Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Environmental Chemical Engineering 10 (2022) 107710

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Enhanced CH4 selectivity for CO2 methanation over Ni-TiO2 by addition of


Zr promoter
Ammarika Makdee a, Pinit Kidkhunthod b, Yingyot Poo-arporn b, Kingkaew
Chayakul Chanapattharapol a, *
a
Materials Chemistry Research Center, Department of Chemistry, Faculty of Science, Khon Kaen University, Khon Kaen 40002, Thailand
b
Synchrotron Light Research institute (Public Organization), Nakhon Ratchasima 30000, Thailand

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

Editor: Dong-Yeun Koh TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts were synthesized and used for CO2
methanation reaction. The catalyst with 10 mol% Zr addition (10 wt%Ni-Zr0.1Ti0.9O2) exhibited the highest CO2
Keywords: conversion and CH4 selectivity with highly stable property. The role of Zr addition on enhancing CO2 metha­
CO2 methanation nation reaction rate was to tune the electronic property of Ni species by electron transfer from Zr to Ni valence
Zr incorporation into TiO2
state. From the donating effect of Zr, an appropriate metal-support interaction was then improved which led to
Electron movement
higher dispersion of Ni species on catalyst surface which was beneficial to the adsorption of H2 molecules.
Metal-support interaction
Moreover, Zr addition also improved the basicity of the catalyst, which can also promote the adsorption of CO2.
Therefore, an appropriate H2 and CO2 adsorption ability can enhance catalytic activity. The effect of tuning
electronic properties by Zr addition on enhancing the catalytic performance resulted from lowering of C-O
bonding dissociation barrier. Upon CO intermediate was adsorbed on Ni electron rich site, the electron of valence
d state of Ni was transferred to π* anti-bonding of CO molecule and the C-O bonding was then weakened and
easily dissociated to carbon and oxygen to further hydrogenation and formed products.

1. Introduction according to Eq. (1), is more interesting for many research groups as it
does not only decrease CO2 concentration but also generates CH4, which
It is well known that the increase of greenhouse gas emissions, is an important energy carrier source [2]. The generated CH4 can be
especially carbon dioxide (CO2) which is the most abundant in the at­ directly injected into natural gas pipelines and used as a starting reagent
mosphere and the main contributor to global warming and climate for the formation of other chemicals. Furthermore, the CO2 methanation
change, is a global concern. Nowadays, the CO2 concentration in the reaction is also considerably faster than other reactions (CO2 hydroge­
atmosphere can be reduced by two main approaches; CO2 capture and nation to hydrocarbons or alcohols) [3] under atmospheric pressure and
storage (CCS) and CO2 capture and utilization (CCU). CO2 capture and low temperature.
storage is a process that reduces CO2 concentration by capturing and
CO2 + 4H2 ⇌ CH4 + 2H2O ΔH298 K = − 165.0 kJ/mol (1)
storing CO2 in the oceans or mineral carbonates. This involves long-term
separation from the atmosphere [1]. Although CO2 capture and storage Normally, to increase the reaction rate, increasing the reaction
are an important way to rapidly decrease CO2 in the atmosphere, it has a temperature is usually employed. However, CO2 methanation is an
problem of CO2 leakage, high operation costs and high energy demands. exothermic reaction which favors to occur at low temperature; thus,
Therefore, CO2 capture and utilization is now the preferred approach. increasing the reaction temperature leads to a reduction of product
CO2 capture and utilization is the technique that captures and uses CO2 formation. Moreover, the CO2 molecule with a fully oxidized state is
as a carbon feedstock to produce useful products and fuels such as car­ inert and highly stable; thus, high energy input is needed to activate this
bon monoxide (CO), methanol (CH3OH), dimethyl ether (CH3OCH3), molecule. Therefore, to overcome the kinetic and thermodynamic lim­
and methane (CH4). Among CO2 conversion processes, CO2 hydroge­ itations, a catalyst must be employed to get high reaction rate at low
nation to CH4, which is called CO2 methanation or Sabatier reaction temperature. The heterogeneous catalyst for this reaction typically

* Corresponding author.
E-mail address: kingkaew@kku.ac.th (K.C. Chanapattharapol).

https://doi.org/10.1016/j.jece.2022.107710
Received 17 January 2022; Received in revised form 17 March 2022; Accepted 11 April 2022
Available online 14 April 2022
2213-3437/© 2022 Elsevier Ltd. All rights reserved.
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

consists of two main parts: active species and support. For active species, emission scanning electron microscopy (FE-SEM), H2-temperature pro­
several works confirmed that noble metals such as Ru and Rh provided grammed reduction (H2-TPR) CO2-temperature programmed desorption
the highest catalytic activity and CH4 selectivity [3–8]. However, they (CO2-TPD), transmission electron microscopy (TEM) and X-ray absorp­
are very expensive. Therefore, other metals with a lower price have been tion spectroscopy (XAS). The catalytic activities of the prepared cata­
employed instead. Ni metal has been widely used as an active species lysts were tested for CO2 methanation reaction in the temperature range
due to its being inexpensive and having high catalytic activity [4,9–13]. of 100–500 ◦ C. Finally, the role of Zr addition on enhancing the CO2
Normally, Ni active species is usually dispersed on high surface area methanation activity of TiO2-supported Ni was discussed using all of
metal oxide support such as SiO2 Al2O3, zeolite and CeO2 etc. These these characterized results.
supports affect the metal-support interaction, CO2 adsorption, and metal
dispersion on the catalyst surface. Those properties are important factors 2. Experiment
to dominate the CO2 methanation catalytic performance. Among metal
oxide supports, TiO2 is now of interested support due to its low price, 2.1. Catalysts preparation
low toxicity, and good chemical and thermal stability.
In the past decade, many research groups reported on improving 2.1.1. Preparation of catalyst support
catalyst properties by many approaches in order to obtain the catalyst TiO2 support was prepared by the sol-gel method. Ti[OCH(CH3)2]4
with high catalytic performance, high stability which avoids deactiva­ was used as a Ti precursor. Ethanol and DI water were employed as
tion during time on stream. One of the most important issues concerning solvents and HNO3 was an acidic catalyst. Ti[OCH(CH3)2]4 was first
catalyst synthesis is that the obtained catalyst must express high CO2 dissolved in ethanol, then DI water and HNO3 were added to the Ti
conversion and CH4 selectivity. Since this reaction has a significant precursor solution and the mixture was continuously stirred for 3 h.
thermodynamic limitation, therefore, at high reaction temperature, CH4 After that, the mixture was dried at 100 ◦ C for 24 h and then calcined
formation is suppressed. Beside thermodynamic limitation at high re­ under air at 400 ◦ C (heating rate of 2 ◦ C/min) for 3 h.
action temperature, reversed water gas shift reaction (CO2 + H2 ⇌ CO + ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) mixed oxide supports were
H2O ΔH298 K = 41.2 kJ/mol), which is a side reaction, can also occur. prepared by using the same procedure as the preparation of TiO2 sup­
Because this side reaction favors to occur at high temperature, the un­ port. ZrO(NO3)2.xH2O was used as a Zr source. The calculated amount of
desired CO product is usually formed and leads to lower CH4 selectivity. Zr precursor was dissolved in DI water. Afterward, both Zr aqueous
Therefore, to increase the CH4 selectivity at high temperature, CO solution and HNO3 were added into the Ti precursor solution. Then, the
methanation is a pathway of interest. The produced CO from side re­ mixture was proceeded along with the preparation method as TiO2.
action can convert to CH4 via CO methanation reaction (CO + 3 H2 ⇌
CH4 + H2O; ΔH298 K = − 206 kJ/mol). In order to increase CH4 forma­ 2.1.2. Preparation of 10 wt%Ni-TiO2 and 10 wt%Ni-ZrxTi1− xO2 catalysts
tion through CO methanation, dissociation of CO molecules must be 10 wt%Ni-doped TiO2 and ZrxTi1− xO2 catalysts were prepared by
facilitated by lowering its dissociation barrier [14–16]. It has been re­ incipient wetness impregnation method. Ni(NO3)2⋅6H2O was used as a
ported that the addition of the second metal into the catalyst can reduce Ni precursor. The desired amount of the Ni precursor was dissolved by
the CO dissociation barrier which resulted in high CO2 conversion with DI water. The solution of the Ni precursor was added to the prepared
high CH4 selectivity at high reaction temperature. The CO methanation supports and then dried overnight at 100 ◦ C and calcined under air at
catalytic activity can be promoted by tuning the interaction between an 400 ◦ C (heating rate of 2 ◦ C/min) for 3 h.
active species and the support. Kikkawa et al. [14] reported Pt-Ni/­
γ-Al2O3 for selective hydrogenation of CO2 toward CH4. They found that 2.2. Catalysts characterization
Ni-Pt alloy can provide the active site for H2 dissociation at low tem­
perature, which resulted in high CO2 methanation catalytic activity. 2.2.1. Standard characterization
This can be described by tuning the interaction between active species X-ray diffraction spectra of all catalysts were obtained from PAN­
and support. The formation of Ni-Pt alloy can promote the C-O disso­ alytical Empyrean with Cu Kα radiation with the wavelength of 1.5406
ciation by increasing the bond strength between the active species on the Å. The XRD patterns were scanned in the 2θ range of 20–80◦ . The
catalyst surface and the CO molecule. Upon strongly anchoring the CO crystalline size (D) was calculated by using Scherrer’s equation by using
molecule on an active species, the C-O bonding was weakened, which (101) and (200) planes of TiO2 and NiO, respectively. Lattice parameters
led to more easily to dissociate to carbon and oxygen atom and further (a, b, c) were calculated from Bragg’s equation by using (101), (004),
hydrogenate with nearby H atom to form products. Therefore, the and (200) planes of TiO2.
addition of the second metal into the catalyst to tune the electronic N2 adsorption-desorption isotherms at 77 K of the prepared catalysts
property of Ni active species is an alternative method. Zr ion is one of the were obtained from the BELSORP-miniX instrument. Before the gas
transition metals that is widely used as the promoter for methanation adsorption experiment, the sample was degassed at 150 ◦ C for 60 min.
reaction [17–20]. It has been reported that the addition of Zr promoted The obtained isotherm was used to calculate the specific surface area of
the thermal stability and catalytic activity of TiO2 [21,22] and also the prepared catalysts using the multipoint-BET method.
improved the structural properties of TiO2 in terms of crystallinity and Three-dimensional morphologies of the prepared catalysts were ob­
surface area [23–25]. Moreover, Zr is an earth-abundant and environ­ tained from an FEI Helios Nanolab G3 CX Dual Beam FIB/SEM instru­
mentally friendly element. Li et al. [26] reported the effect of Zr addition ment. Before scanning, the catalyst was spread on a carbon tape which
on the catalytic activity of anatase and rutile TiO2-supported Co for CO2 was stuck on a specimen stub and coated with gold to avoid the charging
hydrogenation reaction. Zr addition improved the adsorptions of CO, effect.
CO2, and H2 in both the capacity and strength on the catalyst surface and Inductively coupled plasma-optical emission spectroscopy was used
then led to enhancing CO2 conversion. to determine the metal contents in the prepared catalysts. Before the
To the best of our knowledge, this is the first time to report on Zr measuring of metal content, the catalyst (0.02–0.025 g) was digested by
addition into TiO2 as a zirconia-titania mixed oxide support and which is the mixture of concentrated H2SO4 (5 mL) and 30% H2O2 (1 mL) at high
then impregnated by Ni to use as a catalyst for CO2 methanation. In this temperature. The standard solutions in the range of 0–10 ppm were used
work, the role of Zr addition on the properties and the catalytic activity to construct the calibration curve. All samples were analyzed by the
for the CO2 methanation of TiO2-supported Ni catalysts was investi­ optima 100DVICP-OES (PerkinElmer) instrument. The detection wave­
gated. The TiO2-supported Ni catalysts with different amounts of Zr lengths of Zr and Ni were 343.823 and 231.604 nm, respectively.
addition were prepared and then characterized using several techniques In order to study the Ni dispersion in the catalyst, transmission
including X-ray diffraction (XRD), N2 adsorption-desorption, field electron microscopy images of the reduced catalysts were obtained from

2
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Fig. 1. Temperature-dependent (a) CO2 conversion, (b) CH4 yield (c) CH4 and CO selectivity of TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts
for CO2 methanation (CO2:H2 ratio = 1:4) (d) Arrhenius plots of 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts and (e) Long-term stability test over 10 wt
%Ni-Zr0.1Ti0.9O2 catalyst at 450 ◦ C for 50 h (CO2:H2 ratio = 1:4).

a FEI/TECNAI G2 20 instrument. Before the scanning process, the obtained from BELCAT II Version 0.3.0.6 instrument. 0.05 g of the
reduced catalyst powder was dispersed in acetone by sonication and prepared catalyst was first pretreated by N2 gas with a flow rate of 50
then dropped on a copper grid. mL/min at 500 ◦ C for 30 min. After that, the mixed gas of H2 and N2 was
H2-temperature programmed reduction profiles of the catalysts were switched and flowed over the pretreated sample during increasing the

3
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

reduction temperature from 40 ◦ C to 800 ◦ C with a ramping rate of 50 ◦ C until it reached 500 ◦ C and the spectra were recorded as
10 ◦ C/min. H2 consumption was detected by a TCD detector and the TCD described above.
signal was plotted as a function of temperature.
CO2-temperature programmed desorption profiles of all prepared
catalysts were obtained by using the BELCAT II Version 0.3.0.6 instru­ 2.3. CO2 methanation catalytic activity test
ment. 0.05 g of the prepared catalyst was first pretreated by He gas with
a flow rate of 50 mL/min at 300 ◦ C for 60 min. After finishing the pre­ The catalytic activities of all prepared catalysts were tested for CO2
treatment process, the temperature was decreased to 50 ◦ C under He methanation reaction in a continuous flowed fixed-bed reactor. The
flow and then the gas flow was switched to CO2 with a flow rate of 30 catalyst (50 mg) was placed between two layers of quartz wool in the
mL/min for 30 min. After that, the excess CO2 and weakly chemisorbed Pyrex tube. The Pyrex tube was placed inside the controllable temper­
CO2 were removed by purging He (30 mL/min) for 60 min. The CO2-TPD ature furnace. The K-type thermocouple was placed at the top of the
was started by increasing temperature from 50 ◦ C to 750 ◦ C with a catalyst to detect the reaction temperature. The feed gases flow rate was
heating rate of 10 ◦ C/min under He flow. Finally, the desorbed CO2 was controlled by mass flow controllers. Before starting CO2 methanation,
monitored by a TCD detector and the TCD signal was plot as a function of the catalyst was reduced at 450 ◦ C under H2 with a flow rate of 50 mL/
temperature. The amount of desorbed CO2 (mmol/gcat.) was calculated min for 90 min. After that, the feed gas was changed to the mixture of
using the peak area. CO2, H2, and He with a flow rate of 6, 24, and 10 mL/min, respectively.
The mixed feed gas was passed through the catalyst under a reaction
2.2.2. X-ray absorption spectroscopy (XAS) temperature of 100–500 ◦ C. Finally, the remaining reactants and prod­
X-ray absorption spectroscopy was studied on a SUT-NANOTEC-SLRI ucts were analysed by online Agilent Technologies 6890 N gas chro­
beamline (BL 5.2) [27,28] and a Time-resolved X-ray absorption spec­ matograph with a HEYSEP D packed column and TCD was used as a
troscopy beamline (BL 2.2) at the Synchrotron Light Research Institute detector. The percentages of CO2 conversion, CH4 yield, and CH4
(SLRI), Nakhon Ratchasima, Thailand. For BL 5.2, synchrotron radiation selectivity and the rate of CO2 reaction were calculated by using the
was tuned by a double crystal monochromator (DCM) equipped with Ge following equations:
(220) crystals for Ni K-edge and InSb(111) crystals for Zr L3-edge, (
[CO2 ]out
)
respectively. The monochromatic flux was obtained between 108 and % CO2 conversion = 1 -
[CO2 ]out + [CH4 ]out + [CO]out
× 100
1010 photons/sec at 100 mA. Additionally, BL 2.2, the bent crystal was
(2)
used as an energy dispersive monochromator and Si(111) crystal was
used as a monochromator with an energy range of 2.4–10 keV. The (
[CH4 ]out
)
XANES spectra were obtained by subtracting I1 by I0 with the pre­ % CH4 yield = × 100 (3)
[CO2 ]out + [CH4 ]out + [CO]out
processing software before being exported to Athena software. There are
2 experimental sets for investigating the catalyst properties with (
CH4 yield
)
different aims as follows;. % CH4 selectivity = × 100 (4)
CO2 conversion

(i) Ex-situ X-ray absorption spectroscopy for investigating the elec­ CO2 conversion x FCO2, in
tronic state of Ni, Ti, and Zr of freshly prepared catalysts. The Ni K CO2 methanation rate = (5)
W
edge and Zr L3 edge XANES spectra were obtained from BL 5.2
while Ti K edge XANES spectra were collected at BL 2.2. For Ni K where [CO2]out, [CH4]out, and [CO]out are outlet concentrations of CO2,
edge and Zr L3 edge XANES experiment, the spectra were CH4, and CO, respectively. FCO2, in is inlet molar flow rate of CO2 and W
collected by using fluorescence mode with 4-element silicon drift is the sample weight [29].
as a detector which was placed beside the sample to detect the
fluorescence (IF) signal. For Ti K edge XANES spectra, the 3. Results and discussion
experiment was conducted in a transmission mode with ioniza­
tion chamber as a detector. The ionization chamber was placed in 3.1. Catalytic activity for CO2 methanation reaction
front of and behind the sample to detect the incident (I0) and
transmitted (I1) X-ray beam. The catalytic activities of the prepared catalysts were tested for CO2
(ii) In-situ X-ray absorption spectroscopy for monitoring the elec­ methanation reaction in the temperature range of 100–500 ◦ C. Fig. 1(a)–
tronic state change of Ni during CO2 methanation reaction. This (c) show the percentages of CO2 conversion, CH4 yield, and products
experiment was carried out at BL 2.2 by monitoring on the Ni K (CO and CH4) selectivity as a function of reaction temperature, respec­
absorption edge. Firstly, the sample was mixed with boron nitride tively. It is clearly seen that the CO2 conversion and CH4 yield of all
(BN) to obtain a homogeneous powder and pressed into a pellet. prepared catalysts was enhanced with increased reaction temperature
The pellet of the sample was placed in an in-situ cell and then from 100 to 450 ◦ C and then slightly decreased after the reaction tem­
pretreated by heating the sample cell from room temperature to perature reached 500 ◦ C. The reduction of CH4 selectivity at high tem­
450 ◦ C (ramp rate of 5 ◦ C/min) under H2 flowing at 24 mL/min. perature was due to the formation of the reversed-water gas shift side
The spectra were collected at every 10 ◦ C increment during the reaction, which produced an undesired product (CO). The bare TiO2
increasing of temperature. Next, the temperature was kept at catalyst was not active for this reaction. After impregnating Ni onto
450 ◦ C for 90 min and the spectra were collected every 10 min. TiO2, the CO2 conversion was dramatically enhanced, which resulted
After the pretreatment step, the temperature was reduced to from the existence of an active site which was reactive toward both
100 ◦ C and the mixed feed gas of H2 (24 mL/min) and CO2 (6 mL/ reactant molecules (H2 and CO2). Upon incorporating Zr into TiO2 and
min) was then switched over the sample to start the CO2 then impregnating it with Ni species, the CO2 methanation catalytic
methanation reaction. The reaction temperature was kept at performance was shifted toward lower reaction temperature with
100 ◦ C for 20 min and the spectra were collected every 10 min at improving the CO2 conversion percentage. The highest CO2 methana­
this point. After that, the reaction temperature was increased to tion catalytic activity was obtained by Zr addition only 10 mol% (10 wt
150 ◦ C (ramp rate of 5 ◦ C/min) and the spectra were collected %Ni-Zr0.1Ti0.9O2), with CO2 conversion of 68.0% at 450 ◦ C. However,
every 10 ◦ C of increment. When the reaction temperature reached the catalytic activity decreased with further increased Zr content. CH4
150 ◦ C, the spectra were collected every 10 min during holding selectivity of 10 wt%Ni-Zr0.1Ti0.9O2 was higher than 96% in a wide
for 20 min. Then, the reaction temperature was increased every range of reaction temperatures (150–350 ◦ C) when compared with other

4
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Table 1
Catalytic system based on TiO2 support by using Ni as active metal for CO2 methanation.
Catalysts Ni doping CO2 methanation conditions CO2 CH4 Ref.
preparation methods conversion (%) selectivity (%)

15%Ni/TiO2- Conventional Fixed-bed reactor (id: 8 mm), 350 ◦ C, 1 atm, 1.0 g catalyst, pretreated at 450 ◦ C under 80 95.5 [30]
IMP impregnation 100 mL/min H2 +N2 for 4 h, H2:CO2:Ar = 12:3:5, total flow rate = 40 mL/min− 1, GHSV
= 2400 h− 1
10%NiO/TiO2- incipient Fixed-bed reactor, 450 ◦ C, 1 atm, 50 mg catalyst diluted by 0.1 silica carbide, pretreated 60 91.7 [31]
C wetness at 700 ◦ C under 40 mL/min H2 for 1 h, H2:CO2 = 4:1, GHSV = 60,000 h− 1
impregnation
15%Ni/ impregnation Fixed-bed reactor (id: 6 mm), 400 ◦ C, 1 atm, 9 g catalyst, pretreated at 600 ◦ C under 57 97 [32]
TiO2@SiO2 10 L h− 1 H2 + 20 L h− 1 N2 for 2 h, H2:CO2 = 4:1, total flow rate of 25 L h1
10%Ni-Ce/ incipient wetness Fixed bed reactor (id: 6 mm), 450 ◦ C, 1 atm, 50 mg catalyst, pretreated at 450 ◦ C under 65 95 [33]
TiO2 impregnation 40 mL/min H2 for 1.5 h, H2:CO2:He = 24:6:10, total floe rate = 40 mL/min, GHSV
= 4250 h− 1
12%Ni/TiO2 sol–gel Fixed bed reactor (id: 6 mm), 450 ◦ C, 1 atm, 50 mg catalyst, pretreated at 450 ◦ C under 40 97 [34]
40 mL/min H2 for 1.5 h, H2:CO2:He = 24:6:10, total floe rate = 40 mL/min
10%Ni/TiO2- impregnation Fixed bed reactor, 340 ◦ C, 6.1 bar, 200 mg catalyst diluted by 1800 mg silicon carbide, 42.6 79 [35]
anatase pretreated at 400 ◦ C under N2 +H2 atmosphere for 4 h, N2:H2:CO2 = 1:4:1
10%Ni-TiO2 incipient Fixed bed reactor (id: 6 mm), 450 ◦ C, 1 atm, 50 mg catalyst, pretreated at 450 ◦ C under 55 83 This
10%Ni-Zr/ wetness 40 mL/min H2 for 1.5 h, H2:CO2:He = 24:6:10, total floe rate = 40 mL/min, GHSV 68 94 work
TiO2 impregnation = 4250 h− 1

Fig. 2. (a) XRD patterns of ZrxTi1− xO2 (x = 0, 0.15, and 0.2) supports (b) Magnification of (101) peak at ~25.4◦ for ZrxTi1− xO2 (x = 0, 0.15, and 0.2) supports (c)
XRD patterns of TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts (d) Magnification of (101) peak at ~25.4 ◦ for all synthesized catalysts.

prepared catalysts. Furthermore, the addition of optimal Zr content can reaction rate. To avoid the mass and heat transfer effect, the kinetic
reduce the formation of undesired CO product. Table 1 presents the experiment was studied in the temperature range of 190–280 ◦ C in
catalytic activity for CO2 methanation by using the catalysts based on which the CO2 conversion was less than 10%. To determine the apparent
Ni-doped TiO2 in this work compared with other works. activation energy, an Arrhenius plot, which shows a relationship be­
The kinetic study in a terms of apparent activation energy (Ea) was tween ln (rate) and the reciprocal of absolute temperature was con­
also carried out to investigate the role of the catalysts in enhancing the structed as shown in Fig. 1(d). The slope of this plot was used to

5
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Table 2
Crystalline size, unit cell parameters, surface area, and metals content of all prepared catalysts.
Catalyst Crystalline size (Å) Lattice parametersa BET surface areab (m2/g) Metal contentc
d
NiO TiO2e a = b (Å) c (Å) 3
cell volume (Å ) Ni (%wt) Zr (%mol)

TiO2 – 92.5 3.790 9.551 137.2 86.36 – –


10 wt%Ni-TiO2 129.3 95.6 3.789 9.538 136.9 56.20 8.71 –
10 wt%Ni-Zr0.1Ti0.9O2 148.0 67.9 3.807 9.624 139.5 84.05 8.79 7.34
10 wt%Ni-Zr0.15Ti0.85O2 192.3 73.5 3.815 9.658 140.6 103.12 8.90 14.51
10 wt%Ni-Zr0.2Ti0.8O2 201.4 81.6 3.821 9.668 141.2 135.75 8.83 16.34
a
Calculated from (101), (004), (200) planes of TiO2 by using Bragg’s equation.
b
Calculated by using Multipoint BET method.
c
Calculated from ICP-OES results.
d
Calculated from (101) plane of TiO2 by using Sherrer’s equation.
e
Calculated from (200) plane of NiO by using Sherrer’s equation.

calculate the apparent activation energy. The calculated apparent acti­


vation energies of 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) were
79.34, 76.04, 77.18 and 79.99 kJ/mol, respectively. It is clearly seen
that 10 wt%Ni-Zr0.1Ti0.9O2 exhibited the lowest apparent activation
energy (76.04 kJ/mol) which was in agreement with the highest cata­
lytic activity for CO2 methanation.
The long-term stability of the best catalyst (10 wt%Ni-Zr0.1Ti0.9O2)
for CO2 methanation reaction was also investigated because metallic Ni
can be sintered under exothermic conditions and carbon deposition can
also be formed on the catalyst surface, which led to catalyst deactiva­
tion. In this work, a long-term stability test was investigated at 450 ◦ C
for 50 h and the results are shown in Fig. 1(e). The percentages of CO2
conversion, CH4 yield, and CH4 selectivity were almost stable without an
obvious decrease of those values in the range of stability test (50 h). This
indicated that 10 wt%Ni-Zr0.1Ti0.9O2 catalyst had great stability.

3.2. Catalyst characterization

XRD patterns of TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15,


and 0.2) catalysts are revealed in Fig. 2(a). The diffraction peaks of bare Fig. 3. N2 adsorption-desorption isotherms of TiO2 and 10 wt%Ni-ZrxTi1− xO2
TiO2 were observed at around 25.4, 37.8, 48.2, 54.1, 55.1, and 62.8◦ (x = 0, 0.1, 0.15, and 0.2) catalysts (solid symbol: adsorption branch, open
which corresponded to the (101), (004), (200), (105), (211), and (204) symbol: desorption branch).
crystallographic planes of anatase structure, respectively (JCPDS file
number 21–1272) [36–39]. All modified catalysts exhibited all above TiO2 which confirmed that NiO was dispersed on the catalyst surface. In
diffraction peaks corresponding to anatase TiO2, which indicated that contrast, unit cell enlargement occurred upon the addition of Zr into the
the structure of those catalysts remained unchanged after the addition of TiO2 lattice. Larger unit cell expansion of host lattice was observed with
both Ni and Zr. Fig. 2(a) illustrates XRD patterns of bare TiO2 compared increased Zr loading. This indicated that Zr ion incorporated into TiO2
with modified supports (ZrxTi1− xO2; x = 0.15, and 0.20). Only high Zr lattice to form a solid solution, which can be evidenced by the absence of
addition samples were selected to investigate a clearly shifting of the XRD peaks of ZrO2 even though the Zr content reached 20 mol%. The
diffraction peaks in Fig. 2(a). It is seen that upon addition of Zr into TiO2, TiO2 crystalline size tended to decrease upon the addition of Zr, which
the diffraction peaks corresponding to zirconium oxide phase were not was probably due to the inhibition of TiO2 crystal growth by the co-
observed. Moreover, the diffraction peaks were shifted to lower addition of Zr during the synthesis procedure. Upon increased Zr con­
diffraction angle (Fig. 2(b)) which indicated to an expansion of the host tent, the crystalline size of TiO2 was slightly increased. On the other
lattice. This can indicate that Zr incorporated into TiO2 lattice formed as hand, NiO crystalline size was further increased with increased Zr
a solid solution. The incorporation of larger Zr4+ (0.072 nm) ions into amount.
smaller Ti4+ (0.060 nm) sites inside TiO2 host lattice would lead to N2 adsorption-desorption isotherms of TiO2 and 10 wt%Ni-impreg­
distortion and enlargement of the host supporting material lattice. nated mixed oxide catalysts are shown in Fig. 3. All catalysts exhibited
Increased Zr content in the catalysts resulted in more shifting of (101) the type IV isotherm with hysteresis loop, which indicated the meso­
crystallographic plane to lower diffraction angle, which indicated a porous structure [38,41]. The calculated specific surface area of TiO2
higher amount of substitution of Ti4+ by Zr4+. Moreover, upon and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts by BET
impregnating Ni onto bare TiO2 and titania-zirconia mixed oxide sup­ method are shown in Table 1. It can be seen that the BET surface area
port, an additional diffraction peak at around 43.4◦ which corresponded decreased from 86.36 m2/g to 56.20 m2/g after 10 wt%Ni was added on
to (200) plane of cubic NiO (JCPDS file number 47-1049) [39,40] was TiO2, which was probably due to a partial blockage or collapse of
observed as shown in Fig. 2(c). The existence of NiO diffraction peak mesopores during the impregnation process [11]. The surface area of
together with no shifting of TiO2 characteristic peak for 10 wt%Ni-TiO2 impregnated 10 wt%Ni on mixed oxide support was higher than that on
in Fig. 2(d) when compared with other catalysts can indicate that added bare TiO2, which corresponded to higher surface area of supporting
Ni species formed on catalyst surface as NiO. material due to their low crystalline sizes as mentioned in XRD results.
Table 2 summarized the cell parameters, cell volumes, and crystal­ However, upon further increasing Zr content, the surface area was still
line sizes of all prepared catalysts. There was no significant difference increased, which was opposite to their crystalline size. This was prob­
between the cell parameters of bare TiO2 and 10 wt%Ni-impregnated ably due to the formation of the interparticle cavities (self-assembled

6
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

TiO2 10wt%Ni-TiO2 10wt%Ni-Zr0.1Ti0.9O2

10wt%Ni-Zr0.15Ti0.85O2 10wt%Ni-Zr0.2Ti0.8O2

Fig. 4. FE-SEM images of TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts.

To investigate the reduction behavior and interaction of NiO with the


support of the prepared catalysts, H2-TPR profiling was performed.
Fig. 5 shows the H2-TPR profiles of TiO2 and 10 wt%Ni-ZrxTi1− xO2
(x = 0, 0.1, 0.15 and 0.2) catalysts. For TiO2, only a small peak was
observed at 619.6 ◦ C, which indicated that TiO2 catalyst was hardly
reduced in the temperature range of 50–800 ◦ C. When Ni was impreg­
nated onto TiO2 support (10 wt%Ni-TiO2), two main reduction peaks at
340.5 ◦ C and 471.3 ◦ C were found. Those two reduction peaks were
attributed to the reduction of NiO to metallic Ni (Ni2+ → Ni0). It has been
reported that different reduction temperature of Ni2+ to Ni0 was related
to its binding strength to supporting material or metal-support interac­
tion [30,43,44]. The strength of interaction between Ni and support can
be categorized into three regions. The first low reduction temperature
ranges around 250–300 ◦ C was attributed to the reduction of bulk NiO
particle without interaction with the support. The second reduction
temperature range around 300–600 ◦ C was assigned as the reduction of
NiO with moderately binding strength to the support. Finally, the last
high reduction temperature (above 600 ◦ C) was assigned as the reduc­
tion of NiO with a strong binding strength to the support [44]. There­
Fig. 5. H2-TPR profiles of TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15 and fore, those two reduction peaks on 10 wt%Ni-TiO2 were attributed to
0.2) catalysts. moderately binding strength between NiO and TiO2. Moreover, those
two reduction peaks were also indicated to two difference kinds of
NPs) [42] of the catalyst. interaction between NiO and TiO2; i.e., the reduction temperature
Fig. 4 presents the FE-SEM images of TiO2 and 10 wt%Ni-ZrxTi1− xO2 around 340.5 ◦ C and 471.3 ◦ C were related to lower and higher binding
(x = 0, 0.1, 0.15, and 0.2) catalysts with the same magnification of strength between them, respectively. Incorporation of Zr into TiO2 lat­
200000x. It can be seen that the morphologies of TiO2, 10 wt%Ni-TiO2, tice resulted in shifting those two reduction peaks toward higher tem­
and 10 wt%Ni-Zr0.1Ti0.9O2 showed an agglomeration of spherical par­ perature (360–376 ◦ C for low reduction temperature side and
ticles to form a rough surface with the average particle size in the range 476–492 ◦ C for high reduction temperature side). Normally, the
of 13–15 nm. For 10 wt%Ni-Zr0.15Ti0.85O2 and 10 wt%Ni-Zr0.2Ti0.8O2, reduction temperature of the NiO can indicate the interaction between
the particles clearly exhibited an almost spherical shape with particle NiO and catalyst support, i.e., higher reduction temperature implied to
sizes of 26.7 and 30.1 nm, respectively. The 10 wt%Ni-Zr0.2Ti0.8O2 higher or stronger interaction between them. However, it can be seen
catalyst showed the largest particle size. However, the highest surface that the amount of H2 consumption for those two reduction peaks in
area was found for this catalyst, which was probably due to the forma­ Fig. 5 were significantly varied upon addition of Zr. Therefore, the
tion of the highest interparticle cavities. amount of H2 consumption of those two reduction peaks, which were

7
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Table 3
The dispersion of Ni species expressed by AH/(AH+AL) ratio which calculated from H2-TPR results and the desorbed CO2 amount from CO2-TPD result.
Catalyst AH/ (AH+AL) Amount of H2 consumption (mmol/g) Desorbed CO2 (mmol/g)

Lower temperature Higher temperature Percentage of strong interaction

10 wt%Ni-TiO2 0.685 0.313 0.760 70.8 0.204


10 wt%Ni-Zr0.1Ti0.9O2 0.674 0.219 0.451 67.3 0.236
10 wt%Ni-Zr0.15Ti0.85O2 0.356 0.552 0.305 35.6 0.292
10 wt%Ni-Zr0.2Ti0.8O2 0.443 0.651 0.517 44.3 0.337

10Ni-10Zr

(a) (b)
Mean = 19.41 nm Mean = 20.17 nm
Mean = 12.48 nm Mean = 14.25 nm (c) (d)

Counts
Counts
Counts

Counts

0 20 40 60 80
0 5 10 15 20 25 30 35 40 0 10 20 30 40 50 60 70 Particle diameter (nm)
Particle diameter (nm) 0 10 20 30 40 50 Particle diameter (nm)
Particle diameter (nm)

Fig. 6. TEM images of reduced (a) 10 wt%Ni-TiO2, (b) 10 wt%Ni-Zr0.1Ti0.9O2, (c)10 wt%Ni- Zr0.15Ti0.85O2 and (d) 10 wt%Ni-Zr0.2Ti0.8O2 catalysts.

obtained by integrating of the peak area, were also taken into account
beside the shifting of the reduction temperature. The amount of H2
consumption of both two reduction temperatures were summarized in
Table 3. To more clarify on the amount of lower and higher binding
strength between NiO and TiO2 in all samples, the percentage of higher
binding strength (higher reduction peak in the range of 472–493 ◦ C)
were reported and compared. It can be seen that 10 wt%Ni-TiO2 and
10 wt%Ni-Zr0.1Ti0.9O2 exhibited higher fraction of higher binding
strength or strong interaction (70% and 67%, respectively). In contrast,
10 wt%Ni-Zr0.15Ti0.85O2 and 10 wt%Ni-Zr0.2Ti0.8O2 exhibited higher
fraction of lower binding strength or weak interaction (65% and 56%,
respectively). Since, the impregnated amount of Ni for all sample (with
and without Zr addition) were almost the same, therefore, different
fraction of weak and strong interaction of all catalysts should be caused
by Zr addition. These phenomena can indicate that Zr can tune or tailor
the metal (Ni)-support (TiO2) interaction and then also affect the cata­
lyst reducibility and NiO dispersion on the catalyst surface. Therefore, it
implied that the electronic state of Ni species must be tuned by the
addition of Zr. For this electronic state tuning phenomenon, X-ray ab­
sorption spectroscopy results were evidenced, which will be discussed
later.
To estimate the dispersion of Ni species on the catalyst’s surface, the Fig. 7. CO2-TPD profiles of 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and
ratio of peak area at higher temperature (AH) normalized by summation 0.2) catalysts.
of the peak area at lower (AL) and higher temperature (AH) was calcu­
lated [30] and expressed in Table 3. The results showed that 10 wt% were in agreement with the similar value of AH/(AH+AL) for both cat­
Ni-TiO2 catalyst exhibited the highest ratio, which indicated that NiO alysts which was calculated from H2-TPR results. The rest two catalysts
was highly dispersed on the catalyst surface. This ratio tended to reduce (10 wt%Ni-Zr0.15Ti0.85O2 and 10 wt%Ni-Zr0.2Ti0.8O2) showed low
upon increased Zr content, which indicated that lower NiO dispersion on dispersed Ni0 with larger particle size of reduced Ni species which also
the catalyst surface. This trend was correlated to the crystalline sizes of corresponded to XRD and H2-TPR results.
NiO obtained from the XRD results; i.e., a higher crystalline size of NiO The basicity and binding strength between CO2 molecule and the
with increased Zr loading. Larger crystalline size of NiO would result in catalyst was evaluated by CO2-TPD technique. In the CO2-TPD experi­
lower NiO surface area for activating the H2 molecule. ment, the surface of the catalyst was saturated by CO2, and then the
For more insight on the Ni species dispersion, transmission electron adsorbed CO2 was desorbed from the catalyst surface by increasing the
microscopy was used. Fig. 6 reveals the TEM images of 10 wt%Ni- temperature. Therefore, the amount of desorbed CO2 was proportional
ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts which was reduced at to the amount of CO2 adsorption. Moreover, the binding strength be­
450 ◦ C for 90 min (pretreatment step in CO2 methanation activity test). tween CO2 and catalyst surface can be indicated by the desorption
It can be seen that all TEM images could be divided into two regions; temperature. The CO2-TPD profiles of 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1,
dark and bright particles which corresponded to Ni0 and support area, 0.15, and 0.2) catalysts are shown in Fig. 7. It can be seen that the CO2-
respectively. The results showed that both 10 wt%Ni-TiO2 and 10 wt% TPD profiles exhibited three different regions: weak, medium, and
Ni-Zr0.1Ti0.9O2 exhibited similar dispersion of metallic Ni. These results strong regions. These regions were attributed to weak, medium, and

8
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Fig. 8. (a) XANES Spectra of 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) compared with Ni foil and NiO standards. (b) White line intensity comparison of NiO
standard and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2).

strong basicity of the catalyst, which related to the strength of binding imply higher basicity of catalysts.
between CO2 and catalyst surface with weak, medium, and strong
interaction, respectively. These three sites were found for 10 wt%Ni- 3.3. The role of modified catalyst on changes of electronic properties of
TiO2, and 10 wt%Ni-Zr0.1Ti0.9O2 catalysts while only two sites (weak catalyst investigated by X-ray absorption spectroscopy
and medium) were found for 10 wt%Ni-Zr0.15Ti0.85O2 and 10 wt%Ni-
Zr0.2Ti0.8O2 catalysts. However, only adsorbed CO2 on weak and me­ From the above results, the optimal Zr addition into the catalyst can
dium basic sites are favor to involve in the CO2 methanation reaction improve the CO2 conversion percentage and also shift the reaction to
while the strong adsorbed CO2 on the catalysts cannot participate in the occur at lower reaction temperature side. This is because the results
reaction, since the reaction temperature was in the range of 100–500 ◦ C, from H2-TPR indicated that the metal-support interaction was tuned
thus, the adsorbed CO2 in the strong region (over 550 ◦ C) should be upon the incorporation of Zr into the TiO2-supported Ni catalyst, which
excluded [45]. Therefore, only those two former peaks (weak and me­ might be due to the changing of electronic state of Ni by Zr addition.
dium) were taken into account. The summation of desorbed CO2 amount Therefore, in this section, X-ray absorption spectroscopy was used to
of weak and medium basic sites was calculated and presented in Table 2. evidence the tailoring of those electronic perturbations. There are 2
It can be seen that the CO2 adsorption of Zr-doped catalysts (10 wt% experimental sets for conducting X-ray absorption spectroscopy with the
Ni-ZrxTi1− xO2, x = 0.1, 0.15, and 0.2) was higher than that of different aims as follows;.
un-modified support (10 wt%Ni-TiO2) and the amount of CO2 uptake
was increased with further increased Zr content. This indicated that Zr 1) Ex-situ X-ray absorption spectroscopy for investigating the electronic
addition can increase the CO2 affinity on the catalyst surface which can state of Ni, Ti and Zr of freshly prepared catalysts.

Fig. 9. (a) Ti K edge XANES spectra of anatase and rutile TiO2 standards, synthesized TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0 and 0.1) catalysts. (b) Comparison
between pre-edge of Ti K edge XANES spectra of TiO2 and 10 wt%Ni-Zr0.1Ti0.9O2 catalysts.

9
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

2) In-situ X-ray absorption spectroscopy for monitoring the electronic


state change of Ni during CO2 methanation reaction.

3.3.1. Ex-situ X-ray absorption spectroscopy for investigating the electronic


state of Ni, Ti and Zr of freshly prepared catalysts
Fig. 8(a) illustrates the normalized Ni K edge XANES spectra of 10 wt
%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts compared with NiO
(Ni2+) and Ni foil (Ni0) standards. It is seen that the Ni K edge XANES
spectra features of all prepared catalysts was similar to that of NiO
standard. The white line peak at 8350 eV was ascribed to the transition
from 1 s to unoccupied 4p state and the peak at 8366 eV was a signature
of six-fold coordinated nickel by oxygen [46,47]. In order to confirm the
electronic state of Ni species on the catalysts, comparison of all above
catalysts with NiO standard was constructed, as shown in Fig. 8(b). All
prepared catalysts exhibited the same absorption edge as NiO which
indicated that the oxidation state of Ni species on catalyst was 2+.
It can be noticed that the white line intensities of all modified cat­
alysts were different (clearly seen in inset) which can indicate the dif­
ference of the electron density within their d valence state. In the X-ray
absorption process, the corresponding core electron of probe atom
Fig. 10. Zr L3 edge XANES spectra of Zr foil and ZrO2 standards and 10 wt%Ni-
ZrxTi1− xO2 (x = 0.1, 0.15 and 0.2) catalysts.
absorbed an appropriate energy to its binding energy and was excited
into unoccupied valenced state. The absorption signal of those probe
atoms as a function of increasing incident photon energy is detected and
then the relationship between the absorption coefficient and photon
energy is constructed, which is called XAS spectrum. The different
electron density in unoccupied valence state can determine the feature
of XAS spectrum. According to exciting the core electron to low electron
density in unoccupied valence state, more electrons can be moved into

Fig. 11. Ni K edge XANES spectra during pretreatment and CO2 methanation conditions for (a) 10 wt%Ni-TiO2 and (b)10 wt%Ni-Zr0.1Ti0.9O2. (c) The Normalized
peak intensity at 8350 eV from Ni K edge XANES spectra as a function of pretreatment temperature for 10 wt%Ni-TiO2 and 10 wt%Ni-Zr0.1Ti0.9O2.

10
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

this band resulting in highly intense peak in the XAS spectra which is methanation reaction was monitored by time-resolved X-ray absorption
seen by as a white line. Therefore, a comparison of the white line in­ spectroscopy (TRXAS). In the in-situ XAS experiment, the procedure was
tensities of probe atoms can indicate the electron density in the unoc­ similar to the catalytic activity test. The sample was firstly pretreated
cupied valence band. From Fig. 8(b), when comparing only NiO and under H2 atmosphere at 450 ◦ C for 90 min to reduce NiO to metallic Ni
10 wt%Ni-TiO2, the Ni K edge white line intensity of the latter sample active species. After the pretreatment step, the system was cooled down
was lower than that of the former sample. This indicated that the elec­ under H2 flow to 100 ◦ C and then the CO2 methanation reaction was
tron density in valence d state of Ni in 10 wt%Ni-TiO2 was higher than started by switching to the mixed feed gas (CO2:H2 with ratio of 1:4)
that of NiO which is due to electron movement from Ti to Ni probe atom. over the sample and then the Ni K edge XANES spectra were collected.
Moreover, when comparing only the modified catalysts, 10 wt%Ni- After that, the reaction temperature was increased for each 50 ◦ C and
ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2), the white line intensity of Ni K the XANES spectra were collected until the reaction temperature
edge was reduced upon increased Zr loading. This result implied that reached 500 ◦ C. Fig. 11(a) and 11(b) illustrate the Ni K edge XANES
some electron movement to valence d state of Ni also occurred and those spectra during the pretreatment step and CO2 methanation reaction for
electrons would come from added Zr in the TiO2 lattice. Therefore, Zr 10 wt%Ni-TiO2 and 10 wt%Ni-Zr0.1Ti0.9O2 catalysts, respectively. These
can perturb the electronic properties of Ni and this tuning phenomenon two catalysts exhibited a similar changing of XANES spectra during the
can dominate the catalytic activity for CO2 methanation. two steps (pretreatment and CO2 methanation reaction). For the pre­
Fig. 9(a) displays normalized Ti K edge XANES spectra of TiO2 and treatment step, the oxidation state of Ni was changed from Ni2+ to Ni0
10 wt%Ni-ZrxTi1− xO2 (x = 0 and 0.1) catalysts compared with those of with increased temperature from 50 to 450 ◦ C and held at 450 ◦ C for
anatase and rutile TiO2 standards. All prepared catalysts exhibited a 90 min. During the pretreatment step, Ni2+ in both catalysts were
similar feature to the anatase TiO2 standard. Therefore, all catalysts completely reduced to Ni0. However, the starting reduction temperature
were in the anatase (Ti4+) phase which corresponded to the XRD results. of both catalysts was different, which can point to the strength of the
Confirmation of the incorporation of Zr into the TiO2 lattice can be metal-support interaction. To evidence the strength of the Ni species
identified by the changing of the pre-edge of Ti K edge. The pre-edge of with the catalysts surface, the plotting of normalized peak intensity at
Ti K edge XANES spectra exhibited three main peaks; A1, A2 and A3. A1 8350 eV (white line peak of Ni2+) as a function of temperature was
peak was ascribed to the transition from 1 s to 3d-t2 g state. A2 and A3 constructed, as shown in Fig. 11(c). The normalized peak intensity can
peaks were ascribed to the transition from 1 s to hybridized p-d (t2 g) and be calculated from the difference of peak intensity at 8350 eV between
p-d (eg), respectively [48,49]. The higher those peak intensities were the initial state and that of each temperature divided by that of the initial
originated from the distortion of TiO6 symmetry by substituting of Zr4+ state. Therefore, an increase of normalized peak intensity indicated that
ion. Therefore, the higher the pre-edge peak intensity of 10 wt% Ni2+ was more rapidly reduced to Ni0. The difference in starting
Ni-Zr0.1Ti0.9O2 than TiO2 as shown in Fig. 9(b) indicates the distortion of reduction temperature was due to different binding strength between Ni
the TiO2 lattice, which was caused by the incorporation of Zr4+ ion. and support. According to the results from H2-TPR, the Ni species of
Normalized Zr L3 edge XANES spectra of 10 wt%Ni-ZrxTi1− xO2 10 wt%Ni-Zr0.1Ti0.9O2 exhibited higher binding strength to the support
(x = 0.1, 0.15, and 0.2) catalysts and standards (ZrO2 and Zr foil) are than Ni of the 10 wt%Ni-TiO2 sample, which is evidenced by shifting the
shown in Fig. 10. It is seen that all modified catalysts exhibited the same reduction temperature of Ni2+ to higher temperature side. It can be seen
absorption edge as ZrO2 standard which indicated that the Zr probe in all that the starting reduction temperature of 10 wt%Ni-TiO2 was observed
catalysts were in the Zr4+ state. It is noticed that the spectral feature of at lower temperature than that of the 10 wt%Ni Zr0.1Ti0.9O2. This indi­
all modified catalysts was different, which can indicate the different cated that Ni2+ in the 10 wt%Ni-TiO2 catalyst was more easily reduced
environment around Zr ion. It has been reported that the Zr L3 edge than that of the 10 wt%Ni-Zr0.1Ti0.9O2 catalyst. Although 10 wt%Ni-
XANES spectral shape depended on the coordination number and the TiO2 was more easily reduced, however, Ni2+ in both prepared catalysts
symmetry around the Zr ion [50]. Normally, the Zr L3 edge XANES was completely reduced to Ni0 after finishing the pretreatment step.
spectrum exhibited two peaks labeled as A and B which were attributed Therefore, a few differences in the starting reduction temperature did
to the electron transition from 2p3/2 to 4d(t2 g) and 4d(eg), respectively. not affect the CO2 methanation activity of both catalysts since the initial
The degree of difference between the energy level of t2 g and eg can Ni species for the CO2 methanation reaction was in the same species
estimate the XANES spectral feature and can also indicate the coordi­ (Ni0). Furthermore, under CO2 methanation conditions, the oxidation
nation number around Zr ion. A larger difference between those two state of Ni for both catalysts remained unchanged from the initial state
energy levels would lead to a separation of A and B peaks which indi­ (Ni0).
cated a lower coordination number around Zr ion. For 10 wt%Ni-Zr0.1
Ti0.9O2, the A and B peaks were clearly observed, which indicated that 3.4. The role of Zr addition on properties and catalytic activity of the
the t2 g and eg energy levels were well separated. This indicated that catalyst
10 wt%Ni-Zr0.1Ti0.9O2 had the lowest coordination number around Zr
since it had the lowest amount of Zr4+ (10 mol%) incorporation into the From the results of CO2 methanation catalytic activity, the catalytic
TiO2 lattice. In contrast, the A and B peaks of 10 wt%Ni-Zr0.15Ti0.85O2 performance of bare TiO2 was much enhanced by impregnation of Ni,
were less separated and those peaks were almost merged for the 10 wt% which was due to the existence of the active site for dissociative
Ni-Zr0.2Ti0.8O2 sample. This indicated that a smaller difference of the adsorption of H2 molecules. CO2 methanation catalytic activity was
energy level between t2 g and eg occurred upon increased Zr content. further increased upon doping of Zr4+ into TiO2 support. Since the bare
Therefore, it implied that higher coordination number with increased Zr TiO2 and zirconia-titania mixed oxide supports were firstly synthesized
loading occurred, which was due to higher amount of substituted Zr4+ and then the same amount of Ni was impregnated onto those supports,
into the TiO2 lattice. However, comparing the white line intensity of the increased CO2 methanation catalytic activity of Ni-impregnated zirco­
Zr L3 edge to evidence the electron movement from Zr to Ni cannot be nia-titania mixed oxide support resulted from Zr4+ addition. From all of
used since the amount of Ni impregnated onto Zr-Ti support was almost the above characterization techniques, the role of Zr doping on catalysts
the same, but the loading amount of Zr was different; i.e.; the properties which also affected their catalytic activity were summarized
electron-withdrawing effect by Ni was constant, but the Zr content was as follows;.
increased.
i) the incorporation of Zr4+ ion into the TiO2 lattice improved the
3.3.2. In-situ X-ray absorption spectroscopy for monitoring the electronic physical properties of the catalysts. The added Zr incorporated
state change of Ni during CO2 methanation reaction into TiO2 lattice to form a solid solution structure which resulted
In this part, the change of Ni oxidation state during the CO2 in a reduction of TiO2 crystalline size. This was because the

11
­
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Fig. 12. Dispersion of Ni species (expressed as ratio of AH/(AH+AL)) and des­


orbed CO2 of all modified catalysts as a function of Zr content.

addition of Zr could inhibit TiO2 crystal growth, leading to sur­ Fig. 13. Temperature-dependent CO conversion of 10 wt%Ni-TiO2 and 10 wt%
face area enhancement when compared with unmodified support Ni-Zr0.1Ti0.9O2 catalysts for CO methanation (CO:H2 = 1:4).
(10 wt%Ni-TiO2). However, the TiO2 crystalline size increased
with further increased Zr content, which was probably due to an inhibit H2 adsorption on the catalyst surface. Therefore, optimal
excessive Zr addition decreased the amount of phase boundary. H2 and CO2 adsorption ability on the surface of 10 wt%
Furthermore, the surface area also increased with increased Zr Ni-Zr0.1Ti0.9O2 catalyst would lead to its highest catalytic
content, which was possibly due to higher interparticle cavities activity.
formation. In the case of NiO, the larger crystalline size was ob­ iii) From the above discussion, it can be concluded that the role of Zr
tained with increased Zr loading. It is noticed that the impreg­ addition was to tune the interaction between NiO and TiO2 sup­
nated Ni for all catalysts were an almost constant amount, but port and resulting in improving the catalyst properties which can
their crystalline size was different, which indicated that NiO promote the CO2 methanation reaction rate. This tuning phe­
would also exist in a different active area. This implied that Zr can nomenon was caused by the electronic property changing of Ni
alter the Ni species property such as its binding strength with the species when Zr was doped into TiO2. Upon the addition of Zr into
support which can be evidenced by, the H2-TPR results. the TiO2 support which is then impregnated by Ni, electrons from
ii) the incorporation of Zr4+ into TiO2 lattice facilitated the Zr were transferred to Ni. This was due to higher electronega­
adsorption ability of H2 and CO2 by tuning the interaction be­ tivity of Ni (1.91) than that of Zr (1.33). From the electron
tween NiO-support and increasing the catalyst basicity, respec­ movement phenomena, Ni had a high electron density in valence
tively. According to the H2-TPR results, the reduction d state, which was beneficial to the activation of CO molecules. It
temperature of NiO was shifted toward higher temperature upon is known that the CO2 methanation reaction mechanism can
the addition of Zr. This indicated that higher binding strength proceed via CO intermediate, which originated from dissociation
between NiO and TiO2 support existed when Zr was incorporated of CO2 on the catalyst surface. To elucidate that CO was probably
into the TiO2 lattice. An appropriate binding strength between occurred as an intermediate for this reaction, CO methanation
them would promote the highest Ni dispersion on the catalyst catalytic activity of two catalysts (with and without Zr addition)
surface, which can facilitate the adsorption and dissociation of H2 was examined as shown in Fig. 13. For CO methanation, CO
molecules to involve in the reaction pathway. Therefore, tuning molecule was directly adsorbed on catalyst surface and then
the interaction between metal and support was one of the main further associated with adsorbed H atom into various interme­
effects for dominating the catalytic activity. This point can be diate and then finally converted to CH4. Therefore, we used this
described by the results from XAS experiment in the next sub­ reaction to indirect support that Zr addition can promote the
section. Another one important effect of Zr addition was to in­ activation of CO on catalyst surface. Higher CH4 production
crease the catalyst basicity. From the CO2-TPD results, doping of implied that CO was favor to adsorb on catalyst surface and then
Zr led to an increase of CO2 adsorption amount, which indicated convert to product [52]. The results in Fig. 13 illustrated that
higher basicity upon the addition of Zr. Chen and co-workers [51] catalyst with Zr addition (10 wt%Ni-Zr0.10Ti0.90O2) had much
reported a DFT study of the basic property of TiO2 and ZrO2 by higher CO conversion than catalyst without Zr addition (10 wt%
adsorption of CO2. They found that ZrO2 was more reactive to­ Ni-TiO2). Therefore, it can be expected that CO molecule was
ward CO2 than TiO2 or it can be proposed that ZrO2 has more favor to adsorb on catalyst surface and also favor to convert to
basicity than TiO2. From all of the above results, it can be CH4 on the former catalyst. From above results, it can be seen that
concluded that the co-addition of Zr into the TiO2 lattice to form a Zr addition played the role on activating of adsorbed CO molecule
mixed oxide support can improve the surface properties of the on catalyst surface. From above XAS results, the electron move­
catalysts which can promote the H2 and CO2 adsorption. Fig. 12 ment from Zr to Ni led to higher electron density in its valence
illustrates the Ni dispersion (expressed as ratio of AH/(AH+AL)) d state, or Ni acts as an electron rich site. After CO intermediate,
and the amount of desorbed CO2 which directly related to the which originated from CO2 dissociation, was adsorbed on Ni
adsorption amount of H2 and CO2 on the catalyst surface, electron rich site, the electron in d state of Ni was donated to π*
respectively, as a function of Zr content. A lower area of active Ni antibonding of CO molecule. Upon the donation of electrons from
species on the catalyst surface resulted in a lowering of dissoci­ the Ni electron rich site, the bonding between carbon and oxygen
ated H atoms to involve the hydrogenation reaction. Likewise, of the CO molecule was weakened and then the CO molecule was
higher CO2 adsorption amount can competitively adsorb and easily dissociated to carbon and oxygen. The dissociated carbon

12
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

Fig. 14. Proposed the effect of Zr addition on enhancing CH4 selectivity over Ni-TiO2 for CO2 methanation.

atom was then further reacted with nearby H atom by hydroge­ CRediT authorship contribution statement
nation reaction to form CH4. Oxygen atoms also reacted with H
atoms to produce H2O. Therefore, the role of Zr addition was to Ammarika Makdee: Investigation, Writing – original draft. Pinit
activate the CO intermediate leading to further involvement in Kidkhunthod: Methodology. Yingyot Poo-arporn: Methodology.
hydrogenation reaction to form CH4 by a lowering of the C-O Kingkaew Chayakul Chanapattharapol: Conceptualization, Writing –
bonding dissociation barrier. Therefore, facilitating the dissoci­ review & editing.
ation of CO molecule not only sped up the rate of reaction but also
improved the selectivity to CH4 production, since undesired CO
Declaration of Competing Interest
product vanished. The above discussion on the role of Zr addition
on enhancing CO2 methanation catalytic activity was expressed
The authors declare that they have no known competing financial
by a proposed schematic in Fig. 14. However, high loading of Zr
interests or personal relationships that could have appeared to influence
resulted in lowering of CO2 methanation catalytic activity and
the work reported in this paper.
CH4 selectivity. This was caused by a larger amount of electron
movement to Ni and leading to a favoring of CO adsorption on
this electron rich site. Therefore, higher CO adsorption amount Acknowledgements
on Ni would lead to inhibiting of the reactant adsorption on the
catalyst surface, thus suppressing the catalytic activity. The authors would like to thank the Department of Chemistry, Fac­
ulty of Science, Khon Kaen University for providing research facilities.
4. Conclusion Thanks to the Synchrotron Light Research Institute (Public Organiza­
tion) for XAS analysis and generous beamtime. The financial support
TiO2 and 10 wt%Ni-ZrxTi1− xO2 (x = 0, 0.1, 0.15, and 0.2) catalysts from the Science Achievement Scholarship of Thailand (SAST) and
were prepared, characterized by several techniques, and tested for their Materials Chemistry Research Center, Department of Chemistry, Faculty
catalytic activity for the CO2 methanation reaction. The 10 wt%Ni- of Science, Khon Kaen University are gratefully acknowledged.
Zr0.1Ti0.9O2 catalyst exhibited excellent CO2 methanation catalytic ac­
tivity with great stability at 350 ◦ C for 50 h. The role of Zr addition in References
the TiO2 lattice on the enhancement of the catalytic activity for CO2
[1] B. Metz, O. Davidson, H. de Coninck, M. Loos, L. Meyer, IPCC Special Report on
methanation reaction was tuning the interaction between NiO and TiO2 Carbon Dioxide Capture and Storage, Cambridge University Press, United Kingdom
support. This effect can result in an optimal metal-support interaction and NewYork, 2005, https://doi.org/10.1002/9783527818488.ch15.
and high active site dispersion which can facilitate the adsorption of H2 [2] A. Quindimil, U. De-La-Torre, B. Pereda, J.A. González-Marcos, J.R. González-
Velasco, Ni catalysts with La as promoter supported over Y- and BETA- zeolites for
molecules. Meanwhile, Zr addition can also improve the catalyst ba­
CO2 methanation, Appl. Catal. B Environ. 238 (2018) 393–403, https://doi.org/
sicity, which also increases the CO2 adsorption amount; thus, an 10.1016/j.apcatb.2018.07.034.
appropriate amount of H2 and CO2 adsorption can promote the CO2 [3] M.A.A. Aziz, A.A. Jalil, S. Triwahyono, A. Ahmad, CO2 methanation over
heterogeneous catalysts: recent progress and future prospects, Green Chem. 17
methanation catalytic activity. The tuning of electronic property resul­
(2015) 2647–2663, https://doi.org/10.1039/c5gc00119f.
ted from electron movement from Zr to Ni valence d state. After CO [4] I. Fechete, J.C. Vedrine, Nanoporous materials as new engineered catalysts for the
intermediate, which originated from CO2 dissociation, was adsorbed on synthesis of green fuels, Molecules 20 (2015) 5638–5666, https://doi.org/
Ni electron rich site, the bonding between carbon and oxygen atom of 10.3390/molecules20045638.
[5] D.C. Upham, A.R. Derk, S. Sharma, H. Metiu, E.W. McFarland, CO2 methanation by
the CO molecule was then weakened and easily dissociated to carbon Ru-doped ceria: the role of the oxidation state of the surface, Catal. Sci. Technol. 5
and oxygen to hydrogenate with nearby H atom to form products. (2015) 1783–1791, https://doi.org/10.1039/c4cy01106f.
[6] J.H. Kwak, L. Kovarik, J. Szanyi, CO2 reduction on supported Ru/Al2O3 catalysts:
cluster size dependence of product selectivity, ACS Catal. 3 (2013) 2449–2455,
https://doi.org/10.1021/cs400381f.

13
A. Makdee et al. Journal of Environmental Chemical Engineering 10 (2022) 107710

[7] C. Swalus, M. Jacquemin, C. Poleunis, P. Bertrand, P. Ruiz, CO2 methanation on TiO2 catalyst, Catal. Sci. Technol. 3 (2013) 2627–2633, https://doi.org/10.1039/
Rh/γ-Al2O3 catalyst at low temperature: “In situ” supply of hydrogen by Ni/ c3cy00355h.
activated carbon catalyst, Appl. Catal. B Environ. 125 (2012) 41–50, https://doi. [31] R. Zhou, N. Rui, Z. Fan, C. jun Liu, Effect of the structure of Ni/TiO2 catalyst on
org/10.1016/j.apcatb.2012.05.019. CO2 methanation, Int. J. Hydrogen Energy 41 (2016) 22017–22025, https://doi.
[8] A. Karelovic, P. Ruiz, Mechanistic study of low temperature CO2 methanation over org/10.1016/j.ijhydene.2016.08.093.
Rh/TiO2 catalysts, J. Catal. 301 (2013) 141–153, https://doi.org/10.1016/j. [32] C. Jia, Y. Dai, Y. Yang, J.W. Chew, Nickel-cobalt catalyst supported on TiO2-coated
jcat.2013.02.009. SiO2 spheres for CO2 methanation in a fluidized bed, Int. J. Hydrogen Energy 44
[9] P. Riani, G. Garbarino, M.A. Lucchini, F. Canepa, G. Busca, Unsupported versus (2019) 13443–13455, https://doi.org/10.1016/j.ijhydene.2019.04.009.
alumina-supported Ni nanoparticles as catalysts for steam/ethanol conversion and [33] A. Makdee, K.C. Chanapattharapol, P. Kidkhunthod, Y. Poo-Arporn, T. Ohno, The
CO2 methanation, J. Mol. Catal. A Chem. 383–384 (2014) 10–16, https://doi.org/ role of Ce addition in catalytic activity enhancement of TiO2-supported Ni for CO2
10.1016/j.molcata.2013.11.006. methanation reaction, RSC Adv. 10 (2020) 26952–26971, https://doi.org/
[10] D. Pandey, K. Ray, R. Bhardwaj, S. Bojja, K.V.R. Chary, G. Deo, Promotion of 10.1039/d0ra04934d.
unsupported nickel catalyst using iron for CO2 methanation, Int. J. Hydrogen [34] P. Unwiset, K.C. Chanapattharapol, P. Kidkhunthod, Y. Poo-arporn, B. Ohtani,
Energy 43 (2018) 4987–5000, https://doi.org/10.1016/j.ijhydene.2018.01.144. Catalytic activities of titania-supported nickel for carbon-dioxide methanation,
[11] J. Lin, C. Ma, Q. Wang, Y. Xu, G. Ma, J. Wang, H. Wang, C. Dong, C. Zhang, Chem. Eng. Sci. 228 (2020), 115955, https://doi.org/10.1016/j.ces.2020.115955.
M. Ding, Enhanced low-temperature performance of CO2 methanation over [35] D. Messou, V. Bernardin, F. Meunier, M.B. Ordoño, A. Urakawa, B.F. Machado,
mesoporous Ni/Al2O3-ZrO2 catalysts, Appl. Catal. B Environ. 243 (2019) 262–272, V. Collière, R. Philippe, P. Serp, C. Le Berre, Origin of the synergistic effect between
https://doi.org/10.1016/j.apcatb.2018.10.059. TiO2 crystalline phases in the Ni/TiO2-catalyzed CO2 methanation reaction,
[12] S. Tada, S. Ikeda, N. Shimoda, T. Honma, M. Takahashi, A. Nariyuki, S. Satokawa, J. Catal. 398 (2021) 14–28, https://doi.org/10.1016/j.jcat.2021.04.004.
Sponge Ni catalyst with high activity in CO2 methanation, Int. J. Hydrogen Energy [36] N. Li, X. Zou, M. Liu, L. Wei, Q. Shen, R. Bibi, C. Xu, Q. Ma, J. Zhou, Enhanced
42 (2017) 30126–30134, https://doi.org/10.1016/j.ijhydene.2017.10.138. visible light photocatalytic hydrogenation of CO2 into methane over a Pd/Ce-TiO2
[13] S.K. Ryi, S.W. Lee, K.R. Hwang, J.S. Park, Production of synthetic natural gas by nanocomposition, J. Phys. Chem. C 121 (2017) 25795–25804, https://doi.org/
means of a catalytic nickel membrane, Fuel 94 (2012) 64–69, https://doi.org/ 10.1021/acs.jpcc.7b07298.
10.1016/j.fuel.2011.12.001. [37] Y. Jiang, Z. Jin, C. Chen, W. Duan, B. Liu, X. Chen, F. Yang, J. Guo, Cerium-doped
[14] S. Kikkawa, K. Teramura, H. Asakura, S. Hosokawa, T. Tanaka, Isolated platinum mesoporous-assembled SiO2/P25 nanocomposites with innovative visible-light
atoms in Ni/γ-Al2O3 for selective hydrogenation of CO2 toward CH4, J. Phys. sensitivity for the photocatalytic degradation of organic dyes, RSC Adv. 7 (2017)
Chem. C. 123 (2019) 23446–23454, https://doi.org/10.1021/acs.jpcc.9b03432. 12856–12870, https://doi.org/10.1039/c7ra00191f.
[15] H. Yuan, X. Zhu, J. Han, H. Wang, Q. Ge, Rhenium-promoted selective CO2 [38] M. Tahir, N.A.S. Amin, Indium-doped TiO2 nanoparticles for photocatalytic CO2
methanation on Ni-based catalyst, J. CO2 Util. 26 (2018) 8–18, https://doi.org/ reduction with H2O vapors to CH4, Appl. Catal. B Environ. 162 (2015) 98–109,
10.1016/j.jcou.2018.04.010. https://doi.org/10.1016/j.apcatb.2014.06.037.
[16] B. Miao, S.S.K. Ma, X. Wang, H. Su, S.H. Chan, Catalysis mechanisms of CO2 and [39] S. Rajendran, D. Manoj, K. Raju, D.D. Dionysiou, M. Naushad, F. Gracia, L. Cornejo,
CO methanation, Catal. Sci. Technol. 6 (2016) 4048–4058, https://doi.org/ M.A. Gracia-Pinilla, T. Ahamad, Influence of mesoporous defect induced mixed-
10.1039/c6cy00478d. valent NiO (Ni2+/Ni3+)-TiO2 nanocomposite for non-enzymatic glucose biosensors,
[17] M. Cai, J. Wen, W. Chu, X. Cheng, Z. Li, Methanation of carbon dioxide on Ni/ Sens. Actuators B Chem. 264 (2018) 27–37, https://doi.org/10.1016/j.
ZrO2-Al2O3 catalysts: effects of ZrO2 promoter and preparation method of novel snb.2018.02.165.
ZrO2-Al2O3 carrier, J. Nat. Gas Chem. 20 (2011) 318–324, https://doi.org/ [40] H. Yan, D. Zhang, J. Xu, Y. Lu, Y. Liu, K. Qiu, Y. Zhang, Y. Luo, Solution growth of
10.1016/S1003-9953(10)60187-9. NiO nanosheets supported on Ni foam as high-performance electrodes for
[18] S. Wang, Q. Pan, J. Peng, T. Sun, D. Gao, S. Wang, CO2 methanation on Ni/ supercapacitors, Nanoscale Res. Lett. 9 (2014) 1–7, https://doi.org/10.1186/1556-
Ce0.5Zr0.5O2 catalysts for the production of synthetic natural gas, Fuel Process. 276X-9-424.
Technol. 123 (2014) 166–171, https://doi.org/10.1016/j.fuproc.2014.01.004. [41] Z. Fan, F. Meng, J. Gong, H. Li, Z. Ding, B. Ding, One-step hydrothermal synthesis
[19] J. Zhang, Y. Bai, Q. Zhang, X. Wang, T. Zhang, Y. Tan, Y. Han, Low-temperature of mesoporous Ce-doped anatase TiO2 nanoparticles with enhanced photocatalytic
methanation of syngas in slurry phase over Zr-doped Ni/γ-Al2O3 catalysts prepared activity, J. Mater. Sci. Mater. Electron. 27 (2016) 11866–11872, https://doi.org/
using different methods, Fuel 132 (2014) 211–218, https://doi.org/10.1016/j. 10.1007/s10854-016-5329-0.
fuel.2014.04.085. [42] M. Dosa, M. Piumetti, S. Bensaid, T. Andana, C. Galletti, D. Fino, N. Russo,
[20] F. Ocampo, B. Louis, A.C. Roger, Methanation of carbon dioxide over nickel-based Photocatalytic abatement of volatile organic compounds by TiO2 nanoparticles
Ce0.72Zr0.28O2 mixed oxide catalysts prepared by sol-gel method, Appl. Catal. A doped with either phosphorous or zirconium, Materials 12 (2019), https://doi.org/
Gen. 369 (2009) 90–96, https://doi.org/10.1016/j.apcata.2009.09.005. 10.3390/ma12132121.
[21] M.D. Hernández-Alonso, I. Tejedor-Tejedor, J.M. Coronado, J. Soria, M. [43] V. García, J.J. Fernández, W. Ruíz, F. Mondragón, A. Moreno, Effect of MgO
A. Anderson, Sol-gel preparation of TiO2-ZrO2 thin films supported on glass rings: addition on the basicity of Ni/ZrO2 and on its catalytic activity in carbon dioxide
Influence of phase composition on photocatalytic activity, Thin Solid Films 502 reforming of methane, Catal. Commun. 11 (2009) 240–246, https://doi.org/
(2006) 125–131, https://doi.org/10.1016/j.tsf.2005.07.256. 10.1016/j.catcom.2009.10.003.
[22] A. Kitiyanan, S. Sakulkhaemaruethai, Y. Suzuki, S. Yoshikawa, Structural and [44] V.G. Deshmane, S.L. Owen, R.Y. Abrokwah, D. Kuila, Mesoporous nanocrystalline
photovoltaic properties of binary TiO2-ZrO2 oxides system prepared by sol-gel TiO2 supported metal (Cu, Co, Ni, Pd, Zn, and Sn) catalysts: effect of metal-support
method, Compos. Sci. Technol. 66 (2006) 1259–1265, https://doi.org/10.1016/j. interactions on steam reforming of methanol, J. Mol. Catal. A Chem. 408 (2015)
compscitech.2005.10.035. 202–213, https://doi.org/10.1016/j.molcata.2015.07.023.
[23] A. Mbiri, D.H. Taffa, E. Gatebe, M. Wark, Zirconium doped mesoporous TiO2 [45] Q. Pan, J. Peng, T. Sun, S. Wang, S. Wang, Insight into the reaction route of CO2
multilayer thin films: influence of the zirconium content on the photodegradation methanation: Promotion effect of medium basic sites, Catal. Commun. 45 (2014)
of organic pollutants, Catal. Today 328 (2019) 71–78, https://doi.org/10.1016/j. 74–78, https://doi.org/10.1016/j.catcom.2013.10.034.
cattod.2019.01.043. [46] T. Tangcharoen, W. Klysubun, C. Kongmark, Synthesis of nanocrystalline NiO/ZnO
[24] S. Ranjbar, K. Saberyan, F. Parsayan, A new process for preparation of Zr doped heterostructured composite powders by sol-gel auto combustion method and their
TiO2 nanopowders using APCVS method, Mater. Chem. Phys. 214 (2018) 337–344, characterizations, J. Mol. Struct. 1156 (2018) 524–533, https://doi.org/10.1016/j.
https://doi.org/10.1016/j.matchemphys.2018.04.119. molstruc.2017.12.019.
[25] N. Venkatachalam, M. Palanichamy, B. Arabindoo, V. Murugesan, Enhanced [47] T. Tangcharoen, W. Klysubun, C. Kongmark, Synchrotron X-ray absorption
photocatalytic degradation of 4-chlorophenol by Zr4+ doped nano TiO2, J. Mol. spectroscopy and cation distribution studies of NiAl2O4, CuAl2O4, and ZnAl2O4
Catal. A Chem. 266 (2007) 158–165, https://doi.org/10.1016/j. nanoparticles synthesized by sol-gel auto combustion method, J. Mol. Struct. 1182
molcata.2006.10.051. (2019) 219–229, https://doi.org/10.1016/j.molstruc.2019.01.049.
[26] W. Li, G. Zhang, X. Jiang, Y. Liu, J. Zhu, F. Ding, Z. Liu, X. Guo, C. Song, CO2 [48] C. Wattanawikkam, W. Pecharapa, K.N. Ishihara, X-ray absorption spectroscopy
hydrogenation on unpromoted and M‑promoted Co/TiO2 catalysts (M = Zr, K, Cs): analysis and magnetic properties of M-doped TiO2 nanoparticles (M=Co, Mn, Ni
effects of crystal phase of supports and metal− support interaction on tuning and Zn) prepared by co-precipitation method, Ceram. Int. 43 (2017) S397–S402,
product distribution Wenhui, ACS Catal. 9 (2019) 2739–2751, https://doi.org/ https://doi.org/10.1016/j.ceramint.2017.05.188.
10.1021/acscatal.8b04720. [49] M. Sahoo, A.K. Yadav, S.N. Jha, D. Bhattacharyya, T. Mathews, N.K. Sahoo,
[27] P. Kidkhunthod, Structural studies of advanced functional materials by S. Dash, A.K. Tyagi, Nitrogen location and Ti-O bond distances in pristine and N-
synchrotron-based x-ray absorption spectroscopy: BL5.2 at SLRI, Thailand, Adv. Doped TiO2 anatase thin films by X-ray absorption studies, J. Phys. Chem. C 119
Nat. Sci. Nanosci. Nanotechnol. 8 (2017), https://doi.org/10.1088/2043-6254/ (2015) 17640–17647, https://doi.org/10.1021/acs.jpcc.5b03295.
aa7240. [50] H. Ikeno, M. Krause, T. Höche, C. Patzig, Y. Hu, A. Gawronski, I. Tanaka, C. Rüssel,
[28] W. Klysubun, P. Kidkhunthod, P. Tarawarakarn, P. Sombunchoo, C. Kongmark, Variation of Zr-L2,3 XANES in tetravalent zirconium oxides, J. Phys. Condens.
S. Limpijumnong, S. Rujirawat, R. Yimnirun, G. Tumcharern, K. Faungnawakij, Matter 25 (2013), https://doi.org/10.1088/0953-8984/25/16/165505.
SUT-NANOTEC-SLRI beamline for X-ray absorption spectroscopy, J. Synchrotron [51] H.Y.T. Chen, S. Tosoni, G. Pacchioni, A DFT study of the acid–base properties of
Radiat. 24 (2017) 707–716, https://doi.org/10.1107/S1600577517004830. anatase TiO2 and tetragonal ZrO2 by adsorption of CO and CO2 probe molecules,
[29] S. Tada, O.J. Ochieng, R. Kikuchi, T. Haneda, H. Kameyama, Promotion of CO2 Surf. Sci. 652 (2016) 163–171, https://doi.org/10.1016/j.susc.2016.02.008.
methanation activity and CH4 selectivity at low temperatures over Ru/CeO2/Al2O3 [52] S. Tada, T. Shimizu, H. Kameyama, T. Haneda, R. Kikuchi, Ni/CeO2 catalysts with
catalysts, Int. J. Hydrogen Energy 39 (2014) 10090–10100, https://doi.org/ high CO2 methanation activity and high CH4 selectivity at low temperatures, Int. J.
10.1016/j.ijhydene.2014.04.133. Hydrogen Energy 37 (2012) 5527–5531, https://doi.org/10.1016/j.
[30] J. Liu, C. Li, F. Wang, S. He, H. Chen, Y. Zhao, M. Wei, D.G. Evans, X. Duan, ijhydene.2011.12.122.
Enhanced low-temperature activity of CO2 methanation over highly-dispersed Ni/

14

You might also like