Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Applied Surface Science 581 (2022) 152327

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

Mn doped CeO2-MoO3/γ-Al2O3 catalysts for the enhanced adsorption and


catalytic oxidation of Hg0 in oxygen atmosphere
Pengfei Cao a, b, Yan Zhang c, *, FengJuan Song b, c, Haitao Zhao d, e, Cheng Heng Pang b, e,
Tao Wu b, e, *
a
School of Chemical Engineering and Technology, Xi’an Jiaotong University, Xi’an 710049, China
b
Key Laboratory of Carbonaceous Wastes Processing and Process Intensification of Zhejiang Province, University of Nottingham Ningbo China, Ningbo 315100, China
c
College of Safety and Environmental Engineering, Shandong University of Science and Technology, Qingdao 266000, China
d
Materials Interfaces Center, Shenzhen Institutes of Advanced Technology, Chinese Academy of Sciences, Shenzhen 518055, Guangdong, China
e
Municipal Key Laboratory of Clean Energy Conversion Technologies, The University of Nottingham Ningbo China, Ningbo 315100, China

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

Keywords: A series of Mn-doped CeO2-MoO3/γ-Al2O3 catalysts were synthesized for the adsorption and catalytic oxidation
Elemental mercury of Hg0 from coal-fired flue gas using oxygen as the oxidizing agent. The results showed that the addition of Mn
Catalytic oxidation significantly promoted the adsorption and catalytic performances of CeO2-MoO3/γ-Al2O3 catalysts. The 6 wt%
Flue gas
Mn-doped catalysts exhibited the best Hg0 catalytic oxidation efficiency of 86%. The catalysts were characterized
Coal-fired power plant
Adsorption
to illustrate the structure-activity relationship of Mn-doped catalysts. It is found that the addition of Mn to the
CeO2 and MoO3 lattice increases the concentration of oxygen vacancies on the catalyst surface which promotes
the catalytic oxidation of mercury. HRTEM results indicate that the incorporation of Mn cracked CeO2 crystal
plane with an angle of 105◦ and may generate more active sites at the interface. Besides, the in-situ DRIFT
spectra show that the acidic sites on the Mn-doped catalyst surface were dominated by Brønsted acid sites at
lower temperatures which play an essential role in the adsorption of mercury. Overall, This research reveals the
effect of Mn additives on the promotion of adsorption capacity and catalytic oxidation efficiency of Hg0,
providing a promising approach to widen the effective temperature window of the catalyst.

1. Introduction matured methods in the application of Hg0 removal and has been widely
adopted. In terms of this method, the metal oxide catalysts have recently
Elemental mercury (Hg0) is toxic and presents a persistent threat to attracted more attention, and they are mainly divided into selective
the ecosystem and human health [1,2]. The emission of elemental catalytic reduction (SCR) catalysts and transition metal oxide catalysts
mercury from coal combustion processes in coal-fired power plants has [14,15]. The traditional SCR catalysts can remove Hg0 and NOx simul­
gained increasing attention in recent years because it causes a series of taneously, but they faced the drawbacks of low mercury oxidation ef­
severe environmental issues, one of which is the biomagnification of Hg0 ficiency without HCl, and high-temperature deactivation [16,17]. The
in aquatic food chains [3]. It was reported that coal-fired power plants Hg0 oxidation efficiency of traditional commercial SCR catalysts is only
contributed to about one-third of anthropogenic mercury emissions 65% [18]. In power plants, the concentration of HCl or Cl2 in real flue
[4,5]. Therefore, the management of mercury emissions from coal-fired gas is usually lower than the reaction needs [19]. On the contrary, the O2
flue gas is crucial in the control of anthropogenic mercury emissions. content of coal-fired flue gas in power plants is about 5%, which is a
Generally, elemental mercury in coal-fired flue gas is hard to be suitable and sufficient oxidant for Hg0 removal. It was reported that
captured by the existing air pollution control devices because it is vol­ transition metal oxide catalysts could use O2 as the oxidizing agent for
atile and insoluble in water [5–8]. With more strict legislation on Hg0 Hg0 removal [20]. Thus, the development of novel environmental-
emission [9,10], many strategies have been investigated for Hg0 cap­ friendly transition metal oxide catalysts with a wide temperature win­
ture, including adsorbent injection [11], catalytic oxidation [12], and dow is a promising approach to achieve cost-effective Hg0 removal using
photochemical oxidation [13], etc. Catalytic-oxidation is one of the most oxygen in flue gas as the oxidizing agent.

* Corresponding authors.
E-mail addresses: tougaoktz@183.com (Y. Zhang), tao.wu@nottingham.edu.cn (T. Wu).

https://doi.org/10.1016/j.apsusc.2021.152327
Received 26 October 2021; Received in revised form 14 December 2021; Accepted 22 December 2021
Available online 27 December 2021
0169-4332/© 2022 Elsevier B.V. All rights reserved.
P. Cao et al. Applied Surface Science 581 (2022) 152327

Recently, various transition metal oxide catalysts such as V2O5[21], deionized water and then impregnated on γ-Al2O3 support via IWI
MnOx [22], FeOx [23], CoOx [24], CeOx [25], and CuO [26] have been method, the well-impregnated solid solution was then dried in an oven
widely studied for Hg0 adsorption and oxidation due to their low cost, operated at 120 ◦ C for 24 h and calcined in air at 520 ◦ C for 12 h. Ce
high catalytic oxidation reaction activity, and good thermal stability. In (NO3)3⋅6H2O and Mn(NO3)2⋅4H2O precursors were then impregnated
our previous research, a high throughput screen of metal oxide catalysts onto the support in the same way. As the loading of Ce and Mo is 2 wt%
was performed by using a dynamic Hg0-TPSR [27]. Among these cata­ and 8 wt%, CeO2-MoO3/γ-Al2O3 catalysts are assigned as 2Ce8Mo/
lysts, Mn-based catalysts after the modification of rare-earth metal ox­ γ-Al2O3. In this research, the 2Ce8Mo/γ-Al2O3 catalysts with different
ides have been tested, showed a better performance in Hg0 adsorption Mn weight ratios were prepared, named 2Ce8MoXMn/γ-Al2O3 (X =
[22,27,28]. CeO2 was found to have promotion effects on the perfor­ 0–8). The detailed procedure for sample preparation was described
mance of Mn-based catalysts due to its high oxygen storage capacity and elsewhere [37].
redox activity [29–31]. In addition, the synergy between Mn and Ce may
accelerate the transfer of electrons and oxygen, thus, further increasing 2.2. Hg0 removal experiments
the catalytic activity [16]. It was reported that the Mo additive had a
promotional effect on exposuring CeO2 (1 1 1) surface which boost the As shown in the schematic diagram of the experimental system in
catalytic oxidation of Hg0 [20,32]. Besides, molybdenum (Mo) doping as Fig. 1, the mercury capture performance of the catalyst was tested on the
a promoter can improve the catalyst’s crystal dispersion and catalytic designed experimental device. The Hg0 was supplied by a mercury
Hg0 removal performance [33]. However, no detailed studies were generator (Tekran 3310, USA), which could generate Hg0 in N2 at an
carried out using MnOx-CeO2-MoO3 as catalysts to enhance the effi­ accurately controlled level. Hg0 (30 µg/m3) was mixed with O2 (5%)
ciency of removing Hg0. Regarding the cost effective of the catalysts, balanced in N2. The catalyst was heated from 25 to 700 ◦ C at a heating
γ-Al2O3, which is usually used as a support for Hg0 removal [22,23], is rate of 1oC/min for dynamic-state analysis. The concentration of the
selected as the catalyst support in this work due to its outstanding me­ total amount of mercury (HgT), Hg0, and Hg2+ (Hg2+ = [HgT] – [Hg0],
chanical properties, good thermal stability, and lower price [34–36]. [38]) were recorded continuously by using the mercury analysis system
Hereby, a series of MnOx-CeO2-MoO3/γ-Al2O3 catalysts with (Tekran 3300RS, USA). Therefore, the efficiency of Hg0 oxidized to Hg2+
different Mn-doped ratios were synthesized by the incipient wetness for each prepared catalyst can be determined. The maximum instant
impregnation (IWI) method for Hg0 removal. Comprehensive charac­ mercury removal efficiency (ΔXmax) and catalytic oxidation ratio could
terization and performance testing were conducted to investigate the be obtained by using the following formula:
effects of Mn addition on the adsorption and catalytic oxidation of Hg0.
Also, the structure–activity relationship between Mn-doped CeO2- [Hg0 ]in − [Hg0 ]out
ΔXmax = × 100% (1)
MoO3/γ-Al2O3 catalyst and Hg0 removal was elucidated. [Hg0 ]in

2. Material and methods Hg2+


Catalytic Oxidation Ratio = × 100% (2)
[Hg0 ]in
2.1. Preparation of catalysts
2.3. Characterization of catalysts
Mn-doped CeO2-MoO3/γ-Al2O3 catalysts with different loading were
synthesized by a three-step IWI method using (NH4)6Mo7O24⋅4H2O, Ce The specific BET surface area, pore volume, and average pore size of
(NO3)3⋅6H2O, and Mn(NO3)2⋅4H2O (Sino-pharm Chemical Reagent Co, samples prepared were determined N2 adsorption-desorption at − 196 ◦ C
Ltd) as the precursors. The (NH4)6Mo7O24⋅4H2O was dissolved in using Micromeritics ASAP 2020 [39]. Before the test, all samples were
degassed at 200 ◦ C for 6 h to remove moisture and adsorbed gases under

Fig. 1. Schematic diagram of the experimental system.

2
P. Cao et al. Applied Surface Science 581 (2022) 152327

a high vacuum. The Brunauer-Emmett-Teller (BET) method was catalyst spectra). The adsorption process was carried out until the
employed to calculate the surface area, and the pore size distribution adsorption of NH3 reached a steady state, after which a desorption
was estimated utilizing the Barret-Joyner-Halenda (BJH) with the process with N2 purging was conducted.
desorption branches. To determine the crystal phase structure of the
samples, X-Ray diffraction (XRD, Bruker D8 A25, Germany) with Cu Kα 3. Results and discussion
radiation (λ = 1.5406 Å) was conducted at 2θ of 20–80◦ at the rate of
0.067◦ /sec [40]. Transmission electron microscope (TEM) and high- 3.1. Hg0 adsorption and catalytic oxidation
resolution transmission electron microscopy (HRTEM) images were
captured using a JEM 2100 microscope operated at 200 Kv, the sample 3.1.1. Qualitative analysis for Hg0 removal performance.
was dispersed into ethanol assisted by ultra-sonication, and then drop­ The Hg0 temperature-programmed surface reaction (Hg0-TPSR) was
ped on copper grid support and dried in air before analysis [37]. X-Ray used to qualitative and quantitatively analyze the mercury adsorption
photoelectron spectroscopy (XPS) analysis was carried out using a efficiency and catalytic oxidation characteristics of 2Ce8Mo/γ-Al2O3
Kratos AXIS Ultra DLD spectrometer with a monochromatic Al Kα (hv = catalysts doped with different loading of Mn. The dynamic transient-
1486.7 eV) radiation source under a vacuum pressure of 10− 7 Pa. The state results of Hg0 and Hg2+ are displayed in Fig. 2. Fig. 2(a) shows
surface charging was neutralized by using low-energy electrons and the the adsorption and oxidation profiles for HgT and Hg0 from 25 ◦ C to
pass energy of the electron energy analyzer is 20 eV [41]. The binding 700 ◦ C. Lines with solid and hollow symbols present HgT and Hg0,
energy was calibrated with respect to the C 1s level 284.6 eV of respectively. It could be seen that HgT and Hg0 exhibit a similar
contaminated carbon. The CasaXPS solftware (Version 2.2.73) was used changing trend for all 2Ce8MoXMn/γ-Al2O3 catalysts. However, an
for XPS peaks fitting. The samples were firstly dropped onto tin foil appreciable difference between the concentration of HgT and Hg0 was
paper with a smooth surface. After that, a copper sheet (around 1 × 1 observed when increasing the Mn loading. The catalyst doped with 6 wt
cm2) with sticky carbon tape was pressed into powders on a foil paper, % of Mn exhibited the most significant distinction, but the difference
which then was loaded into a sample chamber for the test. Raman decreased as Mn loaded up to 8 wt%. The conversion profiles of Hg2+ are
spectra were recorded by using a Raman Renishaw RM2000. A 514 nm presented in Fig. 2(b), which were calculated by the difference between
diode laser supplied the pump radiation to obtain the spectral range of HgT and Hg0. The results reveal that 2Ce8Mo/γ-Al2O3 catalysts with 6
1000–2000 cm− 1, and the Raman emission was focused through a 50× wt% Mn loadings are optimal, generating the highest volume of Hg2+ at
objective [9]. In-situ diffuse reflectance infrared Fourier transform the low-temperature range.
spectroscopy (In-situ DRIFTS) was employed to study the surface acidic The characteristic temperatures of Hg0 adsorption and desorption
sites of the catalysts by using NH3 as the alkaline probe molecule [10]. processes were summarized in Table 1, including the initial adsorption
The IR spectra were recorded on a Fourier transform infrared spec­ temperature (Ta0), the adsorption rate peak (Tra, peak), the best effective
trometer (Bruker V70, USA) equipped with a Praying Mantis™ reaction adsorption range (Ta, region), and the initial desorption temperature (Td0).
chamber (Harrick, USA). The catalyst was initially heated to 400 ◦ C in As shown in Table 1, when Mn composition increased from 0 wt% to 8
N2 to remove impurities adsorbed by the materials. The background wt%, Ta0 gradually decreased from 58.8 ◦ C to 27.8 ◦ C. The results
spectrum was collected in N2 atmosphere (and subtracted from the indicate that the addition of Mn contributed to lowering the Ta0. Also,

Fig. 2. Hg0 and HgT TPSR profiles of (a) 2Ce8MoXMn/γ-Al2O3 catalysts with different loading of Mn doping. HgT and Hg0 are represented by the line with a solid
symbol and a hollow symbol, respectively. (b) Hg2+ profiles of 2Ce8MoXMn/γ-Al2O3 catalysts.

3
P. Cao et al. Applied Surface Science 581 (2022) 152327

Table 1 wt%. It shows that the addition of Mn improves the adsorption and
Characteristic temperatures of CeMo-catalyst promoted by Mn. catalytic oxidation effect, which is probably related to the synergy be­
Sample Ta0 Tra,peak Ta,region Td0 tween CeO2 and Mn additives.
(◦ C) (◦ C) (◦ C) (◦ C) Furthermore, a steady-state experiment over the 2Ce8Mo6Mn/
2Ce8Mo 58.8 104.3 342.2 479.4 γ-Al2O3 catalyst was conducted to investigate the effect of the operating
2Ce8Mo2Mn 31.8 61.8 209.2 458.7 temperatures on Hg0 total removal efficiency and the catalytic oxidation
2Ce8Mo4Mn 29.2 52.6 203.1 452.7 efficiency, and the results are shown in Fig. 3. The operating tempera­
2Ce8Mo6Mn 30.3 57.2 191 443.8 tures are within the range of 100–400 ◦ C that is below the initial Hg0
2Ce8Mo8Mn 27.8 60.1 183.6 432.8
desorption temperature. The total removal efficiency of Hg0 was higher
than 80% at the temperature range of 150–350 ◦ C, with the highest
Tra peak decreased with an increase of manganese composition. The efficiency of 93% at 225 ◦ C. Regarding the catalytic oxidation efficiency
adsorption rate of 2Ce8Mo/γ-Al2O3 catalysts reached its maximum at of Hg0, it is lower than 34% at temperatures below 225 ◦ C. However, it is
104.3 ◦ C. After adding Mn, the maximum rate of adsorption was at significantly improved above 80% between 250 and 350 ◦ C. Both Hg0
60 ◦ C. This indicates that the addition of Mn could lower the tempera­ adsorption and catalytic oxidation contribute to the total removal effi­
ture peaks of Ta0 and Tra. Thus, the addition of Mn leads to the ciency of Hg0. On the contrary, the capacity of Hg0 adsorption increases
adsorption of catalysts starting at a lower temperature, which widens with the temperature rise to 225 ◦ C, and then it decreases at a higher
the effective temperature range. The adsorption rate and optimal temperature. A possible explanation is that surface active acidic sites are
effective adsorption range are criteria of the adsorption capacity of not stable at higher temperatures, which is further confirmed by the in
catalysts. Although the adsorption rate peak was not spotted earliest for situ DRIFTS characterization. It can be concluded that Hg0 adsorption
the catalysts doped with 6 wt% of Mn, it can reach the maximum prefers at a lower temperature, while catalytic Hg0 oxidation always
instantaneous adsorption rate (in Fig. 2(a)). occurs at a higher temperature. The results show a good agreement with
Besides, the removal efficiency is the key parameter indicating its the observation from the dynamic-state analysis.
effective adsorption range. It is defined to be the best effective adsorp­
tion range when the removal efficiency is over 80%. Both Hg0 adsorp­ 3.2. Structure characterization
tion and desorption for samples without doping manganese begin at
higher temperatures. The optimal effective adsorption range gradually 3.2.1. XRD crystallographic analysis
shifts toward lower temperatures as the manganese content increases. As shown in Fig. 4(a), the XRD patterns of 2Ce8Mo/γ-Al2O3 catalyst
The results show that the adsorption amount of 6 wt% manganese is the with different Mn doping were investigated to understand each cata­
largest and starts at a lower temperature, while the amount of Hg0 lyst’s crystallization and monolayer coverage. The XRD spectra indicate
adsorption for 0 wt% manganese is the lowest. This suggests that the that two main components in the sample are assigned as γ-Al2O3 and
incorporation of Mn broadens the effective temperature window for CeO2. For all the samples, no peaks of Mn and Mo were observed,
adsorption and catalysis and moves to a lower temperature. implying their good dispersion in γ-Al2O3 or their existence in the
amorphous or low crystalline state. The diffraction peaks at 37◦ , 46◦ ,
3.1.2. Quantitative analysis of Hg0 catalytic oxidation and 67◦ are correspond to the phase of γ-Al2O3. As the amount of Mn
The characteristic of mercury catalytic oxidation with respect to increases, the peak of γ-Al2O3 gradually decreases. It could be attributed
catalysts were quantitatively analyzed according to The Hg0-TPSR re­ to the strong interaction between the support and the active metal oxide.
sults of the 2Ce8Mo/γ-Al2O3 catalysts promoted by various composi­ Ce peaks were detected at 28.6◦ , 47.5◦ , and 56.3◦ , corresponding to
tions of Mn. The catalytic capacity study of these catalyst samples had (1 1 1), (2 0 0), and (3 1 1) planes of CeO2 (JCPDS No. 34-0394), respec­
been implemented by integrating the TPSR results to compute the tively. It is found that the diffraction peak of CeO2 gradually decreases
amount of Hg2+. The result is shown in Table 2. The adsorption area (Sa) and shifts to a higher Bragg angle when increases the manganese
and desorption area (Sd) were calculated by integrating results in Fig. 2 loading. Due to the interaction between the two metal oxides, the
(a), representing the amount of Hg0 adsorbed and desorbed, respec­ incorporated Mnx+ cations into the CeO2 lattice leads to a decrease in
tively. From Table 2, it is clear that the adsorption amount of Hg in
2Ce8Mo6Mn/γ-Al2O3 is more than twice that in 2Ce8Mo/γ-Al2O3. When
the experimental temperature increased to about 700 ◦ C, the residual
Hg0 in the samples was negligible because of its total desorption.
Therefore, the significant difference between Sa and Sd was considered
as Hg2+ formation by catalytic oxidation. Another alternative calcula­
tion of Hg2+ was to integrate the results from Fig. 2(b) directly. Two sets
of results have shown the same trend, where the sample with 6 wt% Mn
showed the best catalytic capacity, followed by the catalyst with 8 wt%
Mn. Sa and SHg2+ have significantly increased after the addition of Mn.
Furthermore, 2Ce8Mo6Mn/γ-Al2O3 catalysts show the highest effi­
ciency of Hg adsorption and catalytic oxidation. In particular, the cat­
alytic oxidation efficiency reached 86% when the Mn doping rate was 6

Table 2
Characteristic areas of CeMo-catalyst promoted by Mn.
Sample Sa Sd Sa-Sd SHg2+
(min μg/m3) (min μg/m3) (min μg/m3) (min μg/m3)

2Ce8Mo 1748.9 318.9 1430.7 1424.6


2Ce8Mo2Mn 2807.6 375.79 2431.9 2424.5
2Ce8Mo4Mn 3030.8 590.6 2440.3 2436.2
2Ce8Mo6Mn 3898.6 514.9 3383.6 3374.8
Fig. 3. Steady-state analysis of total removal efficiency and catalytic oxidation
2Ce8Mo8Mn 3642.9 574.8 3068.1 3067.0
efficiency of Hg0 on the 2Ce8Mo6Mn/γ-Al2O3 catalyst.

4
P. Cao et al. Applied Surface Science 581 (2022) 152327

Fig. 4. (a) XRD patterns of 2Ce8MoXMn/γ-Al2O3 catalysts with ranges of Mn loadings. (b) XRD semi-quantitative analysis of 2Ce8Mo/γ-Al2O3 and 2Ce8MoXMn/
γ-Al2O3 and catalysts.

lattice parameters [42]. Additionally, the synergistic effect between Mn


Table 3
and Mo increases the diffusion effect, resulting in higher single layer
The surface properties of Ce-Mo catalysts with different ratios of Mn.
coverage [38]. Fig. 4(b) shows the semi-quantitative analysis of CeO2
(1 1 1) crystallites modified by different Mn loadings. As we can see that Catalysts Surface area(m2/ Total pore volume Average pore width
g) (cm3/g) (Å)
the peak intensity of CeO2 gradually decreases with the introduction of
MnOx. This indicates that the doping of MnOx reduces the crystallinity of γ-Al2O3 242.2 0.48 79.7
2Ce8Mo 173.6 0.43 100.1
CeO2 but improves its dispersibility.
2Ce8Mo2Mn 180.9 0.42 93.8
2Ce8Mo4Mn 178.3 0.42 94.2
3.2.2. BET surface area and pore structure 2Ce8Mo6Mn 165.2 0.39 94.7
The effect of modified Mo-based catalysts with different Mn ratios on 2Ce8Mo8Mn 162.2 0.39 95.8
the surface area, pore volumes, and pore width of catalysts was further
evaluated. The results for the 2Ce8Mo/γ-Al2O3 and 2Ce8MoXMn/
γ-Al2O3 catalysts are displayed in Table 3. Obviously, pure γ-Al2O3 has

5
P. Cao et al. Applied Surface Science 581 (2022) 152327

the largest specific surface area and pore volume. As shown in Table 3, a
slight rise was found when 2 wt% Mn was added to the 2Ce8Mo/γ-Al2O3
catalysts. However, when the loading of Mn was further increased, the
surface area of BET began to decrease gradually. This may be due to the
phase separation of massive MnOx when more Mn was doped into CeO2
[43]. A similar trend was observed on the specific surface area and pore
volume of the sample prepared by the impregnation process. This phe­
nomenon is mainly because the metal oxide is impregnated onto the
γ-Al2O3 surface and blocks its pores resulting in a decrease in surface
area and total pore volume. Moreover, the average pore width shifted to
slightly higher values indicating that the small pores are filled. Also, this
phenomenon demonstrates that the growth and agglomeration of
nanoparticles lead to the formation of a crystal with a high loading of
manganese. Moreover, the specific surface area of the crystalline
structure is usually smaller than that of the amorphous structure.
Although the BET of the sample decreased with the addition of man­
ganese, the activity of the catalyst increased. It suggests that surface area
is not a key factor affecting mercury catalysis.

3.2.3. XPS valence states analysis


The elemental valence state of the surface of the 2Ce8Mo6Mn/
γ-Al2O3 catalyst was verified by XPS spectral analysis. Fig. 5 shows the
distracted XPS spectra of Ce 3d, Mn 2p, and Mo 3d. The XPS map of the
tested and fitted Ce 3d is displayed in Fig. 5(a). Two sets of spin–orbit
multiples u(u-u′′′ ) and v(v-v′′′ ) are observed corresponding to Ce 3d3/2
and Ce 3d5/2, respectively. Among them, u’ and v’ peaks were attributed
to Ce3+, while the remaining six peaks refer to Ce4+. It is obvious that the
proportion of Ce3+ in the catalyst with added Mn is significantly
enhanced. As a result, the redox performance of catalysts is enhanced as
the ratio of Ce3+ is proportional to oxygen vacancy [42]. It indicated
that Mnx+ had been incorporated into the CeO2 lattice resulting in a
change in the Ce3+ ratio. This is consistent with the XRD results. Ce3+
could create an unbalanced charge, vacancy, and unsaturated chemical
bond on the catalyst surface, which is conducive to the attachment of
chemical adsorption oxygen on the catalyst surface. Thus, the redox
conversion between Ce3+ and Ce4+ could improve the removal rate of
mercury [44].
Fig. 5(b) shows the XPS spectrum of the Mn 2p. The two prominent
peaks’ positions were approximately 654 eV and 643 eV, corresponding
to Mn2p1/2 and Mn2p3/2, respectively. As shown in Fig. 5(b), both Mn
2p1/2 and 2p3/2 spectra could be divided into two sub-peaks. The peaks
of 641.5 eV and 653.1 eV are attributed to Mn3+, the peaks of 642.7 eV
and 654.1 eV are assigned to Mn4+. The fitted results show that the
valence state of manganese in catalyst 2Ce8Mo6Mn/γ-Al2O3 is mainly
Mn4+. Previous studies [16] reported that Mn4+ promotes the adsorp­
tion and oxidation of Hg0 as it adsorbs electrons in the oxygen vacancy
to form Mn3+. Fig. 5(c) also shows the two prominent Mo peaks corre­
sponding to the photoelectron lines of Mo3d3/2 and 3d5/2. For the
catalyst 2Ce8M/γ-Al2O3, peaks centred on 235.6 eV and 232.5 eV [45]
refer to Mo6+. The peaks of 236.1 eV [46] and 233.1 eV [47] for catalyst
2Ce8Mo6Mn/γ-Al2O3 are attributed to Mo6+ as well. All peaks attrib­
Fig. 5. The XPS spectra of (a) Ce3d, (b) Mn2p and (c) Mo3d of 2Ce8Mo/
uted to Mo6+ indicate that the high oxygen concentration on the catalyst
γ-Al2O3 and 2Ce8Mo6Mn/γ-Al2O3 catalysts.
surface may be beneficial to the oxidation of mercury. In Fig. 5(c), after
the addition of Mn, there is a shift for Mo6+ peaks of the photoelectron
lines of Mo3d3/2 and 3d5/2 toward higher binding energy. It can be nanoparticles are observed, which formed layer structure (in Fig. 6(b)).
explained that Mn doped into the lattice of MoO3 resulting in the for­ The HRTEM images of Fig. 6(c) clearly show a well-crystallized (1 1 1)
mation of Mn-Mo-O [48,49]. Meanwhile, the lattice distortion induces lattice plane with a lattice spacing of 0.31 nm, demonstrating that the
the oxygen vacancies [50,51], which enhances the catalytic oxidation of crystal structure of CeO2 is fluorite-type in the samples [42]. The results
Hg0 [52]. are consistent with our previous research [46]. After the incorporation
of MnOx, the lattice spacing of (1 1 1) plane in Fig. 6(d) slightly reduced
3.2.4. TEM and high-resolution TEM (HRTEM) morphology study from 0.31 nm to 0.30 nm. This indicates that Mnx+ cations with a smaller
To investigate the morphology and structural characteristics of the ion radius have been incorporated into the cerium lattice to cause lattice
catalyst, 2Ce8Mo/γ-Al2O3 and 2Ce8Mo6Mn/γ-Al2O3 catalysts were distortion. No molybdenum oxides with a clear lattice spacing were
characterized by TEM and HRTEM. Without Mn additives, the 2Ce8Mo/ found in the HRTEM image, it might be because MoO3 exists mainly in
γ-Al2O3 catalyst is consists of nanoparticles that formed a porous an amorphous structure. Furthermore, compared with the complete
structure (in Fig. 6(a)). While for Mn-doped catalyst, flakes instead of CeO2 crystal plane in Fig. 4(c) and (d) CeO2 crystal plane has cracked,

6
P. Cao et al. Applied Surface Science 581 (2022) 152327

Fig. 6. TEM image of (a) 2Ce8Mo/γ-Al2O3 catalyst, (b) 2Ce8Mo6Mn/γ-Al2O3 catalyst; HRTEM image of (c) 2Ce8Mo/γ-Al2O3 and (d) 2Ce8Mo6Mn/γ-Al2O3; top view
of structural configuration of the CeO2 (1 1 1) surface of (e) 2Ce8Mo/γ-Al2O3 and (f) 2Ce8Mo6Mn/γ-Al2O3.

which may generate more active sites at the interface to improve the
catalytic oxidation effect of Hg0 [53]. Fig. 6(e) and (f) are simulated
microstructures diagrams of the (1 1 1) plane of Ce. It gives evidence that
the cracking angle of CeO2 crystal planes of 2Ce8Mo6Mn/γ-Al2O3 is
105◦ .

3.3. Raman spectroscopy surface study

Raman spectrum is an effective way to examine the variations in


vibrational and rotational modes associated with 2Ce8Mo/γ-Al2O3 and
2Ce8Mo6Mn/γ-Al2O3 catalyst, of which the results presented in Fig. 7.
When there is no MnOx is incorporated, the Raman spectrum of the
2Ce8Mo/γ-Al2O3 catalyst displays a weak peak and a strong peak at
positions of about 300 cm− 1 and 950 cm− 1, respectively. Both peaks
refer to MoO3, which corresponds to the stretching vibration of the
Mo–– O bond. When doped 6 wt% of MnOx on the catalysts, the peak
intensity of MoO3 decreased significantly, but a new peak appeared
around 600 cm− 1 attributed to MnOx [38]. Additionally, two peaks Fig. 7. The Raman spectra of 2Ce8Mo/γ-Al2O3 and 2Ce8Mo6Mn/
located at 920 cm− 1 and 990 cm− 1 were detected, indicating that the Mn γ-Al2O3 catalysts.
cations dope into the lattice of MoO3 to form Mo-Mn-O solid solution
scattered on the alumina carrier [42]. This result shows an agreement

7
P. Cao et al. Applied Surface Science 581 (2022) 152327

with XPS results. The incorporation of Mnx+ in MoO3 lattice leads to the
formation of oxygen vacancy, promoting the catalytic oxidation of Hg0.
Even though the peaks corresponding to manganese and molybdenum
are weak and not apparent in XRD patterns, they are detected by Raman
spectrum analysis. It suggests that MnOx and MoO3 components are well
dispersed on the surface of the catalyst. There are also no CeO2 and
Al2O3 peaks detected here which is consistent with XRD results.

3.4. In situ DRIFT characterization of surface acid sites

The properties of the surface acidic sites on the catalysts were


analyzed by in situ DRIFT analysis using NH3 as a molecular probe.
Previous literature has shown that the removal efficiency of Hg0 posi­
tively correlated with the number of Brønsted acid sites on the catalyst
surface [37,54]. To further understand the effect of acidic sites on Hg0
adsorption and catalytic oxidation, the in-situ NH3-DRIFT on
2Ce8Mo6Mn/γ-Al2O3 catalyst was performed at different temperatures.
As shown in Fig. 8 (a), the peaks located around 1680 and 1460 cm− 1
were attributed to the coordinated NH4+ on the Brønsted acid sites,
while the peaks at 1600 and 1230 cm− 1 were attributed to the adsorp­
tion of NH3 on the Lewis acid sites. As the temperature increase, the
amount of both Brønsted acid and Lewis acid sites perform an apparent
decrease, but the decline rate of Brønsted acid is more significant. It
reveals that Brønsted acid sites are not thermal stability at high tem­
peratures. Hg0-TPSR results show that Hg0 adsorption mainly occurs at
temperatures below 200℃, and catalytic oxidation starts at higher
temperatures. Moreover, the capacity of Hg0 adsorption decreases when
increases the temperatures, which performs a similar trend to the in-situ
NH3-DRIFT results. This result indicates that Brønsted acid on the
catalyst surface significantly affects Hg0 adsorption other than catalytic
oxidation. More Brønsted acid on the catalyst surface, higher capacity of
Hg0 adsorption. Fig. 8(b) presents a semi-quantitative analysis of
Brønsted acid and Lewis acid, including the integrated area of Fig. 8(a)
and the ratio of Brønsted acid and Lewis acid (B/L). The Brønsted acid Fig. 8. In suit DRIFT study of surface acid sites of 2Ce8Mo6Mn/
accounts for a large proportion at temperatures below 200 ◦ C, the value γ-Al2O3 catalysts.
of B/L gradually decreases with the temperature rising. It confirms that
Hg0 adsorption is positively correlated with the amount of Brønsted acid CRediT authorship contribution statement
sites more than Lewis acid sites on the catalyst surface. Additionally, Mn
additives increase the amount of Brønsted acid, which lowers the tem­ Pengfei Cao: Conceptualization, Methodology, Software, Formal
perature of Hg0 absorption during the Hg0-TPSR analysis. analysis, Investigation, Data curation, Writing – original draft. Yan
Zhang: Data curation, Validation, Writing – review & editing. Feng­
4. Conclusions Juan Song: Methodology, Investigation, Data curation. Haitao Zhao:
Data curation, Validation, Writing – review & editing. Cheng Heng
The 2Ce8MoXMn/γ-Al2O3 catalysts exhibited an excellent capacity Pang: Validation, Writing – review & editing. Wu Tao: Supervision,
for Hg0 removal at lower temperatures under O2 atmosphere. Among Resources, Writing – review & editing, Project administration.
them, 2Ce8Mo6Mn/γ-Al2O3 was found as the most efficient one (around
86% Hg0 oxidation efficiency under 5% O2). Mn incorporated catalysts Declaration of Competing Interest
perform better Hg0 removal performances, of which the efficiency is
more than twice of the 2Ce8Mo/γ-Al2O3 catalyst. It is found that Mnx+ The authors declare that they have no known competing financial
cations doped into CeO2 lattice resulted in lattice distortion. The higher interests or personal relationships that could have appeared to influence
oxidation efficiency is attributed to the increased concentration of ox­ the work reported in this paper.
ygen vacancies and the strong interaction between CeO2 and Mn. The
reason is that when a manganese ion substitutes the Ce4+ ion, an oxygen Acknowledgments
vacancy is generated to maintain the charge balance. It demonstrated
that the addition of manganese widens the effective temperature win­ Following funding bodies are acknowledged for partially sponsoring
dow and lowers the Hg0 removal temperature. In general, Mn-doped this research: National Key R&D Program of China (2017YFB0602602
2Ce8Mo/γ-Al2O3 catalysts have greatly improved on adsorption and and 2017YFC0210400) and National Natural Science Foundation of
catalytic oxidation performance. Specifically, the catalytic oxidation China (51606106, 52074176, 52173234 and 22008191), Shandong
efficiency reached 86% and the total adsorption capacity also increased Society for Environmental Science (202017), Ningbo Natural Science
significantly. This is mainly due to the synergistic effect between CeO2 Foundation (2017A610233) and Shaanxi Natural Science Foundation
and Mn. It is concluded that the doping of manganese leads to an in­ (2020JQ-014). Authors 1 and 2 contributed equally to this work.
crease in the concentration of oxygen vacancies and an increase in the
number of surface active sites, thereby improving the efficiency of the
catalytic oxidation and adsorption of mercury.

8
P. Cao et al. Applied Surface Science 581 (2022) 152327

References [28] H. Xu, Z. Qu, C. Zong, F. Quan, J. Mei, N. Yan, Catalytic oxidation and adsorption
of Hg0 over low-temperature NH3-SCR LaMnO3 perovskite oxide from flue gas,
Appl. Catal. B 186 (2016) 30–40.
[1] Y. Wang, B. Shen, C. He, S. Yue, F. Wang, Simultaneous Removal of NO and Hg(0)
[29] R. Jin, Y. Liu, Z. Wu, H. Wang, T. Gu, Low-temperature selective catalytic
from Flue Gas over Mn-Ce/Ti-PILCs, Environ. Sci. Technol. 49 (15) (2015)
reduction of NO with NH3 over MnCe oxides supported on TiO2 and Al2O3: a
9355–9363.
comparative study, Chemosphere 78 (9) (2010) 1160–1166.
[2] J. Yang, Y. Zhao, J. Zhang, C. Zheng, Removal of elemental mercury from flue gas
[30] G. Qi, R.T. Yang, Low-temperature selective catalytic reduction of NO with NH3
by recyclable CuCl2 modified magnetospheres catalyst from fly ash. Part 1. Catalyst
over iron and manganese oxides supported on titania, Appl. Catal. B 44 (3) (2003)
characterization and performance evaluation, Fuel 164 (2016) 419–428.
217–225.
[3] L. Zhang, M.H. Wong, Environmental mercury contamination in China: sources and
[31] N.-Z. Yang, R.-T. Guo, W.-G. Pan, Q.-L. Chen, Q.-S. Wang, C.-z. Lu, The promotion
impacts, Environ. Int. 33 (1) (2007) 108–121.
effect of Sb on the Na resistance of Mn/TiO2 catalyst for selective catalytic
[4] D. Jampaiah, K.M. Tur, P. Venkataswamy, S.J. Ippolito, Y.M. Sabri, J. Tardio, S.
reduction of NO with NH3, Fuel 169 (2016) 87–92.
K. Bhargava, B.M. Reddy, Catalytic oxidation and adsorption of elemental mercury
[32] H. Zhao, et al., Co-regulation of dispersion, exposure and defect sites on CeO2 (111)
over nanostructured CeO2–MnOx catalyst, RSC Adv. 5 (38) (2015) 30331–30341.
surface for catalytic oxidation of Hg0, J. Hazard. Mater. 424 (2022), 126566.
[5] A.A. Presto, E.J. Granite, Survey of catalysts for oxidation of mercury in flue gas,
[33] Q. Wang, L. Huang, X. Quan, G. Li Puma, Sequential anaerobic and electro-Fenton
Environ. Sci. Technol. 40 (18) (2006) 5601–5609.
processes mediated by W and Mo oxides for degradation/mineralization of azo dye
[6] P. Wang, S. Su, J. Xiang, F. Cao, L. Sun, S. Hu, S. Lei, Catalytic oxidation of Hg0 by
methyl orange in photo assisted microbial fuel cells, Appl. Catal. B 245 (2019)
CuO–MnO2–Fe2O3/γ-Al2O3 catalyst, Chem. Eng. J. 225 (2013) 68–75.
672–680.
[7] X. Li, L. Zhang, D. Zhou, W. Liu, X. Zhu, Y. Xu, Y. Zheng, C. Zheng, Elemental
[34] W. Hou, J. Zhou, C. Yu, S. You, X. Gao, Z. Luo, Pd/Al2O3 sorbents for elemental
mercury capture from flue gas by a supported ionic liquid phase adsorbent, Energy
mercury capture at high temperatures in syngas, Ind. Eng. Chem. Res. 53 (23)
Fuels 31 (1) (2017) 714–723.
(2014) 9909–9914.
[8] J. Wang, Y. Zhang, L. Han, L. Chang, W. Bao, Simultaneous removal of hydrogen
[35] W. Du, L. Yin, Y. Zhuo, Q. Xu, L. Zhang, C. Chen, Performance of CuOx–neutral
sulfide and mercury from simulated syngas by iron-based sorbents, Fuel 103 (2013)
Al2O3 sorbents on mercury removal from simulated coal combustion flue gas, Fuel
73–79.
Process. Technol. 131 (2015) 403–408.
[9] H. Zhao, X. Mu, G. Yang, M. George, P. Cao, B. Fanady, S. Rong, X. Gao, T. Wu,
[36] W. Du, L. Yin, Y. Zhuo, Q. Xu, L. Zhang, C. Chen, Catalytic oxidation and
Graphene-like MoS2 containing adsorbents for Hg0 capture at coal-fired power
adsorption of elemental mercury over CuCl2-impregnated sorbents, Ind. Eng.
plants, Appl. Energy 207 (2017) 254–264.
Chem. Res. 53 (2) (2014) 582–591.
[10] G. Yang, H. Zhao, X. Luo, K. Shi, H. Zhao, W. Wang, Q. Chen, H. Fan, T. Wu,
[37] H. Zhao, et al., Hg(0) Capture over CoMoS/gamma-Al2O3 with MoS2 nanosheets at
Promotion effect and mechanism of the addition of Mo on the enhanced low
low temperatures, Environ. Sci. Technol. 50 (2) (2016) 1056–1064.
temperature SCR of NOx by NH3 over MnOx/γ-Al2O3 catalysts, Appl. Catal. B 245
[38] H. Zhao, C.I. Ezeh, S. Yin, Z. Xie, C.H. Pang, C. Zheng, X. Gao, T. Wu, MoO3-
(2019) 743–752.
adjusted δ-MnO2nanosheet for catalytic oxidation of Hg0 to Hg2+, Appl. Catal. B
[11] Z. Tan, L. Sun, J. Xiang, H. Zeng, Z. Liu, S. Hu, J. Qiu, Gas-phase elemental mercury
263 (2020) 117829, https://doi.org/10.1016/j.apcatb.2019.117829.
removal by novel carbon-based sorbents, Carbon 50 (2) (2012) 362–371.
[39] Y. Shao, J. Li, H. Chang, Y. Peng, Y. Deng, The outstanding performance of LDH-
[12] H. Li, C.-Y. Wu, Y. Li, J. Zhang, Superior activity of MnOx-CeO2/TiO2 catalyst for
derived mixed oxide Mn/CoAlOx for Hg0 oxidation, Catal. Sci. Technol. 5 (7)
catalytic oxidation of elemental mercury at low flue gas temperatures, Appl. Catal.
(2015) 3536–3544.
B 111-112 (2012) 381–388.
[40] H. Zhao, X. Luo, J. He, C. Peng, T. Wu, Recovery of elemental sulphur via selective
[13] Y. Yuan, J. Zhang, H. Li, Y. Li, Y. Zhao, C. Zheng, Simultaneous removal of SO2, NO
catalytic reduction of SO2 over sulphided CoMo/γ-Al2O3 catalysts, Fuel 147 (2015)
and mercury using TiO2-aluminum silicate fiber by photocatalysis, Chem. Eng. J.
67–75.
192 (2012) 21–28.
[41] G. Greczynski, L. Hultman, The same chemical state of carbon gives rise to two
[14] D. Jampaiah, S.J. Ippolito, Y.M. Sabri, B.M. Reddy, S.K. Bhargava, Highly efficient
peaks in X-ray photoelectron spectroscopy, Sci. Rep. 11 (1) (2021) 11195.
nanosized Mn and Fe codoped ceria-based solid solutions for elemental mercury
[42] P. Venkataswamy, D. Jampaiah, F. Lin, I. Alxneit, B.M. Reddy, Structural
removal at low flue gas temperatures, Catal. Sci. Technol. 5 (5) (2015) 2913–2924.
properties of alumina supported Ce–Mn solid solutions and their markedly
[15] Z. Liu, Z. Zhou, G. Qi, T. Zhu, Selective catalytic reduction of NOx with NH3 over
enhanced catalytic activity for CO oxidation, Appl. Surf. Sci. 349 (2015) 299–309.
MoO3/Mn-Zr composite oxide catalyst, Appl. Surf. Sci. 466 (2019) 459–465.
[43] Y. Liao, M. Fu, L. Chen, J. Wu, B. Huang, D. Ye, Catalytic oxidation of toluene over
[16] H. Li, Y. Wang, S. Wang, X.u. Wang, J. Hu, Removal of elemental mercury in flue
nanorod-structured Mn–Ce mixed oxides, Catal. Today 216 (2013) 220–228.
gas at lower temperatures over Mn-Ce based materials prepared by co-
[44] H. Li, S. Wang, X.u. Wang, N. Tang, S. Pan, J. Hu, Catalytic oxidation of Hg0 in flue
precipitation, Fuel 208 (2017) 576–586.
gas over Ce modified TiO2 supported Co-Mn catalysts: characterization, the effect
[17] D. Pudasainee, S.J. Lee, S.-H. Lee, J.-H. Kim, H.-N. Jang, S.-J. Cho, Y.-C. Seo, Effect
of gas composition and co-benefit of NO conversion, Fuel 202 (2017) 470–482.
of selective catalytic reactor on oxidation and enhanced removal of mercury in
[45] J.-G. Choi, L.T. Thompson, XPS study of as-prepared and reduced molybdenum
coal-fired power plants, Fuel 89 (4) (2010) 804–809.
oxides, Appl. Surf. Sci. 93 (2) (1996) 143–149.
[18] R. Stolle, H. Koeser, H. Gutberlet, Oxidation and reduction of mercury by SCR
[46] M.R. Smith, U.S. Ozkan, The partial oxidation of methane to formaldehyde: role of
DeNOx catalysts under flue gas conditions in coal fired power plants, Appl. Catal. B
different crystal planes of MoO3, J. Catal. 141 (1) (1993) 124–139.
144 (2014) 486–497.
[47] T.F. Hayden, J.A. Dumesic, Studies of the structure of molybdenum oxide and
[19] H. Zhao, et al., Catalytic oxidation of Hg(0) with O2 induced by synergistic
sulfide supported on thin films of alumina, J. Catal. 103 (2) (1987) 366–384.
coupling of CeO2 and MoO3, J. Hazard. Mater. 381 (2019), 121037.
[48] T.A. Patterson, J.C. Carver, D.E. Leyden, D.M. Hercules, A surface study of cobalt-
[20] H. Zhao, et al., Catalytic oxidation of Hg0 with O2 induced by synergistic coupling
molybdena-alumina catalysts using x-ray photoelectron spectroscopy, J. Phys.
of CeO2and MoO3, J. Hazard. Mater. 381 (2020), 121037.
Chem. 80 (15) (1976) 1700–1708.
[21] T.H. Vuong, J. Radnik, J. Rabeah, U. Bentrup, M. Schneider, H. Atia,
[49] R.M. Martin-Aranda, M.F. Portela, L.M. Madeira, F. Freire, M. Oliveira, Effect of
U. Armbruster, W. Grünert, A. Brückner, Efficient VOx/Ce1–xTixO2 Catalysts for
alkali metal promoters on nickel molybdate catalysts and its relevance to the
Low-Temperature NH3-SCR: Reaction Mechanism and Active Sites Assessed by in
selective oxidation of butane, Appl. Catal. A 127 (1-2) (1995) 201–217.
Situ/Operando Spectroscopy, ACS Catal. 7 (3) (2017) 1693–1705.
[50] V.B. Sulimov, et al., Asymmetry and long-range character of lattice deformation by
[22] J. Li, N. Yan, Z. Qu, S. Qiao, S. Yang, Y. Guo, P. Liu, J. Jia, Catalytic oxidation of
neutral oxygen vacancy in \ensuremath{\alpha}-quartz, Phys. Rev. B 66 (2)
elemental mercury over the modified catalyst Mn/α-Al2O3 at lower temperatures,
(2002), 024108.
Environ. Sci. Technol. 44 (1) (2010) 426–431.
[51] S. Patil, S. Seal, Y.u. Guo, A. Schulte, J. Norwood, Role of trivalent La and Nd
[23] F. Cao, S. Su, J. Xiang, P. Wang, S. Hu, L. Sun, A. Zhang, The activity and
dopants in lattice distortion and oxygen vacancy generation in cerium oxide
mechanism study of Fe–Mn–Ce/γ-Al2 O3 catalyst for low temperature selective
nanoparticles, Appl. Phys. Lett. 88 (24) (2006) 243110, https://doi.org/10.1063/
catalytic reduction of NO with NH3, Fuel 139 (2015) 232–239.
1.2210795.
[24] B. Yue, R. Zhou, Y. Wang, X. Zheng, Influence of transition metals (Cr, Mn, Fe, Co
[52] X. Zhang, H. Zhang, H. Zhu, C. Li, N. Zhang, J. Bao, G. He, Co3O4 nanorods with a
and Ni) on the methane combustion over Pd/Ce–Zr/Al2O3 catalyst, Appl. Surf. Sci.
great amount of oxygen vacancies for highly efficient Hg0 oxidation from coal
252 (16) (2006) 5820–5828.
combustion flue gas, Energy Fuels 33 (7) (2019) 6552–6561.
[25] M. Jiang, B. Wang, Y. Yao, Z. Li, X. Ma, S. Qin, Q.i. Sun, A comparative study of
[53] H. Zhao, X. Mu, G. Yang, C. Zheng, C. Sun, X. Gao, T. Wu, Microwave-induced
CeO2-Al2O3 support prepared with different methods and its application on MoO3/
activation of additional active edge sites on the MoS2 surface for enhanced Hg0
CeO2-Al2O3 catalyst for sulfur-resistant methanation, Appl. Surf. Sci. 285 (2013)
capture, Appl. Surf. Sci. 420 (2017) 439–445.
267–277.
[54] Y. Wang, H. Li, S. Wang, X.u. Wang, Z. He, J. Hu, Investigation of sulphated CuCl2/
[26] W. Xiang, J. Liu, M. Chang, C. Zheng, The adsorption mechanism of elemental
TiO2 catalyst for simultaneous removal of Hg0 and NO in SCR process, Fuel
mercury on CuO (110) surface, Chem. Eng. J. 200-202 (2012) 91–96.
Process. Technol. 188 (2019) 179–189.
[27] H. Zhao, G. Yang, X. Gao, C. Pang, S. Kingman, E. Lester, T. Wu, Hg 0 -temperature-
programmed surface reaction and its application on the investigation of metal
oxides for Hg 0 capture, Fuel 181 (2016) 1089–1094.

You might also like