Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Environmental Chemical Engineering 9 (2021) 106433

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Biohydrogen production by glycerol Aqueous-Phase Reforming: Effect of


promoters (Ce or Mg) in the NiAl2O4 spinel-derived catalysts
A. Morales-Marín a, J.L. Ayastuy a, *, U. Iriarte-Velasco b, M.A. Gutiérrez-Ortiz a
a
Chemical Technologies for the Environmental Sustainability Group (TQSA), Department of Chemical Engineering, Faculty of Science and Technology, University of the
Basque Country UPV/EHU, Sarriena S/N, 48940 Leioa, Spain
b
Department of Chemical Engineering, Faculty of Pharmacy, University of the Basque Country UPV/EHU, Paseo de la Universidad, 7, 01006 Vitoria, Spain

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

Editor: Dr. G. Palmisano Biohydrogen production by Aqueous-Phase Reforming (APR) of glycerol was studied over nickel aluminate
spinel-derived catalysts modified by Ce or Mg. As reference, an alumina supported nickel catalyst was also
Keywords: investigated. Catalysts were characterised by a sort of techniques in order to elucidate any structure-activity
Nickel relationship, and tested for the WGS and CO hydrogenation reactions. It was concluded that the spinel cata­
Promoter
lyst allowed a better dispersion of the metallic Ni, and led to a more APR active and more hydrogen-selective
Aqueous-phase reforming
catalyst. Mg and Ce doping further increased the nickel dispersion and increased basicity what could be
Glycerol
Biohydrogen related to their higher WGS relative reaction rate. Also, Mg doping hindered nickel reduction due to Ni-Mg
interactions. Both promoters improved the stability of the textural properties under APR conditions, mini­
mizing the formation of boehmite. Leaching of Ni was found, at lesser extent after Mg doping.

1. Introduction production in the last two decades has led to a huge surplus of glycerol in
the world market [12]. Its valorisation into higher value-added products
In this context of global warming, the use of H2 as an energy carrier would contribute to a higher profitability of biorefineries.
could be an interesting low-carbon solution. Nowadays, hydrogen is In order to maximize the hydrogen selectivity in the APR of polyols,
mainly produced from fossil-derived feedstocks, normally, through the the catalysts should promote C-H and C-C bond scission and deactivate
steam reforming (SR) of natural gas [1], which major drawback is the C-O bond cleavage reactions. In general, low acidity supports are
large amount of CO2 emitted [2]. Dumesic et al. [3] discovered in 2002 preferred. WGS-active catalysts could provide high hydrogen yields in
an alternative process to obtain biohydrogen from renewable sources at the aqueous-phase reforming what has been ascribed to their inherent
mild temperature (200–270 ºC) and moderate pressure (25–50 bar) by ability for removing the adsorbed CO molecules from the metallic sites
the Aqueous-Phase Reforming (APR) of biomass-derived oxygenates. It [13]. Group VIII metals, especially Pt, fulfil all the above requirements.
should be noted that APR is more advantageous than SR due to its lower However, cheaper metals should be investigated, as for example Ni,
energy requirements (i.e. no need of water evaporation), the reduction which has attracted considerable attention during the last years.
of undesirable side-reactions which decrease the hydrogen yield, the Another major consideration involved in designing an adequate APR
lower CO concentration in the gas, and the ease of separation of the APR catalyst is the hydrothermal stability of the support. Conventional
products. Sort of biomass-derived substrates have been used for γ-alumina is known to have poor thermal and hydrothermal stability
hydrogen production by APR, including ethanol [4], ethylene glycol [5], [14,15]. In our previous work, we investigated Ni-based catalysts
glycerol [6], xylitol [7], sorbitol [8], and glucose [9]. Among these, derived from spinel-like NiAl2O4 precursor [16] in the APR of glycerol.
glycerol is one of the most promising substrates. Glycerol, a major The use of nickel aluminate spinel improved the dispersion of the
by-product (E factor 0.1) of the biodiesel production process [10], can metallic phase and ensured the hydrothermal stability of the support. In
serve as a substrate for the production of various higher value-added addition, NiAl2O4 spinel contained oxygen vacancies, providing crucial
products, and thus, has been included as one of the 12 platform mole­ active sites for WGS activity [17].
cules for biorefineries [11]. The boom experienced by biodiesel With this background in mind, it might be useful to investigate new

* Corresponding author.
E-mail address: joseluis.ayastuy@ehu.eus (J.L. Ayastuy).

https://doi.org/10.1016/j.jece.2021.106433
Received 21 July 2021; Received in revised form 17 September 2021; Accepted 20 September 2021
Available online 6 October 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

methods to improve the activity and hydrogen selectivity of NiAl2O4- calculated from the integration of the H2-TPR profile and defined as
derived catalysts. The strategy followed in the present work is based on TPR950. Additional H2-TPR runs were carried out to determine the
the addition of basic oxides as promoters. Among others, Cu, La, Co, Zn reduction degree and nickel speciation. Firstly, the sample was reduced
and Br have been reported as structural modifiers of Ni/Al2O3 which at 700 ◦ C for 1 h, simulating the reduction step before the catalytic tests
effectively promoted the APR of glycerol [3,18]. Basic oxides such as (H2 uptake in this step was named TPR700). Then, after cooled down to
MgO, La2O3, CeO2 or CaO have been used to modify or neutralize the room temperature, second H2-TPR run was done up to 950 ◦ C (H2 uptake
acidic alumina [19,20]. Magnesium has been used as promoter of was named TPR700/950). Reduction degree (RDNi) was calculated, as
Ni/Al2O3 for the partial oxidation of methane [21] and glycerol APR follows:
[22] and provides structural stabilization of Ni/Al2O3 [23]. On the other
H TPR700
hand, the high redox capacity of ceria facilitates reduction of nickel RDNi (%) = 2
TPR700/950 × 100 (1)
species. Also, Ce-Ni interactions can promote the catalytic behaviour by H TPR + H2
700
2
creation of new species [24]. To the best of our knowledge, this is the
first study on the effect of Ce and Mg used as promoters of nickel where HTPR
2
i
is the hydrogen uptake in stage i. TPR of the spent catalysts
aluminate spinel catalyst for the use in the Aqueous-Phase Reforming. (TPRspent) was carried out with the similar protocol of TPR950, but
Ni/Al2O3 (by impregnation) and NiAl2O4 (by coprecipitation) doped removing the cold trap and following the releases gases by MS (Hiden
with Ce or Mg (at 5 wt%) were synthesized. These were thoroughly Analytical).
characterised by XRD, nitrogen adsorption-desorption isotherms, The micrographs along with elemental EDX-maps of the ex-situ
H2-TPR, electronic microscopy, ICP, CO2-TPD and NH3-TPD techniques, reduced catalysts were obtained by a FEI Titan Cubed G2 60–300 mi­
and tested in the APR of glycerol, CO hydrogenation and WGS reactions croscope, composed of a high-brightness X-FEG Schottky field emission
in order to elucidate any structure-activity relationship. electron gun, monochromator, CEOS Gmbh spherical aberration
corrector and Super-X EDX system with High-Angle Annular Dark-Field
2. Experimental (HAADF) detector for Z contrast imaging in Scanning Transmission
Electron Microscopy (STEM) configuration. The metallic nickel particle
2.1. Catalysts preparation size distribution was determined by measuring the diameter (d) of at
least 240 particles. After that, the average nickel dispersion (DNi0) was
Nickel aluminate spinel with Al/Ni = 2 (mol ratio) was synthesized estimated applying D-FE model [25].
by coprecipitation. Details on the synthesis can be found elsewhere [16]. The average diameter of the Ni0 particle (dNi0 ,TEM ) was estimated
The as-prepared solid was divided into three parts: one part was no from volume to surface ratio. Metallic surface area, (SNi0) was calculated
further modified (labelled NiAl) and the other two parts were subjected taking as follows:
to either Ce or Mg addition (5 wt%) by wet-impregnation (WI). Cerium /
ANio ⋅0.01⋅RDNi ⋅DNi0 ⋅NA
(III) nitrate hexahydrate (Ce(NO3)3.6 H2O, Sigma-Aldrich) and magne­ SNi0 (m2 g) = (2)
MWNi
sium nitrate hexahydrate (Mg(NO3)2.6 H2O, Sigma-Aldrich) were used
as metal precursors. The obtained solids were dried overnight at 110 ºC where, ANio is the atomic area of Ni, NA the Avogadro number and MWNi
in an oven and finally calcined in a muffle. Ce/NiAl catalyst was calcined the molecular weight of Ni.
at 600 ºC for 2 h with a heating rate of 5 ºC/min, while Mg/NiAl catalyst The population, strength and distribution of acid and basic sites were
was calcined at 850 ºC for 4 h with a heating rate of 10 ºC/min. For determined by NH3 and CO2 temperature programmed desorption,
comparison purposes, Ni (from nickel (II) acetate tetrahydrate precur­ respectively, in a Micromeritics AutoChem 2920 equipment. Before both
sor) was also impregnated onto γ-Al2O3 (labelled Ni/Al) at 18 wt% Ni analysis, the sample was cleaned by flowing He at 550 ◦ C for 1 h. After
loading in order to match the extra lattice nickel content of NiAl spinel cooled down to room temperature, the catalyst was reduced at 700 ◦ C
reduced at 700 ºC for 1 h [16]. All the samples were pelletized at for 1 h, and then cooled down to room temperature under He flow. For
0.04–0.16 mm. APR exhausted catalysts were named adding -U to their acidity evaluation, a sequence of 10% NH3/He pulses were injected at
label. 90 ◦ C up to saturation of the sample, and flushed with He flow in be­
tween each ammonia pulse. Then, the sample was heated up to 850 ◦ C at
2.2. Catalyst characterization studies 10 ºC/min under He and the desorbed NH3 was recorded. For basicity
evaluation, the sample was saturated by flowing 50 mL/min 5% CO2/He
ICP-AES was used to analyse the bulk chemical composition of the at 50 ◦ C for 1 h, then flushed out with He flow for 1 h. Then, the sample
samples and the presence of leached metals in the post-reaction liquid was heated up to 850 ◦ C at 10 ºC/min under He and the CO2 desorbed
solutions. The textural properties of the prepared solids were evaluated was recorded.
from N2 adsorption-desorption isotherms at 77 K in a Micromeritics
TriStar II 3020 equipment. Before the experiment, the solid was
degassed at 300 ◦ C for 10 h in order to clean the surface. The specific 2.3. Catalytic performance evaluation
surface area and pore size distribution were determined by BET and BJH
formalisms, respectively. Aqueous-Phase Reforming (APR) of glycerol aqueous solution (10 wt
Powder XRD was measured by means of a PANalytical Xpert PRO % glycerol) was carried out in a fixed-bed reactor (PID Eng&TEch) at
diffractometer in 2θ range 10–80º, at 0.026º steps. The crystallite size 235 ºC and 35 bar and WHSV = 24.5 h− 1. Catalyst was reduced in-situ
was calculated from the X-ray line broadening analysis. The identifica­ with 20% H2/He flow at 700 ºC and atmospheric pressure, for 1 h.
tion of the crystal phases was carried out on the basis of ICDD database. Hereafter, reactor was pressurised with He up to the desired pressure,
Reducibility of the solids was evaluated by temperature- then the He flow was switched to bypass and 0.2 mL/min of the glycerol
programmed reduction with hydrogen (H2-TPR) performed in a Micro­ solution was pumped into the reactor while the temperature was pro­
meritics AutoChem 2920 apparatus. The sample was placed in the U- gressively raised at 5 ºC/min up to the reaction temperature. Catalytic
shaped quartz reactor. Firstly, the solid was cleaned by flushing 50 mL/ performance was measured at 3 h TOS (time on stream). The reactor
min of He at 550 ºC for 1 h, and subsequently cooled down to 50 ºC. effluent was cooled down and the gas and liquid phases separated. The
Then, analysis was carried out by flowing 5% H2/Ar stream though the gaseous products were swept away with a He flow (40 mL/min). Gas
sample and rising the temperature from room temperature to 950 ◦ C at product distribution was continuously analyzed by a μGC (Agilent). The
10 ◦ C/min, monitoring the signal by TCD. The water formed during the liquid product was collected every 15 min in glass vials and off-line
reduction was trapped using a cold trap. The total H2-uptake was analysed by GC-FID (Agilent, 6890 N) and HPLC-RI (Waters, Hi-Plex H

2
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

column). Total organic carbon (TOC) in the liquid phase was measured 100
XGly
on a Shimadzu TOC-L apparatus. The carbon balance was above 97% for (a)
90 Xgas 10
all the experiments.
YH2
Glycerol conversion (XGly) and conversion to gas (Xgas) were calcu­
80
lated as follows: 8
70
F in out
Gly − F Gly
XGly (%) = 100⋅ (3)
F in 60 6

Conversion (%)
Gly

YH2(%)
F in out 50
C atoms − F C atoms,liquid
Xgas (%) = 100⋅ (4)
in
FC atoms 40 4

Selectivity to hydrogen (SH2) was calculated as follows: 30


out
FH2 2
SH2 (%) = 100⋅ out (5) 20
FH2 + (n + 1)⋅FCoutn H2n+2
10
The selectivity to alkanes was calculated as follows: 0

F out 0
SAlkanes (%) = 100⋅ C atoms in Alkanes
(6) Ni/Al NiAl Ce/NiAl Mg/NiAl
F out
C atoms

Hydrogen yield (YH2) was calculated as follows: 5


70
(b) SH2 SAlkanes
F out 1 H2 production rate
YH2 (%) = 100⋅ inH2 ⋅ (7)
F Gly 7
60 4
The yield of the liquid compounds (YLiq,i) was calculated as the

H2 production rate (mmolH2/gNi·min)


carbon molar flow in a given liquid compound (FC_Liq,i) divided by the 50
total carbon flow in the liquid products (FC_Liq): Selectivity (%)
3
FC Liq,i 40
YLiq,i = 100 × (8)
FC Liq
30 2
CO hydrogenation and WGS reactions were performed in a fixed bed
reactor (P&ID EngTech) at 350 ◦ C and atmospheric pressure, at GHSV
12,000 h− 1. The feedstream composition was (in vol%): CO/H2/He: 7.5/ 20
10/82.5 for CO hydrogenation; and CO/H2O/H2/He: 7.5/25/10/57.5 1
for WGS reaction. 10
CO conversion and selectivity to CO2 (SCO2) in CO hydrogenation and
WGS experiments were calculated as follows: 0 0

F in − F out Ni/Al NiAl Ce/NiAl Mg/NiAl


XCO (%) = 100⋅ CO CO
(9)
Fin
CO Fig. 1. (a) Glycerol conversion, conversion to gas and hydrogen yield; (b)
Hydrogen and alkane selectivity (left axis) and H2 production rate per nickel
Fout
SCO2 (%) = 100⋅ in CO2 out (10) mass (right axis). Reaction conditions: 235 ◦ C/35 bar, WHSV = 24.5 h− 1, 10 wt
F CO − F CO % glycerol in water mixture. Data evaluated at 3 h TOS.

3. Results and discussion reflected that APR of glycerol produced a significant amount of liquid-
phase products in addition to hydrogen.
3.1. Catalyst activity Fig. 1b allows us to compare the catalyst performance in terms of
hydrogen and alkane selectivity. Catalyst Ni/Al showed high hydrogen
Glycerol conversion (XGly), conversion to gas (Xgas) and H2 yield selectivity (74%) as due to its poor C–C cleavage activity. Its counter­
(YH2) are shown in Fig. 1. The Weisz-Prater and Mears criteria were part NiAl had 55% selectivity to hydrogen. Ce and Mg doped catalysts
applied and the obtained values (Table S1, Supporting Information) slightly enhanced the SH2 of the parent NiAl (SH2 around 60%). The
indicated that the catalytic APR experiments were done under kinetic highest selectivity to alkanes was achieved by catalyst Ni/Al (SAlkanes
regime, that is, in the absence of mass-transfer limitations. = 56%), with similar ethane and methane production (Table S2, Sup­
Differences were observed in the performance of the unpromoted porting Information). Its counterpart NiAl, showed a much lesser
catalysts, NiAl and Ni/Al (Fig. 1). The sample NiAl was much more selectivity of 38% (close to that obtained by the promoted assays). For
active than its counterpart Ni/Al (XGly: 13% vs 65%; Xgas: 2% vs 27%; the coprecipitated assays, the methane to alkanes mole ratio was above
YH2: 0.3% vs 8%). Clearly, spinel-derived nickel nanoparticles were 0.90. Ce/NiAl showed the highest hydrogen production-rate, of
much more active for hydrogen production by APR than those obtained 3.4 mmolH2/min⋅gNi. The moles of hydrogen produced per mole of
by impregnation onto alumina. Thus, the promoter (Ce or Mg) was glycerol converted (Table 1) was ten-fold larger in the spinel-based
finally added to NiAl catalyst instead of to Ni/Al. catalysts and among them, promoted catalysts increased by 10% that
Fig. 1a show that doping of Mg did not improve glycerol conversion. of NiAl.
Moreover, XGly slightly decreased after Mg addition (64% vs. 60%). Table 1 revealed that H2, CO2, and CH4 were the main products of the
Likewise, a detrimental effect was observed in Xgas (27% vs 24%) APR of glycerol (accounted for more than 98% of the detected com­
whereas the YH2 barely increased (8.0% vs 8.3%). Contrarily, Ce created pounds). CO was also detected, although in much lower amounts. H2
an advantageous effect on the catalyst performance. The marked dif­ concentration varied between 50% and 85% (maximum by catalyst Ni/
ferences between the conversion values (XGly 64–73% vs. Xgas 24–34%) Al). CO2 was the second most abundant compound (27–29%), except for

3
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

Table 1 liquid phase. Several oxygenated liquid compounds were detected:


Gas and liquid products in glycerol APR. propylene glycol, ethylene glycol, ethanol, hydroxyacetone, methanol
Ni/Al NiAl Ce/NiAl Mg/NiAl and acetaldehyde (accounting for over 97% of the carbon in the liquid
H2 production (molH2/molGly converted) 0.09 0.87 0.96 0.96 phase). Propionaldehyde and propanols were detected in trace
Gas-phase (%) quantities.
H2 84.7 49.8 49.7 54.6
CO2 3.3 28.4 29.3 27.0
CH4 5.9 18.7 19.2 16.7 3.2. General reaction path
H2/CO2 25.7 1.8 1.7 2.0
CO/H2 ~0 0.03 0.01 0.01 The obtained gas and liquid product distribution evidenced the
Liquid-phase yields (%) occurrence of both dehydration and hydrogenation reactions. With the
Hydroxyacetone 46.3 9.1 9.1 7.6
Propylene glycol 23.3 40.0 39.1 42.7
observed product distributions for glycerol APR a plausible reaction
Ethylene glycol 4.1 27.4 21.9 23.5 pathway with two reaction paths consistent with bifunctional (metallic
Ethanol 20.4 17.9 22.9 19.3 and Lewis acid functions) nature is depicted in Scheme 1.
Acetaldehyde 1.6 0.7 0.4 0.5 Compared to Ni/Al catalyst, spinel-derived catalysts showed boosted
Methanol 1.9 4.0 5.2 5.6
yield of H2 per mole of glycerol converted. Among the both promoted
Type of scission (liquid-phase products)
C-C only 6.0 31.4 27.2 29.1 catalysts gave the maximum H2 production of 0.96 molH2/molGly con­
C-O 91.6 67.6 71.6 70.2 verted. Dehydrogenation of glycerol followed by decarbonylation (C–C
cleavage) and subsequent WGS reaction led to the formation of H2 and
CO2 (Path I, over metallic sites) [35,36]. When dehydration (C–O
catalyst Ni/Al (< 4%). CH4 concentration varied between 6% and 19%, cleavage) occurs, H2 is consumed in the subsequent hydrogenation re­
being catalyst Ni/Al, again, the one with the lowest value. In an ideal actions (Path II, over acidic and metallic sites) that led to the formation
glycerol APR reaction, stoichiometric H2/CO2 ratio should be 2.3. It of oxygenated liquids, which can be thereafter converted to alkanes.
ranged from 1.7 to 2.0 for our spinel-derived catalysts, indicating that Thus, catalysts with enhanced capacity for C–C bond cleavage typically
some hydrogen was consumed in parallel reactions. Indeed, nickel is favour hydrogen production while a high activity in C–O bond cleavage
known to be very active in the CO and CO2 methanation reactions is typically associated with alkane formation [37,38].
[26–28]. The primary product from Path I (glyceraldehyde) was not detected,
To have light on the gas product distribution, WGS and CO hydro­ due to its high reactivity [39]. The main liquid product of Ni/Al was
genation activity was also investigated (Table 2). Methane was the only hydroxyacetone. Products from Path I were detected at trace levels, in
product in the absence of water and, under WGS conditions, both CO2 line with its extremely low H2 production rate. The spinel-derived cat­
and CH4 were formed. alysts showed a different liquid-phase composition. These produced
In both reaction, NiAl showed higher XCO than its counterpart Ni/Al. mainly propylene glycol, product of C–O cleavage (Path II), followed by
Feeding only CO and H2 the promoted catalysts showed around ten-fold ethylene glycol, product of C–C bond cleavage (Path I). The spinel se­
increased CO conversion with respect NiAl (i.e. XCO>30% vs XCO = 3%). ries showed a greater propylene glycol to hydroxyacetone mole ratio,
This was in line with others that reported that both Ce and Ce promote suggesting an important hydrogenation activity. Interestingly, the Mg
CO hydrogenation reaction [29,30]. Under WGS conditions, these dif­ and Ce doped assays, showed a subtle tendency to decrease the yield of
ferences further increased as XCO raised up to 67% by catalyst Mg/NiAl ethylene glycol and, concomitantly, increase that of small chain alcohols
(XCO < 5% by the unpromoted catalysts). NiAl was 20% more active (i.e. ethanol). This behaviour was ascribed to the ability of Mg and Ce
than Ni/Al for hydrogen production by WGS. Moreover, Ce or Mg doped assays for C–C bond cleavage, in line with their higher selectivity
doping further increased the WGS activity. It is well known that basic to hydrogen.
surfaces promote WGS reaction [31] by facilitating the water dissocia­ Liquid products were grouped by the type of bond scission (Table 1),
tion [32,33]. By means of WGS, CO can be converted into CO2 giving and the most outstanding difference was between Ni/Al and spinel-
additional H2. Therefore, catalysts that promote WGS are crucial for derived catalyst, the former producing vastly C–O scission products,
obtaining a high hydrogen selectivity [13]. However, a high selectivity while no appreciable differences among the doped catalysts and their
to methane was also shown (SCH4 about 30% for both promoted cata­ parent NiAl. It can be concluded that the catalyst synthesis method
lysts), in line with the high ability of nickel for CO methanation [34]. strongly affected its relative activity for the C–C to C–O bond scission.
The CO/CO2 molar ratio during the APR of glycerol (Table S2, The high hydrogenation activity and selectivity to methane observed
Supporting Information) was three-fold lower in the promoted catalysts, for the promoted catalysts suggested that CH4 was mainly produced
in line with the estimated relative WGS reaction rates (Table 2), and through hydrogen consuming reactions (both CO and CO2 hydrogena­
suggested that both Ce and Mg increase the CO conversion through the tion), with the concomitant decrease in hydrogen selectivity. As seen
WGS reaction. This was corroborated by the three-fold lower CO/H2 above CO2 was preferentially produced over CH4 (CO2/CH4 ratio:
values measured for both promoted catalysts (Table 1) in glycerol APR. 1.5–1.6) for all the spinel-derived catalysts, what suggested that Path I
With respect to the liquid-phase product distribution (Table 1), it was promoted.
should be noted that a significant fraction of carbon atoms in the feed
(98% for Ni/Al, 66–76% for spinel-derived catalysts) was retained in the 3.3. Materials characterization

3.3.1. Textural properties


Table 2 The specific surface area (SBET) of the catalysts is shown in Table 3.
CO hydrogenation and WGS activity results. All the calcined solids presented Type IV isotherms with H2 hysteresis
CO+H2 CO+H2 +H2O (Fig. S1, Supporting Information), with the P/P0 position of the inflec­
Catalyst XCO SCO2 XCO SCO2 Normalized rate of H2 tion point corresponding to the mesoporous range. Catalyst NiAl had a
(%) (%) (%) (%) production (WGS) 40% larger specific surface area than Ni/Al. In addition, pore volume
Ni/Al 2.0 100 2.5 68.1 1.0 and the average pore size were two-fold increased (Table S3, Supporting
NiAl 3.0 100 4.5 55.1 1.2 Information). With promoter, the SBET decreased due to pore blockage.
Ce/NiAl 34.6 100 45.4 30.3 1.5 This phenomenon was more pronounced for the Mg doped sample (6%
Mg/ 31.0 100 67.0 35.0 1.4 and 29% decrease in Ce/NiAl and Mg/NiAl, respectively) due to its
NiAl
higher calcination temperature. The average pore size of the parent NiAl

4
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

Scheme 1. Possible reaction pathways for the glycerol APR.

Table 3
Bulk composition and textural properties of the prepared samples.
Sample Metal loading (wt%) SBET (m2/g) Metal leached (%)

Ni Promoter Calcined Reduced Spent Ni Promoter Al

Ni/Al 19.4 n.a. 70 76 122 0.2 n.a. 0.03


NiAl 32.0 n.a. 98 83 116 0.7 n.a. 0.05
Ce/NiAl 30.3 5.5 92 85 97 1.6 0.24 0.03
Mg/NiAl 28.3 4.6 70 71 81 0.4 13.2 0.05

was preserved in the promoted samples.


The effect of reduction treatment (20% H2 flow up to 700 ºC, hold
1 h) on the texture was also investigated. In the case of the unpromoted
solids, an opposed behaviour was observed where SBET of Ni/Al
increased (i.e. 9%) while that of NiAl decreased by 15%. For the pro­
moted samples, SBET of Ce/NiAl decreased by 8% and no appreciable
variation was observed for Mg/NiAl sample. Indeed, the non-reducible
nature of MgO, in contrast to the high reducibility of ceria, could sta­
bilize the support [40].
With respect to the exhausted catalysts, a notable increase in SBET
was observed (Table 3). Among the unpromoted catalysts, Ni/Al-U
showed the highest increase (i.e. 60%) while its counterpart NiAl-U
increased by 40%. The increase in SBET was ascribed to the hydration
of γ-alumina. Boehmite could be leached-off and re-deposited on the
catalyst surface, generating additional porosity in the solid [41,42]. The
notable increase of the SBET of spent Ni/Al is in line with its higher
content in γ-alumina. The promoted catalysts showed the lowest varia­
tion in SBET (i.e. 5–14% increase) suggesting that both Ce and Mg doping
were helpful to maintain the stability of the textural properties under
harsh hydrothermal conditions.
With phase modification of the support metal particle detachment
can occur, causing sintering or leaching [43]. The liquid product of the
APR (Table 3) showed 0.2–1.6% Ni leaching. Leaching of Al was low
(0.05%), due to its re-deposition on the surface [44]. Ce leaching was
very low (i.e. 0.24%) as compared to magnesium (i.e. 13%), in line with
literature [45,46]. In addition, Mg could also be detached by the for­
mation and posterior elution of Mg organometallic complexes.
Fig. 2. XRD spectra of the calcined, reduced (-R) and spent (-U) catalysts. (a)
3.3.2. XRD analysis Ni/Al, (b) NiAl, (c) Mg/NiAl and (d) Ce/NiAl. (Symbols: Δ NiAl2O4; • Ni0; ο
XRD of the calcined Ni/Al contained three crystallographic phases, NiO; ◆ MgNiO2/MgO; ❑ CeO2; þ AlO(OH); × CeCO3OH).
namely, γ-Al2O3 (JCPDS 01–079–1558), NiO (JCPDS 01–078–0643) and
NiAl2O4 spinel (JCPDS 00–010–0339) (Fig. 2). NiO peaks were narrow, clusters. NiAl2O4 was formed upon calcination, as high temperatures
indicative of the presence of large crystallites, consistent with the used favour the migration of nickel atoms in close contact with alumina into
impregnation method, which favours formation of large metal oxide the alumina framework [47]. NiO was not detected in the reduced

5
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

samples while new diffraction peaks corresponding to Ni0 (JCPDS characteristic of metallic Ni phase could be observed. With respect to its
01–087–0712) appeared, indicative of the complete reduction of the counterpart NiAl-U, the absence of boehmite peaks suggested that spinel
nickel oxide particles. In addition, characteristic peaks of the structure was mostly retained under hydrothermal conditions. Metallic
spinel-phase were still identified, reflecting its low reducibility. Ni was detected for NiAl suggesting that the spinel-derived catalysts are
The XRD of the calcined NiAl matched that of NiAl2O4 spinel. Unlike promising for the aqueous-phase reactions, due to their higher phase
for catalyst Ni/Al, NiO was not detected, what suggested that Ni was stability, as compared to alumina-supported catalyst. In fact, it is
mainly forming the spinel or existed as fine particles (2–5 nm). Upon accepted that the presence of metal hinders the hydration of γ-alumina
reduction, the presence of metallic Ni was confirmed, together with to boehmite [54].
NiAl2O4 spinel. During the reduction, nickel cations located in the spinel Mg better stabilized the catalyst structure as compared to Ce. XRD of
framework can readily migrate to the surface, leaving the bulk enriched Mg/NiAl-U was similar to that of the fresh (reduced) sample, which
in γ-Al2O3 [48]. γ-Al2O3 cannot be detected by XRD due to its amor­ contained both spinel and Ni0 phases. No other crystalline phase was
phous nature. detected by XRD. Ce/NiAl-U showed a more complex XRD pattern.
In Ce/NiAl and Mg/NiAl solids, spinel was clearly identified. In Apart of the spinel and Ni0 peaks, a sort of narrow and intense peaks of
catalyst Ce/NiAl characteristic diffraction peaks of CeO2 (JCPDS cerium carbonate hydroxide (JCPDS 00–032–0189) appeared, formed
00–034–0394) could be observed. Catalyst Mg/NiAl showed peaks by the carbonation of ceria by the gas-phase CO [55]. Characteristic
characteristic of MgO (JCPDS 00–045–0946) and MgNiO2 (JCPDS peaks of ceria still could be observed, suggesting oxidation of this
00–024–0712). In sample Mg/NiAl, diffraction peaks of the spinel-phase sample. The absence of boehmite in both promoted spent samples
shifted to lower 2θ as compared to parent NiAl (Fig. S2, Supporting indicated that both promoters increased the hydrothermal resistance,
Information), due to the formation of Ni(Mg)Al2O4 spinel. The high especially in the case of the Mg-doped sample, which even preserved the
calcination temperature applied to the Mg-doped assay could promoted size of the Ni0 crystallites during the APR reaction (i.e. 7–8 nm, Table 4).
the diffusion of Mg2+ ions (ionic radius 0.72 Å) to the sites occupied by
Ni2+ ions (ionic radius 0.69 Å) in the original spinel. The increase in the 3.3.3. H2-TPR analysis
lattice parameters (8.0397 vs 8.0474 Å for NiAl and Mg/NiAl, respec­ The reduction profile of the investigated catalysts is shown in Fig. 3,
tively) confirms the incorporation of Mg ions. Also, formation of the and the H2 consumption is summarised in Table 5. Catalyst Ni/Al
MgO-NiO solid solution was favoured by the high calcination tempera­
ture [49].
Upon reduction of the promoted assays, the spinel-phase could be TPR950 TPR700/950
still detected whereas no characteristic diffraction peaks of NiO were Mg/NiAl
observed. The absence of characteristic peaks for any Ce-containing
phase suggested that Ce2O3 particles existed at below XRD detection
limit or as an amorphous phase [50]. A similar outcome could be ach­
ieved for catalyst Mg/NiAl since no characteristic diffraction peaks of
MgO nor MgNiO2 were detected. Again, in the reduced Mg/NiAl sample,
CeO2 (x10)
diffraction peaks of the spinel-phase downshifted (Fig. S2, Supporting Ce/NiAl
Information), suggesting the presence of Ni(Mg)Al2O4 spinel [20]. This
hypothesis was confirmed by a reduction experiment (up to 1000 ºC in
H2 uptake (a.u.)

5% H2/Ar flow) carried out with a 10 wt% Mg loaded NiAl2O4. Phase


evolution was analysed by XRD (Fig. S3, Supporting Information) which
showed a downshift ascribed to the formation of Ni(Mg)Al2O4.
The average values of the metallic Ni crystallites are given in Table 4. NiAl
By far, Ni/Al contained the largest metallic Ni crystallites (31 nm).
Much finer Ni particles were formed in catalyst NiAl (15 nm). A further
decrease in Ni0 size was observed for both promoted catalysts (decrease
by 50%), suggesting that the addition of Mg and Ce enhanced the
resistance against sintering during reduction. Ni/Al
In order to gain knowledge on the effect of APR reaction conditions
on the structural properties of the catalysts, XRD of the spent catalysts
were also collected (Fig. 2). Interestingly, none of the catalysts presented
characteristic peaks of carbonaceous material, due to the basic charac­
teristics of the catalyst surface [51]. Diffraction peaks from boehmite
(JCPDS 01–083–2384) were identified in the exhausted Ni/Al assay. As 200 400 600 800 200 400 600 800
previously noted, under harsh hydrothermal conditions, γ-AlO(OH) can Temperature (ºC)
be readily formed by hydration of γ-Al2O3 [52,53]. Spinel phase re­
Fig. 3. H2-TPR profiles of catalysts. TPR950: calcined samples; TPR700/950: fresh
flections were visible, suggesting that the hydration of alumina was
reduced at 700 ºC for 1 h.
limited to the surface. Interestingly, intense and narrow peaks

Table 4
Particle size of metallic Ni, exposed metal surface area, and acid/base characteristics of the fresh catalysts. Particle size also given for used assays.
Sample Ni0 size (nm) DNi0 (%) SNi0 (m2Ni0/g) Sites density RBasic/Acid
2 d 2 e
Fresh Used Basic (µmolCO2/m ) Acid (µmolNH3/m )

XRD TEM XRD

Ni/Al 31 48 31 1.7 1.3 0.46 1.86 0.25


NiAl 15 12 10 10.0 14.1 1.12 1.95 0.57
Ce/NiAl 7 15 26 11.2 14.7 2.03 2.49 0.82
Mg/NiAl 6 11 8 11.3 9.1 2.59 2.55 1.02

6
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

Table 5
Results from H2-TPR of the calcined (TPR950), fresh reduced (TPR700/950) and spent (TPRspent) catalysts.
Sample H2 uptake TPR950 TPR700 H2 uptake TPR700/950 H2 uptake RDNi Spent catalysts
(mmolH2/gcat) (mmolH2/g (mmolH2/g) (%)
Nickel speciation (%)

Peak Peak Peak TPRspent H2 uptake % Ni oxidized


(1)
α β γ (mmolH2/g)

Ni/Al 3.0 59.0 0 41.0 2.4 0.6 80 0.98 41


NiAl 5.6 4.1 50.0 45.9 5.0 0.6 89 1.85 37
Ce/NiAl 5.6 8.0 35.5 56.5 4.9 0.7 88 1.73 35
Mg/ 4.9 3.3 59.6 37.1 2.8 2.1 57 1.72 51
NiAl
(1)
percent oxidized with respect to that reduced at 700 ºC (TPR700)

showed two reduction events indicating the presence of, at least, two contrary, the upshift in sample Mg/NiAl (844 ºC) suggested the forma­
Ni2+ species, namely NiO and the nickel in spinel phase. The low tem­ tion of Ni(Mg)Al2O4 spinel during reduction, in line with XRD data, and
perature peak (peak-α, with its maximum at around 460 ºC) was indicated that reduction of nickel was hindered by the strong interaction
assigned to NiO particles weakly interacting with the alumina support. with MgO. It is interesting to note that, for all the samples, the sum of
The asymmetry of peak-α suggested the presence of NiO particles of hydrogen uptake from TPR700 and TPR700/950 coincided with that from
different size and varying strength of interaction with the support [56]. TPR950. The absence of hydrogen consumption below 700 ºC in the
Deconvolution allowed to identify four peaks centred at around 230, TPR700/950 run indicated that NiO particles were readily reduced in all
310, 415 and 470 ºC. The high temperature peak (peak-γ, centred at catalysts by the established pre-reaction reduction protocol. Also, all the
around 810 ºC) was ascribed to nickel in the spinel structure (i.e. nickel in the non-stoichiometric spinel phase (peak-β) of catalysts Ni/Al,
NiAl2O4) [16]. About of 59% of the Ni remained as free NiO (Table 3). NiAl and Ce/NiAl was readily reduced. Moreover, 51%, 77% and 78%,
Catalyst NiAl showed a large and intense reduction event starting at respectively, of the nickel in the spinel phase was reduced. As previously
225 ºC. The left tail indicated the existence of different nickel species. noted, catalyst Mg/NiAl showed quite a different reduction behaviour.
Deconvolution allowed differentiating three main peaks: peak-α (surface This case, around 90% of the nickel in the non-stoichiometric phase was
free NiO particles, at around 400 ºC), peak-β (sub-stoichiometric nickel reduced, however, that in the spinel phase (peak-γ) remained unre­
spinel, at around 685 ºC) and peak-γ (NiAl2O4, at around 790 ºC). Free duced. These results coincide with those obtained by Parmalina et al.
NiO accounted for less than 5% of the total Ni for this sample. Non- who reported that Mg-Ni interactions can hinder the reducibility of
stoichiometric and stoichiometric nickel aluminate spinel were formed nickel in Ni(Mg)Al2O4 spinels [49,60].
in similar amounts during the calcination. Oxidation of the active metallic phase has been identified as one of
The reduction profile of the Ce-doped sample (Ce/NiAl) was similar the fundamental reasons for catalyst deactivation in aqueous-phase re­
to that of the parent catalyst (NiAl) with a subtle shift of the peak-β to actions [61]. Therefore, spent catalysts (tested in glycerol APR at 235 ºC
lower temperatures, in line with others [57]. Taking as a reference the and 35 bar for 3 h) were subjected to TPR analysis (TPRspent) and the
bulk ceria reduction profile, there was an absence of separated, clear evolved gases were analysed by mass spectrometry. Two main hydrogen
reduction peaks of isolated CeO2 species (which presence was confirmed consumption events could be identified (Fig. S5, Supporting Informa­
by XRD) indicating they were masked by the intense peaks β and γ. tion): first, in the low-intermediate temperature range (200–600 ºC) and
Contrarily, catalyst Mg/NiAl presented a notably different reduction second, at high temperatures (> 600 ºC). The low temperature event was
profile. This case, peak-β and peak-γ could be resolved from peak-γ since accompanied by the evolution of methane and water. The release of
the latter appeared shifted to a higher temperature (peak maximum at methane suggested that reduction of nickel species took place simulta­
845 ºC). This behaviour was ascribed to the lesser reducibility of the neously with the hydrogenation of carbonaceous material, likely accu­
spinel phase Ni(Mg)Al2O4 (detected by XRD) caused by a strong mulated into the catalyst pores (glycerol and its APR intermediate liquid
MgO-NiO interaction [58,59]. Instead, peak-β, centred at around 690 ºC, products) [61]. The fact that no methane was detected during the high
remained unaltered in both the position and intensity. temperature hydrogen consumption event, suggested that solely the
As previously noted, around 40% of the Ni atoms in Ni/Al assay were reduction of the spinel was taking place. Once subtracted the hydrogen
incorporated to the spinel, and increased to above 90% for the copre­ consumption required for methane formation (assuming 2 moles of H2
cipitated samples. This was a determinant factor in stabilizing the Ni0 per mole of methane) the hydrogen uptake ascribed to reduction of the
particles since the NiAl2O4 structure suppressed their coarsening upon NiO particles (200–600 ºC range) was calculated. Data shown in Table 3
reduction [47]. This was evidenced by the four times smaller Ni0 size in suggested that a significant amount of metallic Ni was oxidized (i.e.
the NiAl series as compared to Ni/Al (Table 4). These smaller particles 37–51%) during the APR process. However, bulk metallic Ni nano­
were much more active in the glycerol conversion by APR. particles were preserved, as revealed by XRD analyses. Thus, oxidation
Prior to the catalytic runs, the catalysts were reduced at 700 ºC, for seemed to be restricted to the outmost layer of the Ni nanoparticle. Our
1 h. In order to measure the reduction degree of Ni (RDNi) and the results showed that NiAl was slightly more resistant to oxidation than
amount of metallic Ni out of the spinel lattice the so-called TPR700 and the impregnated counterpart (i.e. 37% vs 41%). Also, Ce doping seemed
TPR700/950 procedures were used. TPR700 profiles, shown in Fig. S4 to limit the oxidation of nickel (35% nickel oxidized) while Mg doping
(Supporting Information), coincided with the first part of the TPR950 caused the opposite behaviour.
curve (up to 700 ºC). Peak-α was observable in all samples; peak-β was Metal leaching in aqueous-phase systems proceeds by oxidation, and
partly developed, and peak-γ was absent in all samples. Overall, RDNi subsequent hydration to form hydroxide [62]. Indeed, particle size could
varied in the 57–89% (Table 5), being the highest for Ni/Al and Ce/NiAl be involved as suggested by the observed trend in the leached nickel and
and lowest for Mg/NiAl. the average Ni0 particle size in the fresh catalyst (Fig. S6, Supporting
The TPR700/950 reduction profiles showed a single reduction peak, at Information). It has been reported that smaller Ni nanoparticles are
above 700 ºC. For sample Ni/Al, it was centred at 816 ºC and down­ more susceptible to oxidation and leaching [63]. This behaviour would
shifted in the case of NiAl (780 ºC), indicating that nickel species were explain the increased average size of Ni0 in the spent Ce/NiAl (Table 2),
more readily reduced in the co-precipitated assay. The further downshift the catalyst with the highest Ni leaching. Leaching data showed some
for sample Ce/NiAl (770 ºC) revealed the promotion effect of Ce. On the interesting results. Although Mg-doped assay suffered from oxidation,

7
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

the leaching of Ni was lowered (i.e. from 0.7% to 0.4%, in NiAl and precursor limited the coalescence of smallest metallic Ni particles during
Mg/NiAl, respectively). It seemed that Mg played a kind of protection the reduction stage, likely due to the strong interaction of Ni with Al-rich
role in the leaching of Ni particles. environment (Ni-O-Al). Such interaction stabilized the metallic particles
by limiting their surface mobility and thus, by protecting against sin­
3.3.4. Nickel particle size distribution and dispersion tering under reductive atmosphere.
The particles morphology was studied by TEM (Fig. 4). The obtained In the case of the promoted catalysts, EDS mapping showed the
images revealed that unpromoted catalysts contained agglomerates of formation of nearly spherical-shape metallic Ni particles. The size dis­
nickel particles of nearly spherical-shape. A homogeneous particle size tribution in catalyst Ce/NiAl presented a bimodal distribution with
distribution was observed in both unpromoted samples, being narrower maxima at 10 and 15 nm. This catalyst had the largest amount of surface
for the coprecipitated assay. The average metallic Ni particle size of NiO particles (peak-α in TPR) within the spinel-derived series (Table 3:
reduced samples (Table 4) and indicated, in general, that Ni0 particles 8% of the total nickel). The fact that α-NiO crystallites are expected to
shown by TEM consisted of metallic crystallites agglomerates. produce, upon reduction, larger metallic Ni particles than those from
Among the unpromoted samples, Ni/Al contained metallic Ni parti­ spinel, would explain the observed bimodal distribution. In line with the
cles of a wide range of sizes, from around 7–70 nm; contrarily, the size above mentioned, catalyst Mg/NiAl, with the least amount of free NiO,
distribution was much narrower in the coprecipitated assay, with an showed the smaller size of Ni0 particles (10.6 nm) with a homogenous
average particle of NiAl was around 12 nm, which is four times smaller distribution. Regarding SNi◦ , Ce/NiAl almost equalled the parent NiAl
than the average value measured for Ni/Al (48 nm). Accordingly, both while Mg/NiAl had notable lesser value. The lower specific surface area
the metal dispersion (10% vs 1.7%) and surface area values (14.1 vs (SBET) and reducibility (DRNi) of Mg-doped catalyst were important
1.3 m2Niº/g) were significantly larger for the coprecipitated assay as constraints affecting its catalytic behaviour.
compared to the impregnated one.
These results suggested that the use of nickel aluminate spinel as

Fig. 4. STEM, EDS mapping images and Ni particle size distribution of reduced fresh catalysts: (a) Ni/Al; (b) NiAl; (c) Ce/NiAl and (d) Mg/NiAl.

8
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

3.3.5. Surface chemistry: basic and acid sites density would explain the obtained liquid product distribution, which was
Surface acid-base properties of the catalysts reduced at 700 ºC were mainly composed by dehydration products, with a negligible hydrogen
studied by means of NH3 and CO2-TPD techniques (Figs. S7 and S8, yield.
Supporting Information). The site densities are listed in Table 4. Catalyst It was observed that Ce doping improved the APR activity, whereas it
Ni/Al had a relatively low density of basic sites, of weak and medium was slight decreased by Mg doping. The hydrogen yield and selectivity
strength. Catalyst NiAl contained a much larger amount of basic sites were slightly increased by both promoters, whereas no effect on
(1.12 vs. 0.46 μmolCO2/g) with an appreciable contribution of those of methane selectivity was reported. XRD and TEM measurements
higher strength (14.4%). The density of basic sites increased for both concluded that smaller metallic Ni particles were formed upon reduction
doped samples due to the basic character of both promoters, being the of both promoted catalysts (around half of those in NiAl) and, conse­
increase larger for Mg/NiAl sample. The most remarkable difference quently, SNi◦ , that is, the available metallic sites, increased by 15–36%.
with respect to the parent NiAl was the increase of the medium and Others also found that magnesium, deposited by impregnation on nickel
strong basic sites (71% for NiAl vs. 80–82% for the promoted assays) aluminate spinel, could help to prevent the sintering of the metallic
(Fig. S7b, Supporting Information). nickel particles. The reducibility of the spinel based catalysts has been
Regarding acidity, unpromoted samples showed a similar surface found to increase upon Ce doping, and nickel species in the spinel phase,
acidity, what suggested that the exposed alumina, was the main source could be reduced at 700 ºC. Contrarily, the formation of magnesium-
of acidity. The synthesis method caused a trade-off between medium and nickel aluminate spinel, which is a difficult-to-reduce species, led to a
weak acid sites (Fig. S8b, Supporting Information). Promotion by Ce or lower degree of reduction of the Mg/NiAl catalyst. This case, the nickel
Mg also increased the acid-site density by around 22% and 30%, species in the spinel phase could not be reduced. Doping of either Ce or
respectively. It is interesting to note that Ce doping had no effect on the Mg increased the density of basic and acid functionalities on the cata­
distribution of the acid-sites, whereas incorporation of Mg increased the lysts surface, with a more marked increase of basic sites [67]. Moreover,
amount of strong sites. It is worth to note that acid sites were of Lewis- the strength of the acid and basic sites was increased.
type, consistent with the obtained liquid products distribution. The presence of CO2 and CH4 in the gas stream during glycerol APR
In summary, all of our catalysts presented amphoteric surface indicated that the investigated catalysts were active for the C–C bond
properties. High density of basic-sites (i.e. promotion of WGS reaction) cleavage. According to the reaction pathway, CO2 formation requires a
and low density of acid-sites (i.e. inhibition of dehydration reactions) first dehydrogenation step, to form glyceraldehyde, followed by decar­
have been related to improved hydrogen selectivity [27,31]. Among the bonylation and further conversion of the released CO by WGS. There­
unpromoted samples, Ni/Al had the lowest basic-to-acid-sites ratio fore, 2 moles of H2 are formed for each mole of CO2 produced. However,
(RBasic/Acid = 0.25 and 0.57 for Ni/Al and NiAl, respectively). Moreover, the production of CH4 implies hydrogen consumption. Methane could be
the incorporation of Ce and Mg notably increased the prevalence of basic produced by either hydrogenation of the released CO or by dehydration
sites (i.e. RBasic/Acid increased by 44% and 79%, respectively) (Table 4). of intermediate liquid products (i.e. ethylene glycol) and subsequent
Indeed, this could be related to the observed higher selectivity to C–C bond cleavage and hydrogenation. The CO2/CH4 mole ratio in the
hydrogen. gaseous products by the spinel-derived catalysts was three-fold larger
than Ni/Al (0.55), indicating the higher H2 selectivity of the formers,
3.4. General discussion likely due to a lowered basic to acid-sites ratio and the less accessible
metallic Ni surface. Among the promoted catalysts, the superior basicity
According to the mechanism previously described for glycerol of Mg/NiAl led to the highest CO2/CH4 ratio (1.63). Moreover, metallic
Aqueous-Phase Reforming over Ni-based catalysts, it is recognized that nickel dispersion and surface area values were significantly larger for
both the metal and the support play an essential role in the catalytic the coprecipitated assay as compared to the impregnated one (DNi0: 6%
behaviour [16]. The product-yield can be tailored by the fine-tuning of vs 1% and SNi0: 8.0 vs 0.8 m2Niº/g). This behaviour would explain the
both functionalities [27]. On one hand, the surface acidity directly af­ improved hydrogenation activity of the spinel-derived catalysts, as
fects the activity since glycerol could adsorb on the acid-sites to form pointed out by the composition of the liquid products, as well as their
hydroxyacetone through C–O bond cleavage (dehydration reaction) enhanced APR activity.
[63]. The surface basic sites would promote the desired WGS reaction
[31]. On the other hand, metallic Ni can activate the organic molecule 4. Conclusions
through the C–C bond cleavage to form ethylene glycol. This reaction
route releases one mole of CO and H2 per mole of glycerol converted. Four nickel-based catalysts were synthesized and characterized.
Indeed, the selectivity towards hydrogen will depend whether the WGS Their performance in the aqueous-phase reforming (APR) of glycerol
(beneficial) or the CO and CO2 hydrogenation (detrimental) reactions is was evaluated at 235 ºC and 35 bar, in a fixed-bed up-flow reactor. One
promoted by the metallic sites. of the catalysts was synthesized by impregnation of nickel over
According to activity results, NiAl catalyst, derived from nickel γ-alumina (Ni/Al), another by co-precipitation to form stoichiometric
aluminate spinel, was more advantageous than that prepared by the spinel (NiAl) and the rest by Ce and Mg doping on NiAl support (Ce/NiAl
conventional impregnation procedure (Ni/Al). The latter contained and Mg/NiAl). The characterization results indicated that nickel
coarse metallic Ni particles and, correspondingly, very low SNi◦ (ten-fold strongly anchored to the NiAl2O4 structure what improved the nickel
decrease as compared to NiAl). Indeed, the synthesis procedure notably dispersion upon reduction. This was reflected by the superior catalytic
affected the catalysts physico-chemical properties, and therefore, APR activity of NiAl with respect to Ni/Al, in terms of glycerol conversion
activity. It could be deduced that small metallic Ni particles are crucial and hydrogen selectivity. Addition of Mg and Ce limited the growth of
to achieve highly active catalysts for the APR; however, small nickel metallic Ni particles upon reduction and increased the basicity. In
particles can also activate CO to drive hydrogenation reactions [64], general, Ce promoted the reducibility of the nickel species while Mg
which would reduce selectivity to hydrogen [65]. The observed addition caused the opposite effect due to Ni-Mg interactions. Both
behaviour is in contrast to the reported activity vs. metal particle trend promoters slightly increased the catalyst selectivity to hydrogen. The
for Pt/Al2O3 catalysts, where the cleavage of C–C bonds preferably main difference was reported in the glycerol conversion. It notably
occurred on the facet Pt atoms rather than on edge and corner atoms (i. increased after Ce doping whereas it was slightly lowered after Mg
e., larger Pt particles were more active) [66]. In addition, the observed addition. Also, the promoted catalysts achieved a lower CO/CO2 and
difference in the acid/basic functionalities might be involved in the CO/H2 ratio in the gas-phase products, indicating their ability to pro­
divergent activity and product distribution of the unpromoted catalysts. mote WGS reaction. Regarding the obtained liquid-phase products,
The prevalence of acid sites in catalyst Ni/Al (Rbasic/acid was half of NiAl) spinel-based catalysts showed a similar product distribution, being

9
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

propylene glycol the most abundant product. [14] F. Liu, C. Okolie, R.M. Ravenelle, J.C. Crittenden, C. Sievers, P.C.A. Bruijnincx, B.
M. Weckhuysen, Silica deposition as an approach for improving the hydrothermal
XRD of the spent Ni/Al catalysts indicated the formation of
stability of an alumina support during glycerol aqueous phase reforming, Appl.
boehmite. Also, it was found that both promoters improved the stability Catal. A:Gen. 551 (2018) 13–22, https://doi.org/10.1016/j.apcata.2017.11.025.
of the textural properties under APR conditions. Leaching of metals is [15] M. El Doukkali, A. Iriondo, I. Gandarias, Enhanced catalytic upgrading of glycerol
likely the main issue to be addressed in future investigations in order to into high value-added H2 and propanediols: Recent developments and future
perspectives, Mol. Cat. 490 (2020), 110928, https://doi.org/10.1016/j.
seek for stable catalysts that allow liquid-phase biomass upgrading. mcat.2020.110928.
[16] A. Morales-Marín, J.L. Ayastuy, U. Iriarte-Velasco, M.A. Gutiérrez-Ortiz, Nickel
CRediT authorship contribution statement aluminate spinel-derived catalysts for the aqueous phase reforming of glycerol:
effect of reduction temperature, Appl. Catal. B:Environ. 244 (2019) 931–945,
https://doi.org/10.1016/j.apcatb.2018.12.020.
A. Morales-Marín: Investigation and Writing – original draft. J.L. [17] D. Li, Y. Li, X. Liu, Y. Guo, C.W. Pao, J.L. Chen, Y. Hu, Y. Wang, NiAl2O4 spinel
Ayastuy: Funding acquisition, Conceptualization, Writing – review & supported Pt catalyst: high performance and origin in aqueous-phase reforming of
methanol, ACS Catal. 9 (2019) 9671–9682, https://doi.org/10.1021/
editing. U. Iriarte-Velasco: Formal analysis, Writing – review & editing. acscatal.9b02243.
M.A. Gutiérrez-Ortiz: Resources, Funding acquisition and supervision. [18] P.D. Vaidya, J.A. Lopez-Sanchez, Review of hydrogen production by catalytic
aqueous-phase reforming, Chem. Sel. 2 (2017) 6563–6576, https://doi.org/
10.1002/slct.201700905.
Declaration of Competing Interest [19] A. Arandia, I. Coronado, A. Remiro, A.G. Gayubo, M. Reinikainen, Aqueous-phase
reforming of bio-oil aqueous fraction over nickel-based catalysts, Int. J. Hydrog.
The authors declare that they have no known competing financial Energ. 44 (2019) 13157–13168, https://doi.org/10.1016/j.ijhydene.2019.04.007.
[20] A. Iriondo, V.L. Barrio, J.F. Cambra, P.L. Arias, M.B. Güemez, R.M. Navarro, M.
interests or personal relationships that could have appeared to influence C. Sánchez-Sánchez, J.L.G. Fierro, Hydrogen production from glycerol over nickel
the work reported in this paper. catalysts supported on Al2O3 modified by Mg, Zr, Ce or La, Top. Catal. 49 (2008)
46–58, https://doi.org/10.1007/s11244-008-9060-9.
[21] Z. Boukha, C. Jiménez-González, M. Gil-Calvo, B. de Rivas, J.R. González-Velasco,
Acknowledgments J.I. Gutiérrez-Ortiz, R. López-Fonseca, MgO/NiAl2O4 as a new formulation of
reforming catalysts: tuning the surface properties for the enhanced partial
This research was funded by Mineco (PID2019–106692EB-I00) and oxidation of methane, Appl. Catal. B:Environ. 199 (2016) 372–383, https://doi.
org/10.1016/j.apcatb.2016.06.045.
FEDER. The authors thank for technical support provided by SGIker of [22] H. Lee, G.S. Shin, Y. Kim, Characterization of supported Ni catalysts for aqueous-
UPV/EHU and European funding (ERDF and ESF). phase reforming of glycerol, Korean J. Chem. Eng. 32 (2015) 1267–1272, https://
doi.org/10.1007/s11814-014-0325-7.
[23] T.K. Phung, T.L.M. Pham, A.T. Nguyen, K.B. Vu, H.N. Giang, T.A. Nguyen, T.
Appendix A. Supporting information C. Huynh, H.D. Pham, Effect of supports and promoters on the performance of ni-
based catalysts in ethanol steam reforming, Chem. Eng. Technol. 43 (2020)
Supplementary data associated with this article can be found in the 672–688, https://doi.org/10.1002/ceat.201900445.
[24] L. Smoláková, M. Kout, E. Koudelková, L. Ĉapek, Effect of calcination temperature
online version at doi:10.1016/j.jece.2021.106433. on the structure and catalytic performance of the Ni/Al2O3 and Ni-Ce/Al2O3
catalysts in oxidative dehydrogenation of ethane, Ind. Eng. Chem. Res. 54 (2015)
References 12730–12740, https://doi.org/10.1021/acs.iecr.5b03425.
[25] A. Borodzinski, M. Bonarowska, Relation between crystallite size and dispersion on
supported metal catalysts, Langmuir 13 (1997) 5613–5620, https://doi.org/
[1] L. Chen, Z. Qi, S. Zhang, J. Su, G.A. Somorjai, Catalytic hydrogen production from
10.1021/la962103u.
methane: a review on recent progress and prospect, Catalysts 10 (8) (2020) 858,
[26] J. Sehested, S. Dahl, J. Jacobsen, J.R. Rostrup-Nielsen, Methanation of CO over
https://doi.org/10.3390/catal10080858.
nickel: mechanism and kinetics at high H2/CO ratios, J. Phys. Chem. B 109 (2005)
[2] Y.K. Salkuyeh, B.A. Saville, H.L. MacLean, Techno-economic analysis and life cycle
2432–2438, https://doi.org/10.1021/jp040239s.
assessment of hydrogen production from natural gas using current and emerging
[27] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, A review of
technologies, Int. J. Hydrog. Energ. 42 (30) (2017) 18894–18909, https://doi.org/
catalytic issues and process conditions for renewable hydrogen and alkanes by
10.1016/j.ijhydene.2017.05.219.
aqueous-phase reforming of oxygenated hydrocarbons over supported metal
[3] R.D. Cortright, R.R. Davda, J.A. Dumesic, Hydrogen from catalytic reforming of
catalysts, Appl. Catal. B:Environ. 56 (2005) 171–186, https://doi.org/10.1016/j.
biomass-derived hydrocarbons in liquid water, Nature 418 (2002) 964–967,
apcatb.2004.04.027.
https://doi.org/10.1038/nature01009.
[28] L. Shen, J. Xu, M. Zhu, Y.F. Han, Essential role of the support for nickel-based CO2
[4] M. Alvear, A. Aho, I.L. Simakova, H. Grénman, T. Salmi, D.Y. Murzin, Aqueous
methanation catalysts, ACS Catal. 10 (2020) 14581–14591, https://doi.org/
phase reforming of alcohols over a bimetallic Pt-Pd catalyst in the presence of
10.1021/acscatal.0c03471.
formic acid, Chem. Eng. J. 398 (2020), 125541, https://doi.org/10.1016/j.
[29] Y. Han, Y. Quan, J. Zhao, J. Ren, Promotion effect by Mg on the catalytic behavior
cej.2020.125541.
of MgNi/WO3 in the CO methanation, Int J. Hydrog. Energ. 45 (2020)
[5] J. Tao, L. Hou, B. Yan, G. Chen, W. Li, H. Chen, Z. Cheng, F. Lin, Hydrogen
29917–29928, https://doi.org/10.1016/j.ijhydene.2020.08.121.
production via aqueous-phase reforming of ethylene glycol over a nickel-iron alloy
[30] L. Znak, K. Stołecki, J. Zieliński, The effect of cerium, lanthanum and zirconium on
catalyst: effect of cobalt addition, Energy Fuels 34 (2020) 1153–1161, https://doi.
nickel/alumina catalysts for the hydrogenation of carbon oxides, Catal. Today 101
org/10.1021/acs.energyfuels.9b02149.
(2) (2005) 65–71, https://doi.org/10.1016/j.cattod.2005.01.003.
[6] T. Xie, C.J. Bodenschatz, R.B. Getman, Insights into the roles of water on the
[31] Y. Guo, M.U. Azmat, X. Liu, Y. Wang, G. Lu, Effect of support’s basic properties on
aqueous phase reforming of glycerol, React. Chem. Eng. 4 (2019) 383–392,
hydrogen production in aqueous-phase reforming of glycerol and correlation
https://doi.org/10.1039/C8RE00267C.
between WGS and APR, Appl. Energy 92 (2012) 218–223, https://doi.org/
[7] A.V. Kirilin, B. Hasse, A.V. Tokarev, L.M. Kustov, G.N. Baeva, G.O. Bragina, A.
10.1016/j.apenergy.2011.10.020.
Y. Stakheev, A.R. Rautio, T. Salmi, B.J.M. Etzold, J.P. Mikkola, D.Y. Murzin,
[32] S. Jin, Y. Park, G. Bang, N. Dat Vo, C.H. Lee, Revisiting magnesium oxide to boost
Aqueous-phase reforming of xylitol over Pt/C and Pt/TiC-CDC catalysts: catalyst
hydrogen production via water-gas shift reaction: mechanistic study to economic
characterization and catalytic performance, Catal. Sci. Technol. 4 (2014) 387–401,
evaluation, Appl. Catal. B:Environ 284 (2021), 119701.
https://doi.org/10.1039/C3CY00636K.
[33] S.D. Senanayake, J. Evans, S. Agnoli, L. Barrio, T.L. Chen, J. Hrbek, J.A. Rodriguez,
[8] A. Kirilin, J. Wärnå, A. Tokarev, D.Y. Murzin, Kinetic modeling of sorbitol aqueous-
Water-gas shift and CO methanation reactions over Ni-CeO2(111) catalysts, Top.
phase reforming over Pt/Al2O3, Ind. Eng. Chem. Res. 53 (2014) 4580–4588,
Catal. 54 (2011) 34–41, https://doi.org/10.1007/s11244-011-9645-6.
https://doi.org/10.1021/ie403813y.
[34] J. Gao, Q. Liu, F. Gu, B. Liu, Z. Zhong, F. Su, Recent advances in methanation
[9] R.R. Davda, J.A. Dumesic, Renewable hydrogen by aqueous-phase reforming of
catalysts for the production of synthetic natural gas, RSC Adv. 5 (2015)
glucose, Chem. Commun. 4 (2004) 36–37, https://doi.org/10.1039/B310152E.
22759–22776, https://doi.org/10.1039/C4RA16114A.
[10] R.A. Sheldon, The E factor: fifteen years on, Green. Chem. 9 (2007) 1273–1283,
[35] J.W. Shabaker, R.R. Davda, G.W. Huber, R.D. Cortright, J.A. Dumesic, Aqueous-
https://doi.org/10.1039/B713736M.
phase reforming of methanol and ethylene glycol over alumina-supported platinum
[11] J.J. Bozell, G.R. Petersen, Technology development for the production of biobased
catalysts, J. Catal. 215 (2) (2003) 344–352, https://doi.org/10.1016/S0021-9517
products from biorefinery carbohydrates -the US Department of Energy’s “top 10”
(03)00032-0.
revisited, Green. Chem. 12 (2010) 539–554, https://doi.org/10.1039/B922014C.
[36] L.I. Godina, A.V. Tokarev, I.L. Simakova, P.M. Arvela, E. Kortesmäki, J. Gläsel,
[12] Q.S. He, J. McNutt, J. Yang, Utilization of the residual glycerol from biodiesel
L. Kronberg, B. Etzold, D.Y. Murzin, Aqueous-phase reforming of alcohols with
production for renewable energy generation, Renew. Sustain. Energy Rev. 71
three carbon atoms on carbon-supported Pt, Catal. Today 301 (2018) 78–89,
(2017) 63–76, https://doi.org/10.1016/j.rser.2016.12.110.
https://doi.org/10.1016/j.cattod.2017.03.042.
[13] A. Ciftci, D.A.J.M. Ligthart, A.O. Sen, A.J.F. Van Hoof, H. Friedrich, E.J.M. Hensen,
[37] B. Liu, J. Greeley, Decomposition pathways of glycerol via C-H, O-H, and C-C bond
Pt-Re synergy in aqueous-phase reforming of glycerol and the water-gas shift
scission on Pt(111): a density functional theory study, J. Phys. Chem. C. 115 (2011)
reaction, J. Catal. 311 (2014) 88–101, https://doi.org/10.1016/j.
19702–19709, https://doi.org/10.1021/jp202923w.
jcat.2013.11.011.

10
A. Morales-Marín et al. Journal of Environmental Chemical Engineering 9 (2021) 106433

[38] P.J. Dietrich, T. Wu, A. Sumer, J.A. Dumesic, J. Jellinek, W.N. Delgass, F. Catal. B:Environ. 220 (2018) 31–41, https://doi.org/10.1016/j.
H. Ribeiro, J.T. Miller, Aqueous phase glycerol reforming with Pt and PtMo apcatb.2017.08.030.
bimetallic nanoparticle catalysts: the role of the Mo promoter, Top. Catal. 56 [54] M. Stekrova, A. Rinta-Paavola, R. Karinen, Hydrogen production via aqueous-
(2013) 1814–1828, https://doi.org/10.1007/s11244-013-0115-1. phase reforming of methanol over nickel modified Ce, Zr and La oxide supports,
[39] D. Sun, Y. Yamada, S. Sato, W. Ueda, Glycerol hydrogenolysis into useful C3 Catal. Today 304 (2018) 143–152, https://doi.org/10.1016/j.cattod.2017.08.030.
chemicals, Appl. Catal. B:Environ. 193 (2016) 75–92, https://doi.org/10.1016/j. [55] G. Li, L. Hu, J.M. Hill, Comparison of reducibility and stability of alumina-
apcatb.2016.04.013. supported Ni catalysts prepared by impregnation and co-precipitation, Appl. Catal.
[40] G. Balducci, M.S. Islam, J. Kašpar, P. Fornasiero, M. Graziani, Bulk reduction and A:Gen. 301 (2006) 16–24, https://doi.org/10.1016/j.apcata.2005.11.013.
oxygen migration in the ceria-based oxides, Chem. Mater. 12 (2000) 677–681, [56] A. Horváth, M. Németh, A. Beck, B. Maróti, G. Sáfrán, G. Pantaleo, L.F. Liotta, A.
https://doi.org/10.1021/cm991089e. M. Venezia, V. La Parola, Strong impact of indium promoter on Ni/Al2O3 and Ni/
[41] R.M. Ravenelle, J.R. Copeland, W. Kim, J.C. Crittenden, C. Sievers, Structural CeO2-Al2O3 catalysts used in dry reforming of methane, Appl. Catal. A: Gen. 621
changes of γ -Al2O3-supported catalysts in hot liquid water, ACS Catal. 1 (2011) (2021), 118174.
552–561, https://doi.org/10.1021/cs1001515. [57] Y. Zhan, J. Han, Z. Bao, B. Cao, Y. Li, J. Street, F. Yu, Biogas reforming of carbon
[42] D.J.M. de Vlieger, B.L. Mojet, L. Lefferts, K. Seshan, Aqueous phase reforming of dioxide to syngas production over Ni-Mg-Al catalysts, Mol. Cat. 436 (2017)
ethylene glycol -role of intermediates in catalyst performance, J. Catal. 292 (2012) 248–258, https://doi.org/10.1016/j.mcat.2017.04.032.
239–245, https://doi.org/10.1016/j.jcat.2012.05.019. [58] H. Özdemir, M.A.F. Öksüzömer, Synthesis of Al2O3, MgO and MgAl2O4 by solution
[43] S. Biswas, A. Pal, T. Pal, Supported metal and metal oxide particles with proximity combustion method and investigation of performances in partial oxidation of
effect for catalysis, RSC Adv. 10 (2020) 35449–35472. methane, Powder Technol. 359 (2020) 107–117, https://doi.org/10.1016/j.
[44] Y.Y. Gorbanev, S. Kegnæs, A. Riisager, Selective aerobic oxidation of 5-hydroxy­ powtec.2019.10.001.
methylfurfural in water over solid ruthenium hydroxide catalysts with magnesium- [59] A. Parmaliana, F. Arena, F. Frusteri, N. Giordano, Temperature-programmed
based supports, Catal. Lett. 141 (2011) 1752–1760, https://doi.org/10.1007/ reduction study of NiO-MgO interactions in magnesia-supported Ni catalysts and
s10562-011-0707-y. NiO-MgO physical mixture, J. Chem. Soc. Faraday Trans. 86 (1990) 2663–2669,
[45] C. Sievers, S.L. Scott, Y. Noda, L. Qi, E.M. Albuquerque, R.M. Rioux, Phenomena https://doi.org/10.1039/FT9908602663.
affecting catalytic reactions at solid-liquid interfaces, ACS Catal. 6 (2016) [60] J.W. Shabaker, D.A. Simonetti, R.D. Cortright, J.A. Dumesic, Sn-modified Ni
8286–8307, https://doi.org/10.1021/acscatal.6b02532. catalysts for aqueous-phase reforming: characterization and deactivation studies,
[46] L. Zhou, L. Li, N. Wei, J. Li, J. Basset, Effect of NiAl2O4 formation on Ni/Al2O3 J. Catal. 231 (2005) 67–76, https://doi.org/10.1016/j.jcat.2005.01.019.
stability during dry reforming of methane, ChemCatChem 7 (2015) 2508–2516, [61] M. El Doukkali, A. Iriondo, J.F. Cambra, I. Gandarias, L. Jalowiecki-Duhamel,
https://doi.org/10.1002/cctc.201500379. F. Dumeignil, P.L. Arias, Deactivation study of the Pt and/or Ni-based γ-Al2O3
[47] N. Braidy, S. Bastien, J. Blanchard, C. Fauteux-Lefebvre, I.E. Achouri, catalysts used in the aqueous phase reforming of glycerol for H2 production, Appl.
N. Abatzoglou, Activation mechanism and microstructural evolution of a YSZ/Ni- Catal. A:Gen. 472 (2014) 80–91, https://doi.org/10.1016/j.apcata.2013.12.015.
alumina catalyst for dry reforming of methane, Catal. Today 291 (2017) 99–105, [62] T. Van Haasterecht, M. Swart, K.P. De Jong, J.H. Bitter, Effect of initial nickel
https://doi.org/10.1016/j.cattod.2017.03.006. particle size on stability of nickel catalysts for aqueous phase reforming, J. Energy
[48] W. Gac, Acid-base properties of Ni-MgO-Al2O3 materials, Appl. Surf. Sci. 257 Chem. 25 (2016) 289–296, https://doi.org/10.1016/j.jechem.2016.01.006.
(2011) 2875–2880, https://doi.org/10.1016/j.apsusc.2010.10.084. [63] J. Callison, N.D. Subramanian, S.M. Rogers, A. Chutia, D. Gianolio, C.R.A. Catlow,
[49] B. Li, X. Xu, S. Zhang, Synthesis gas production in the combined CO2 reforming P.P. Wells, N. Dimitratos, Directed aqueous-phase reforming of glycerol through
with partial oxidation of methane over Ce-promoted Ni/SiO2 catalysts, Int. J. tailored platinum nanoparticles, Appl. Catal. B:Environ. 238 (2018) 618–628,
Hydrog. Energ. 38 (2013) 890–900, https://doi.org/10.1016/j. https://doi.org/10.1016/j.apcatb.2018.07.008.
ijhydene.2012.10.103. [64] J. Gao, C. Jia, M. ZHang, F. Gu, G. Xu, F. Su, Effect of nickel nanoparticle size in
[50] J.P.D.S.Q. Menezes, A.P.D.S. Dias, M.A.P. da Silva, M.M.V.M. Souza, Effect of Ni/γ-Al2O3 on CO methanation reaction for the production of syntetic natural gas,
alkaline earth oxides on nickel catalysts supported over γ-alumina for butanol Catal. Sci. Technol. 8 (2013) 2009–2015, https://doi.org/10.1039/C3CY00139C.
steam reforming: coke formation and deactivation process, Int. J. Hydrog. Energ. [65] B. Roy, U. Martinez, K. Loganathan, A.K. Daty, C.A. Leclerc, Effect of preparation
45 (2020) 22906–22920, https://doi.org/10.1016/j.ijhydene.2020.06.187. methods on the performance of Ni/Al2O3 catalysts for aqueous-phase reforming of
[51] J.A. Abi, P. Courty, D. Decottignies, M. Michau, F. Diehl, X. Carrier, E. Marceau, ethanol: part I-catalytic activity, Int. J. Hydrog. Energ. 37 (10) (2012) 8143–8153,
Inhibition by inorganic dopants of γ-alumina chemical weathering under https://doi.org/10.1016/j.ijhydene.2012.02.056.
hydrothermal conditions: identification of reactive sites and their influence in [66] K. Lehnert, P. Claus, Influence of Pt particle size and support type on the aqueous-
fischer-tropsch synthesis, ChemCatChem 9 (2017) 2106–2117, https://doi.org/ phase reforming of glycerol, Catal. Commun. 9 (15) (2008) 2543–2546, https://
10.1002/cctc.201700140. doi.org/10.1016/j.catcom.2008.07.002.
[52] N. Luo, X. Fu, F. Cao, T. Xiao, P.P. Edwards, Glycerol aqueous phase reforming for [67] F. Bastan, M. Kazemeini, A. Larimi, H. Maleki, Production of renewable hydrogen
hydrogen generation over Pt catalyst - effect of catalyst composition and reaction through aqueous-phase reforming of glycerol over Ni/Al2O3-MgO nano-catalyst,
conditions, Fuel 87 (2008) 3483–3489, https://doi.org/10.1016/j. Int. J. Hydrog. Energ. 43 (2018) 614–621, https://doi.org/10.1016/j.
fuel.2008.06.021. ijhydene.2017.11.12.
[53] I.C. Freitas, R.L. Manfro, M.M.V.M. Souza, Hydrogenolysis of glycerol to propylene
glycol in continuous system without hydrogen addition over Cu-Ni catalysts, Appl.

11

You might also like