.5Ni0.5Ru/ Fig. 8. XP spectra of Ce 3d and Ni 2p (a) and Ru 3p (b) of 5Ni1Ru/CeO2.
CeO2, D) 1.5Ni0.5Ru/CeO2, E) 2.0Ni0.4Ru/CeO2, F) 2.5Ni0.3Ru/CeO2, G) 5Ni/
CeO2, H) 2Ru/CeO2, I) CeO2.
(74.6 %), respectively, as well as a satellite line at higher binding energy.
The presence of metallic ruthenium in the bimetallic catalyst can be
ascribed to the metal capacity of activate hydrogen through the explained by the decomposition of RuO2 to Ru◦ upon calcination, as
hydrogen spillover effect [43]. It is known that both Ru and Ni are reported in the literature [47]. By comparing these results with those
capable of absorbing hydrogen on their surface, then the atomic obtained in the XPS analysis of the monometallic catalysts in our pre­
hydrogen that is generated can be transported to the support [44]. These vious work [9], it is noticeable that, in the bimetallic catalyst, Ni and Ru
results suggest that Ru has a greater ability to decrease the reduction are present in a much more reduced state, considering that Ni2+ and
temperature of surface ceria than Ni, since the corresponding peak is Ni3+ in the monometallic catalyst were 75.8 and 24.2 %, respectively of
shifted at lower temperature. The Ni/CeO2 sample (spectrum G) pre­ the total Ni, and Ru4+ and Ru6+ were found, but not metallic Ru. This
sented a large peak at 280 ◦ C with a shoulder at 217 ◦ C, which over­ again can be interpreted as a result of a strong interaction between Ni
lapped the peak of the surface ceria reduction. These two peaks were and Ru in the bimetallic catalyst.
assigned to the reduction of NiO species present on the surface. A smaller
peak at around 587 ◦ C was ascribed to the reduction of strong Ni–OC–e
entities [45]. Regarding the Ru/CeO2 spectrum (H), the peak related to 3.3. Kinetic analysis
ruthenium reduction appeared at 83 ◦ C, attributed to the reduction of
RuO2 [37]. In addition to the peaks related to ceria, the 10Ni2Ru/CeO2 A kinetic analysis was carried out with the experimental results ob­
catalyst (spectrum A) presented various peaks at low temperatures tained with catalysts 5Ni1Ru/CeO2 and 2.5Ni0.5Ru/CeO2. The deter­
attributed to the reduction of Ru (signal at 84 ◦ C) and Ni (signals at 250 mination of the reaction rate was accomplished by testing different
and 300 ◦ C). The presence of other peaks at 100 and 141 ◦ C suggests a kinetic expressions obtained by applying the LHHW method and by
strong interaction between Ni and Ru, in accordance to the presence of considering the following steps (Eqs. 8–13):
bimetallic Ni-Ru particles. The other bimetallic catalysts (spectra B–F)
presented a single or slightly split peak between 80 and 95 ◦ C, which was Step1 : NH3(g) + s⇄NH3(a) (8)
assigned to the co-reduction of Ni and Ru oxides. The fact that only one
low-temperature peak related to NiO and RuO2 reduction is present Step 2 : NH3(a) + s⇄NH2(a) + H(a) (9)
points out to a strong interaction between both metals in the prepared
bimetallic samples [46], which explains the superior catalytic perfor­
Step 3 : NH2(a) + s⇄NH(a) + H(a) (10)
mance of the bimetallic catalysts with respect to their monometallic
Step 4 : NH(a) + s⇄N(a) + H(a) (11)
counterparts.
The surface composition of the 5Ni1Ru/CeO2 catalyst was studied by
Step 5 : 2N(a) →N2(g) + 2s (12)
XPS (Fig. 8) and the surface atomic ratios Ni/Ce and Ru/Ce were 0.7 and
0.1, respectively, which yields a surface atomic Ni/Ru ratio of 7, a value
Step 6 : 2H(a) ⇄H2(g) + 2s (13)
which is clearly lower than the nominal Ni/Ru ratio determined by ICP-
MS (Ni/Ru = 7.2 on a weight basis or Ni/Ru = 12.4 on an atomic basis, where s represents a vacant site of the catalyst surface, (g) stands for gas
Table 1). This is in accordance to the HRTEM results, where RuO2 par­ and (a) stands for adsorbed. Our experimental tests (Fig. 3) showed that
ticles have a smaller size than NiO ones and, therefore, contribute more the rate of ammonia decomposition does not depend on the presence of
to the XPS signal. The deconvolution of the Ce 3d signal indicated that N2, but it is inhibited by the presence of H2. For this reason, Eq. 12 was
ca. 18 % of the ceria surface corresponded to Ce3+. The deconvolution of considered irreversible, while Eq. 13 was reversible. To determine the
the Ni 2p3/2 signal (Fig. 8a) showed bands at 854.3 and 855.9 eV, cor­ reaction rate equation, each step presented in Eqs. 8–13 was considered
responding to Ni2+ (96.6 %) and Ni3+ (3.4 %) species, respectively, as as the limiting step of the reaction, and the expression for which the sum
well as their corresponding satellite lines at higher binding energies. The of the square of the residuals (SSR) with respect to the experimental
deconvolution of the Ru 3p3/2 signal (Fig. 8b) showed bands at 460.9 results is minimum was chosen. The Model Selection Criterion (MSC) was
and 462.8 eV, which were ascribed to metallic Ru (25.4 %) and Ru4+ also used for the kinetic analysis (Eqs. 14, 15) [15]:

7
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

( ) ( )
SST 2p Table 3
MSC = ln − (14) Kinetic parameters of the ammonia decomposition rate equation for 5Ni1Ru/
SSR N
CeO2 and 2.5Ni0.5Ru/CeO2 catalysts.

N
Parameter 5Ni1Ru/CeO2 2.5Ni0.5Ru/CeO2
SST = (xn − x)2 (15)
n=1 k0 (mol NH3 g−cat1 s− 1) 3.9 × 106 ± 0.7 × 106 3.3 × 105 ± 0.8 × 105
Ek (kJ mol− 1) 124 ± 8 107 ± 11
Where p is the number of parameters and N the number of experimental K1 (m3 mol− 1) 1.1 ± 0.2 1.2 ± 0.3
K 4 (mol m− 3) 4.7 × 10− 2 ± 0.1 × 10− 2
4.5 × 10− 2 ± 0.2 × 10− 2
points. The variance of the SSR was calculated to further confirm the
model selection, using the expression (Eq. 16):
√̅̅̅̅̅̅̅̅̅̅̅̅ CeO2, respectively, and those estimated with the kinetic model. A good
VAR =
SSR
(16) correlation is observed within the limits of a ±10 % difference, and most
N− p of the values fall within the ±5 % difference. The largest deviation be­
tween experimental and predicted values is noticed at low temperatures
The comparison of the calculated statistical parameters is presented
due to low conversion values. Anyhow, no systematic deviations are
in Table 2. The kinetic expression obtained by considering the last step
evident, and the model can be considered robust. Fig. 9b and d show a
of the mechanism proposed as the determining step (Eq. 13, hydrogen
comparison between the values obtained experimentally (points) and
desorption) was disregarded, because it results in a zero-order kinetic
those obtained with the simulation (lines) in experiments carried out at
expression, which does not follow the qualitative behaviour observed
different contact times and temperatures for each catalyst. A good cor­
from the experimental data. The expression obtained considering the
relation in both cases is observed.
first dehydrogenation of ammonia (Eq. 9) as the limiting step of the
The kinetic model can now be used to simulate the behaviour of the
reaction was the one which best fit the experimental data obtained with
catalysts in order to obtain the best operation conditions for hydrogen
both 5Ni1Ru/CeO2 and 2.5Ni0.5Ru/CeO2 catalysts, as it presented the
production. Fig. 10 shows three-dimensional graphs showing the influ­
highest values of MSC and R2, and a low variance and SSR (Table 2).
ence of temperature and contact time on ammonia conversion using a
Therefore, the equation of the reaction rate r is (Eq. 17):
pure NH3 stream and the 2.5Ni0.5Ru/CeO2 catalyst. As expected,
r=
kK1 CNH3
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ 2 (17) ammonia conversion is strongly affected by both parameters, being high
(1 + K1 CNH3 + CH2 /K4 ) temperatures and high contact times preferred to maximize the con­
version of ammonia and the concomitant production of hydrogen. It can
where k is the specific constant of the reaction rate (mol NH3 g−cat1 s− 1), K1 be seen that the lowest temperature at which the complete conversion
is the ammonia adsorption constant (m3 mol− 1), K4 is the hydrogen (> 99 %) of ammonia is reached is 450 ◦ C under a pure flow of ammonia
desorption constant (mol m-3) and CNH3 and CH2 are the ammonia and using the catalyst under study, at a contact time of 4.6 s (F/W of 3.2 L
hydrogen concentrations (mol m-3). The specific constant k depends on g− 1 h− 1). If the goal is to optimize the hydrogen production rate, a
the frequency factor k0 (mol NH3 g−cat1 s− 1), on the activation energy Ek contact time of 1.6 s, which corresponds to a F/W of 9 L g− 1 h− 1, and a
(kJ mol− 1) and on the temperature T (K) through the Arrhenius relation temperature of 500 ◦ C, are a suitable combination of operating condi­
(Eq. 18): tions. Under these conditions, 13.5 LH2 g− 1 h− 1 can be obtained.

(18)
− Ek/
k = k0 e RT
4. Conclusions
The kinetic parameters obtained and their confidence intervals are
reported in Table 3. The values of the activation energy, 124 ± 8 kJ Several bimetallic Ni-Ru/CeO2 catalysts containing different
mol− 1 for 5Ni1Ru/CeO2 and 107 ± 11 kJ mol− 1 for 2.5Ni0.5Ru/CeO2, amounts of Ni and Ru were prepared by co-impregnation and tested in
are within the values reported in the literature for other catalysts. The the ammonia decomposition reaction to obtain hydrogen at atmospheric
lowest value of the activation energy, 21 kJ mol− 1, was reported by Itoh pressure. The best catalytic performance was obtained with catalysts
et al. [48] for Fe supported on ceria-zirconia. Deng et al. [49] reported containing between 3 and 6 wt.% metal loading with a Ni/Ru ratio of ca.
an activation energy of 54 kJ mol− 1 for Ni/Ce0.8Zr0.2O2; in the same 7 (w/w). TOFH2 values exceeding 2 s− 1 were obtained at 400 ◦ C, which
work they reported values of 71 and 99.5 kJ mol− 1 for Ni/CeO2 and are among the highest reported for ammonia decomposition. These
Ni/Al2O3, respectively. Hu et al. [50] reported an activation energy of catalysts showed an excellent stability in long-term and multicycle tests,
149.2 kJ mol− 1 for Ru/CeO2. The activation energies of 5Ni1Ru/CeO2 outperforming by a large extent those of the monometallic Ni/CeO2 and
and 2.5Ni0.5Ru/CeO2 fall between those of Ni/CeO2 and Ru/CeO2. Ru/CeO2 counterparts. The decomposition of ammonia over the bime­
Fig. 9a and c show parity plots that compare the ammonia conver­ tallic catalysts was inhibited by the presence of H2, but did not depend
sion values obtained experimentally for 5Ni1Ru/CeO2 and 2.5Ni0.5Ru/ on the presence of N2. A detailed characterization by Raman spectros­
copy, X-ray diffraction, high resolution transmission electron micro­
scopy, X-ray photoelectron spectroscopy, temperature-programmed
Table 2 reduction and H2 chemisorption showed the existence of an intimate
Statistical parameters of the rate expressions obtained according to the step contact between Ni and Ru and with the ceria support, which is
considered as the limiting step of the reaction.
considered the reason of the excellent catalytic activity and stability
5Ni1Ru/CeO2 Limiting step SSR Variance error MSC R2 (%) observed. The experimental data was used to obtain a kinetic model and
1 0.194 0.028 3.509 97 kinetic parameters. The estimated activation energy values for the
2 0.070 0.010 4.523 99 bimetallic Ni-Ru/CeO2 catalysts, 107− 124 kJ mol− 1, is in between those
3 0.069 0.010 4.503 99 of monometallic Ni/CeO2 and Ru/CeO2 reported in the literature. The
4 0.506 0.075 2.471 94
kinetic analysis suggested that the dehydrogenation of ammonia
5 0.080 0.012 4.278 99
adsorbed on the catalysts surface was the limiting step of the reaction on
2.5Ni0.5Ru/CeO2 Limiting step SSR Variance error MSC R2 (%)
Ni-Ru/CeO2. The kinetic model was used to simulate the hydrogen
1 0.305 0.051 2.593 94 production rate under different operation conditions using a pure
2 0.109 0.018 3.622 98 ammonia stream for practical application.
3 0.104 0.018 3.619 98
4 0.481 0.083 2.037 93
5 0.105 0.018 3.509 98

8
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

Fig. 9. a) Parity plot between experimental and predicted ammonia conversion values from the kinetic model developed for 5Ni1Ru/CeO2 and c) 2.5Ni0.5Ru/CeO2.
b) Experimental (points) and predicted (lines) ammonia conversion values at different contact time, Ar:NH3 = 1.2:1 for 5Ni1Ru/CeO2 and d) 2.5Ni0.5Ru/CeO2.

Declaration of Competing Interest

The authors have no competing interests to declare.

Acknowledgments

This work was funded by projects MICINN/FEDERRTI2018-093996-


B-C31 and GC 2017 SGR 128. IL is grateful to MINECO for PhD grant
BES-2016-076507. JL is a Serra Húnter Fellow and is grateful to ICREA
Academia program.

References

[1] G.W. Crabtree, M.S. Dresselhaus, M.V. Buchanan, The hydrogen economy, Phys.
Today 57 (2004) 39–45, https://doi.org/10.1063/1.1878333.
[2] C. Zamfirescu, I. Dincer, Ammonia as a green fuel and hydrogen source for
vehicular applications, Fuel Process. Technol. 90 (2009) 729–737, https://doi.org/
10.1016/j.fuproc.2009.02.004.
[3] J.W. Makepeace, T. He, C. Weidenthaler, T.R. Jensen, F. Chang, T. Vegge,
P. Ngene, Y. Kojima, P.E. de Jongh, P. Chen, W.I.F. David, Reversible ammonia-
based and liquid organic hydrogen carriers for high-density hydrogen storage:
recent progress, Int. J. Hydrogen Energy 44 (2019) 7746–7767, https://doi.org/
Fig. 10. Simulated results for ammonia conversion obtained over 2.5Ni0.5Ru/ 10.1016/j.ijhydene.2019.01.144.
CeO2 at different temperatures and contact times using pure ammonia. [4] S. Mukherjee, S.V. Devaguptapu, A. Sviripa, C.R.F. Lund, G. Wu, Low-temperature
ammonia decomposition catalysts for hydrogen generation, Appl. Catal. B Environ.
226 (2018) 162–181, https://doi.org/10.1016/j.apcatb.2017.12.039.
CRediT authorship contribution statement [5] X. Duan, G. Qian, X. Zhou, D. Chen, W. Yuan, MCM-41 supported Co-Mo bimetallic
catalysts for enhanced hydrogen production by ammonia decomposition, Chem.
Ilaria Lucentini: Investigation, Formal analysis. Germán García Eng. J. 207–208 (2012) 103–108, https://doi.org/10.1016/j.cej.2012.05.100.
[6] Z. Lendzion-Bieluń, W. Arabczyk, Fused Fe-Co catalysts for hydrogen production
Colli: Formal analysis. Carlos D. Luzi: Formal analysis. Isabel Serrano: by means of the ammonia decomposition reaction, Catal. Today 212 (2013)
Investigation. Osvaldo M. Martínez: Conceptualization, Methodology, 215–219, https://doi.org/10.1016/j.cattod.2012.12.014.
Supervision. Jordi Llorca: Conceptualization, Resources, Supervision.

9
I. Lucentini et al. Applied Catalysis B: Environmental 286 (2021) 119896

[7] B. Lorenzut, T. Montini, M. Bevilacqua, P. Fornasiero, FeMo-based catalysts for H2 Catal. B Environ. 52 (2004) 287–299, https://doi.org/10.1016/j.
production by NH3 decomposition, Appl. Catal. B Environ. 125 (2012) 409–417, apcatb.2004.05.002.
https://doi.org/10.1016/j.apcatb.2012.06.011. [30] V. Prasad, A.M. Karim, A. Arya, D.G. Vlachos, Assessment of overall rate
[8] S.B. Simonsen, D. Chakraborty, I. Chorkendorff, S. Dahl, Alloyed Ni-Fe expressions and multiscale, microkinetic model uniqueness via experimental data
nanoparticles as catalysts for NH3 decomposition, Appl. Catal. A Gen. 447–448 injection: ammonia decomposition on Ru/γ-Al2O3 for hydrogen production, Ind.
(2012) 22–31, https://doi.org/10.1016/j.apcata.2012.08.045. Eng. Chem. Res. 48 (2009) 5255–5265, https://doi.org/10.1021/ie900144x.
[9] I. Lucentini, A. Casanovas, J. Llorca, Catalytic ammonia decomposition for [31] Z. Wu, M. Li, J. Howe, H.M. Meyer III, S.H. Overbury, Probing defect sites on CeO2
hydrogen production on Ni, Ru and Ni–Ru supported on CeO2, Int. J. Hydrogen nanocrystals with well-defined surface planes by Raman spectroscopy and O2
Energy 44 (2019) 12693–12707, https://doi.org/10.1016/j.ijhydene.2019.01.154. adsorption, Langmuir 26 (2010) 16595–16606, https://doi.org/10.1021/
[10] E. García-Bordejé, S. Armenise, L. Roldán, Toward practical application of H2 la101723w.
generation from ammonia decomposition guided by rational catalyst design, Catal. [32] Y.M. Chen, A. Korotcov, H.P. Hsu, Y.S. Huang, D.S. Tsai, Raman scattering
Rev. - Sci. Eng. 56 (2014) 220–237, https://doi.org/10.1080/ characterization of well-aligned RuO2 nanocrystals grown on sapphire substrates,
01614940.2014.903637. New J. Phys. 9 (2007) 1–12, https://doi.org/10.1088/1367-2630/9/5/130.
[11] W. Raróg-Pilecka, Z. Kowalczyk, J. Sentek, D. Składanowski, D. Szmigiel, [33] B. Budiana, S. Suasmoro, X. Wang, J. Song, Raman scattering in nanosized nickel
J. Zieliñski, Decomposition of ammonia over potassium promoted ruthenium oxide NiO, J. Phys. Conf. Ser. 93 (2007) 1–5, https://doi.org/10.1088/1742-6596/
catalyst supported on carbon, Appl. Catal. A Gen. 208 (2001) 213–216, https://doi. 93/1/012039.
org/10.1016/S0926-860X(00)00721-3. [34] Z. Ma, S. Zhao, X. Pei, X. Xiong, B. Hu, New insights into the support morphology-
[12] A.K. Hill, L. Torrente-Murciano, Low temperature H2 production from ammonia dependent ammonia synthesis activity of Ru/CeO2 catalysts, Catal. Sci. Technol. 7
using ruthenium-based catalysts: synergetic effect of promoter and support, Appl. (2017) 191–199, https://doi.org/10.1039/c6cy02089e.
Catal. B Environ. J. 172–173 (2015) 129–135, https://doi.org/10.1016/j. [35] E. Sartoretti, C. Novara, F. Giorgis, M. Piumetti, S. Bensaid, N. Russo, D. Fino, In
apcatb.2015.02.011. situ Raman analyses of the soot oxidation reaction over nanostructured ceria-based
[13] W. Raróg-Pilecka, D. Szmigiel, Z. Kowalczyk, S. Jodzis, J. Zieliñski, Ammonia catalysts, Sci. Rep. 9 (2019) 1–14, https://doi.org/10.1038/s41598-019-39105-5.
decomposition over the carbon-based ruthenium catalyst promoted with barium or [36] A.M. Álvarez, M.Á. Centeno, J.A. Odriozola, Ru–Ni catalyst in the combined dry-
cesium, J. Catal. 218 (2003) 465–469, https://doi.org/10.1016/S0021-9517(03) steam reforming of methane: the importance in the metal order addition, Top.
00058-7. Catal. 59 (2016) 303–313, https://doi.org/10.1007/s11244-015-0426-5.
[14] S.F. Yin, B.Q. Xu, X.P. Zhou, C.T. Au, A mini-review on ammonia decomposition [37] H. Huang, Q. Dai, X. Wang, Morphology effect of Ru/CeO2 catalysts for the
catalysts for on-site generation of hydrogen for fuel cell applications, Appl. Catal. A catalytic combustion of chlorobenzene, Appl. Catal. B Environ. 158–159 (2014)
Gen. 277 (2004) 1–9, https://doi.org/10.1016/j.apcata.2004.09.020. 96–105, https://doi.org/10.1016/j.apcatb.2014.01.062.
[15] S. Armenise, E. García-Bordejé, J.L. Valverde, E. Romeo, A. Monzón, [38] M. Kurnatowska, L. Kepinski, W. Mista, Structure evolution of nanocrystalline Ce1-
A Langmuir–Hinshelwood approach to the kinetic modelling of catalytic ammonia xPdxO2-ymixed oxide in oxidizing and reducing atmosphere: reduction-induced
decomposition in an integral reactor, Phys. Chem. Chem. Phys. 15 (2013) activity in low-temperature CO oxidation, Appl. Catal. B Environ. 117–118 (2012)
12104–12117, https://doi.org/10.1039/c3cp50715g. 135–147, https://doi.org/10.1016/j.apcatb.2011.12.034.
[16] J.C. Ganley, A heterogeneous chemical reactor analysis and design laboratory: the [39] L. Bourja, B. Bakiz, A. Benlhachemi, M. Ezahri, S. Villain, J.R. Gavarri, Synthesis
kinetics of ammonia decomposition, Educ. Chem. Eng. 21 (2017) 11–16, https:// and characterization of nanosized Ce1-xbixo2− δ solid solutions for catalytic
doi.org/10.1016/j.ece.2017.08.003. applications, J. Taibah Univ. Sci. 4 (2010) 1–8, https://doi.org/10.1016/S1658-
[17] P. Yu, J. Guo, L. Liu, P. Wang, G. Wu, F. Chang, P. Chen, Ammonia decomposition 3655(12)60021-1.
with manganese nitride-calcium imide composites as efficient catalysts, [40] H. Yan, D. Zhang, J. Xu, Y. Lu, Y. Liu, K. Qiu, Y. Zhang, Yongsong Luo, Solution
ChemSusChem 9 (2016) 364–369, https://doi.org/10.1002/cssc.201501498. growth of NiO nanosheets supported on Ni foam as high-performance electrodes
[18] S. Chiuta, R.C. Everson, H.W.J.P. Neomagus, L.A. Le Grange, D.G. Bessarabov, for supercapacitors, Nanoscale Res. Lett. 9 (2014) 1–7, https://doi.org/10.1186/
A modelling evaluation of an ammonia-fuelled microchannel reformer for 1556-276X-9-424.
hydrogen generation, Int. J. Hydrogen Energy 39 (2014) 11390–11402, https:// [41] Y. Zhang, T. Ren, Silica supported ruthenium oxide nanoparticulates as efficient
doi.org/10.1016/j.ijhydene.2014.05.146. catalysts for water oxidation, Chem. Commun. 48 (2012) 11005–11007, https://
[19] R.W. McCabe, Kinetics of ammonia decomposition on nickel, J. Catal. 79 (1983) doi.org/10.1039/c2cc35272a.
445–450, https://doi.org/10.1016/0021-9517(83)90337-8. [42] G. Jacobs, P.M. Patterson, L. Williams, D. Sparks, B.H. Davis, Low temperature
[20] S. Armenise, F. Cazaña, A. Monzón, E. García-Bordejé, In situ generation of COx- water-gas shift: role of pretreatment on formation of surface carbonates and
free H2 by catalytic ammonia decomposition over Ru-Al-monoliths, Fuel 233 formates, Catal. Lett. 96 (2004) 97–105, https://doi.org/10.1023/B:
(2018) 851–859, https://doi.org/10.1016/j.fuel.2018.06.129. CATL.0000029536.52909.92.
[21] Z. Zhang, S. Liguori, T.F. Fuerst, J.D. Way, C.A. Wolden, Efficient Ammonia [43] M.K. Gnanamani, M.C. Ribeiro, W. Ma, W.D. Shafer, G. Jacobs, U.M. Graham, B.
decomposition in a catalytic membrane reactor to enable hydrogen storage and H. Davis, Fischer-Tropsch synthesis: metal-support interfacial contact governs
utilization, ACS Sustain. Chem. Eng. 7 (2019) 5975–5985, https://doi.org/ oxygenates selectivity over CeO2 supported Pt-Co catalysts, Appl. Catal. A Gen. 393
10.1021/acssuschemeng.8b06065. (2011) 17–23, https://doi.org/10.1016/j.apcata.2010.11.019.
[22] A.S. Chellappa, C.M. Fischer, W.J. Thomson, Ammonia decomposition kinetics [44] V. Sharma, P.A. Crozier, R. Sharma, J.B. Adams, Direct observation of hydrogen
over Ni-Pt/Al2O3 for PEM fuel cell applications, Appl. Catal. A Gen. 227 (2002) spillover in Ni-loaded Pr-doped ceria, Catal. Today 180 (2012) 2–8, https://doi.
231–240, https://doi.org/10.1016/S0926-860X(01)00941-3. org/10.1016/j.cattod.2011.09.009.
[23] A.R. De La Osa, A. De Lucas, A. Romero, J.L. Valverde, P. Sánchez, Kinetic models [45] W. Zheng, J. Zhang, Q. Ge, H. Xu, W. Li, Effects of CeO2 addition on Ni/Al2O3
discrimination for the high pressure WGS reaction over a commercial CoMo catalysts for the reaction of ammonia decomposition to hydrogen, Appl. Catal. B
catalyst, Int. J. Hydrogen Energy 36 (2011) 9673–9684, https://doi.org/10.1016/ Environ. 80 (2008) 98–105, https://doi.org/10.1016/j.apcatb.2007.11.008.
j.ijhydene.2011.05.043. [46] M.Z. Hossain, M.B.I. Chowdhury, Q. Alsharari, A.K. Jhawar, P.A. Charpentier,
[24] D.R. Mullins, S.H. Overbury, D.R. Huntley, Electron spectroscopy of single crystal Effect of mesoporosity of bimetallic Ni-Ru-Al2O3 catalysts for hydrogen production
and polycrystalline cerium oxide surfaces, Surf. Sci. 409 (1998) 307–319, https:// during supercritical water gasification of glucose, Fuel Process. Technol. 159
doi.org/10.1016/S0039-6028(98)00257-X. (2017) 55–66, https://doi.org/10.1016/j.fuproc.2017.01.013.
[25] M.C. Biesinger, B.P. Payne, L.W.M. Lau, A. Gerson, R.S.C. Smart, X-ray [47] Y.-J. Oh, S.H. Moon, C.-H. Chung, Thermodynamic stability of RuO2 bottom
photoelectron spectroscopic chemical state quantification of mixed nickel metal, electrodes and their effect on the Ba-Sr-Ti oxide film quality, J. Electrochem. Soc.
oxide and hydroxide systems, Surf. Interface Anal. 41 (2009) 324–332, https://doi. 148 (2001) F56–F62, https://doi.org/10.1149/1.1355689.
org/10.1002/sia.3026. [48] M. Itoh, M. Masuda, K. Machida, Hydrogen generation by ammonia cracking with
[26] D.J. Morgan, Resolving ruthenium: XPS studies of common ruthenium materials, Iron metal-rare earth oxide composite catalyst, Mater. Trans. 43 (2002)
Surf. Interface Anal. 47 (2015) 1072–1079, https://doi.org/10.1002/sia.5852. 2763–2767, https://doi.org/10.2320/matertrans.43.2763.
[27] R.P. Dias, C.S. Fernandes, M. Mota, J.A. Teixeira, A. Yelshin, Permeability and [49] Q.-F. Deng, H. Zhang, X.-X. Hou, T.-Z. Ren, Z.-Y. Yuan, High-surface-area
effective thermal conductivity of bisized porous media, Int. J. Heat Mass Transf. 50 Ce0.8Zr0.2O2 solid solutions supported Ni catalysts for ammonia decomposition to
(2007) 1295–1301, https://doi.org/10.1016/j.ijheatmasstransfer.2006.09.039. hydrogen, Int. J. Hydrogen Energy 37 (2012) 15901–15907, https://doi.org/
[28] N.J. Mariani, W.I. Salvat, O.M. Martínez, G.F. Barreto, Packed bed structure: 10.1016/j.ijhydene.2012.08.069.
evaluation of radial particle distribution, Can. J. Chem. Eng. 80 (2002) 186–193, [50] X.-C. Hu, X.-P. Fu, W.-W. Wang, X. Wang, K. Wu, R. Si, C. Ma, C.-J. Jia, C.-H. Yan,
https://doi.org/10.1002/cjce.5450800202. Ceria-supported ruthenium clusters transforming from isolated single atoms for
[29] S.J. Wang, S.F. Yin, L. Li, B.Q. Xu, C.F. Ng, C.T. Au, Investigation on modification hydrogen production via decomposition of ammonia, Appl. Catal. B Environ. 268
of Ru/CNTs catalyst for the generation of COx-free hydrogen from ammonia, Appl. (2020) 1–13, https://doi.org/10.1016/j.apcatb.2019.118424.

10

You might also like