Next Article in Journal
Liquid Phase Furfural Oxidation under Uncontrolled pH in Batch and Flow Conditions: The Role of In Situ Formed Base
Next Article in Special Issue
Asymmetric Cyanation of Activated Olefins with Ethyl Cyanoformate Catalyzed by Ti(IV)-Catalyst: A Theoretical Study
Previous Article in Journal
Recent Advances in Rare Earth Complexes Containing N-Heterocyclic Carbenes: Synthesis, Reactivity, and Applications in Polymerization
Previous Article in Special Issue
A Four Coordinated Iron(II)-Digermyl Complex as an Effective Precursor for the Catalytic Dehydrogenation of Ammonia Borane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Promoting Effect of the Core-Shell Structure of MnO2@TiO2 Nanorods on SO2 Resistance in Hg0 Removal Process

State Key Laboratory of Fine Chemicals, School of Chemical Engineering at Panjin, Dalian University of Technology, Panjin 124221, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 72; https://doi.org/10.3390/catal10010072
Submission received: 5 December 2019 / Revised: 2 January 2020 / Accepted: 2 January 2020 / Published: 3 January 2020
(This article belongs to the Special Issue Advanced Strategies for Catalyst Design)

Abstract

:
Sorbent of αMnO2 nanorods coating TiO2 shell (denoted as αMnO2-NR@TiO2) was prepared to investigate the elemental mercury (Hg0) removal performance in the presence of SO2. Due the core-shell structure, αMnO2-NR@TiO2 has a better SO2 resistance when compared to αMnO2 nanorods (denoted as αMnO2-NR). Kinetic studies have shown that both the sorption rates of αMnO2-NR and αMnO2-NR@TiO2, which can be described by pseudo second-order models and SO2 treatment, did not change the kinetic models for both the two catalysts. In contrast, X-ray photoelectron spectroscopy (XPS) results showed that, after reaction in the presence of SO2, S concentration on αMnO2-NR@TiO2 surface is lower than on αMnO2-NR surface, which demonstrated that TiO2 shell could effectively inhibit the SO2 diffusion onto MnO2 surface. Thermogravimetry-differential thermosgravimetry (TG-DTG) results further pointed that SO2 mainly react with TiO2 forming Ti(SO4)O in αMnO2-NR@TiO2, which will protect Mn from being deactivated by SO2. These results were the reason for the better SO2 resistance of αMnO2-NR@TiO2.

Graphical Abstract

1. Introduction

The emission of mercury from coal-fired power plants has drawn wide public concern in modern society. Mercury emissions are a long-term threat to human health and the environment because of extreme toxicity, persistence, and bioaccumulation. Therefore, controlling mercury emitted from coal-fired power plants has practical significance. Mercury in coal combustion flue gas is mainly present in three forms: Elemental mercury (Hg0), oxidized mercury (Hg2+), and particulate-bound mercury (Hgp). Particulate-bound mercury (Hgp) can be removed by electrostatic precipitators (ESP) and fabric filters (FF), while oxidized mercury (Hg2+) can be captured by wet flue gas desulfurization system (WFGD). However, existing air pollution control devices can hardly remove Hg0 due to its high volatility and low solubility.
Hg0 capture with specific adsorbents is a usual way to control Hg0 emissions from coal-fired power plants [1]. Activated cabon (AC) has been widely used for the adsorption of Hg0 in coal-fired flue gas [2,3]. However, a huge amount of AC needs to be injected into flue gas because of its low Hg0 capture capacity, which leads to a high operating cost of this technology. Sulfur or halogen modification can enhance adsorption ability of AC [4,5]. However, the injected AC is usually captured together with fly ash by particulate control device, and the Hg0 adsorbed on AC will influence the fly ash utilization [6]. Therefore, alternative economic sorbents with high Hg0 removal efficiency are necessary.
Oxides, such as CuOx [7,8], FeOx [9,10], CeOx [11,12] and MnOx [13,14,15], with high redox properties, exhibit great potential for Hg0 adsorption. Among these oxides, MnOx is a commonly available and inexpensive material has received extensive attention due to the redox couples of Mn2+/Mn3+ and Mn3+/Mn4+ [16]. Electronic shift between the different valence states of Mn is active and leads to a high redox capacity. Stefano Cimino et al. [14] investigated the Hg0 removal performance of Mn/TiO2 and found that Hg0 capture efficiency was about 57% at 70 °C. After modification by some other transition metal oxides, Mn-based materials, such as Mn-FeOx [15], Mn-ZrOx [17], Mn-CeOx [18], and Mn-CuOx [19] can remove Hg0 better. Furthermore, it has been reported that the shape and crystallographic phases of Mn based sorbents have serious effects on Hg0 removal performance. Xu et al. [20] synthesized three different crystallographic phases of MnO2 and found that α-MnO2 had the highest capacity due to its larger surface area and oxidizability. Chalkidis et al. [21] pointed out that MnO2 nano-rods possessed good Hg0 removal capacity owing to the higher surface adsorbed oxygen species.
However, Mn-based sorbents usually have a poor SO2 resistance as SO2 can easily react with Mn, thereby forming MnSO4 and leading to a largely suppressed Hg0 removal activity. Even a little amount of SO2 will results in serious inhibited effects on Hg0 removal process. Our previous work has indicated that Ce-Zr modified Mn sorbent will be totally deactivated in 1h after the introduction of 50 ppm SO2 due to SO2 poisoning Mn forming MnSO4 [22]. TiO2 is a traditional way to enhance the SO2 resistance of MnOx [23] as TiO2 can inhibit the deposition of sulfates on sorbents surface [24]. But the Hg0 removal activity of MnOx/TiO2 is unsatisfactory because the active component of Mn is still exposed in SO2 atmosphere. Core-shell is a structure with active component core and supporting components shell. The shell can inhibit the interaction between SO2 and sorbent surface and efficiently protect active component core [25]. Therefore, synthesizing a core-shell structure with MnOx core and TiO2 shell may obtain a better SO2 resistance.
Inspired by this, αMnO2 nanorods and αMnO2 nanorods coating TiO2 shell were synthesized in the present work to investigate the Hg0 removal efficiency in the presence of SO2. Thermo-gravimetric (TG) and X-ray photoelectron spectroscopy (XPS) were performed to determine the role of SO2 in the Hg0 oxidation and adsorption processes and a probable mechanism of SO2 influence was deduced based on XPS and TG results. The kinetic model of the Hg0 adsorption process was examined as well.

2. Results and Discussion

2.1. Structure Characterization

Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) were performed to investigate the morphologic and structural properties of αMnO2-NR and αMnO2-NR@TiO2. Figure 1a,a’ show SEM and TEM images of αMnO2-NR. It can be seen that αMnO2-NR has a uniform nanorod structure with an average diameter of about 100 nm. As shown in Figure 1b, for αMnO2-NR@TiO2, the uniform nanorod structure is well-retained after being coated with TiO2 and the packing state of this sample is similar to αMnO2-NR. The surface of αMnO2-NR@TiO2 are rougher when compared to αMnO2-NR, and the average diameter increases to 150 nm due to the TiO2 coating. The average length of the αMnO2-NR@TiO2 is about 2–3 μm (shown in Figure 1c). As shown in Figure 1b’, an obvious dividing line can be detected between MnO2 core and TiO2 shell, and the shell with thickness of about 30 nm is well dispersed outside of the αMnO2-NR.
N2 sorption-desorption isotherms of the samples are shown in Figure 2. Both αMnO2-NR and αMnO2-NR@TiO2 exhibit a type IV adsorption isotherm, according to the definition of IUPAC, which means that αMnO2-NR and αMnO2-NR@TiO2 have a mesoporous structure. The surface areas, pore volumes, and average pore diameters of the sorbents are illustrated in Table 1. BET surface areas of the two sorbents are similar, suggesting that TiO2 coating does not change the structure of αMnO2-NR a lot. This result consists with SEM results.
X-ray diffractometer (XRD) patterns of the two catalysts are shown in Figure 3. All the peaks in XRD pattern of αMnO2-NR and αMnO2-NR@TiO2 were indexed to cryptomelane type α-MnO2 (JCPDS 44-0141, tetragonal, I4/m, a = b = 0.978 nm, c = 0.286 nm). The intensity of diffraction peaks for the two samples is almost the same. It means that TiO2 shell does not influence the dispersion of αMnO2-NR, which is great agreement with BET and SEM results.

2.2. Hg0 Adsorption

2.2.1. Hg0 Adsorption Performance

Breakthrough experiments were performed to investigate the Hg0 adsorption performance of the two sorbents. A blank test was also performed and the results is shown in Figure S1. It can be seen that the outlet Hg0 concentration is stable when no sorbent was loaded in the fixed-bed reactor. As shown in Figure 4, the Hg0 removal efficiency of αMnO2-NR is about 92% at the beginning of the test and it decreases to 41% after 130 min reaction. When it comes to αMnO2-NR@TiO2, the Hg0 removal efficiency at the beginning of the test is about 81% which is lower than that of αMnO2-NR. But it is about 43% at the end of the test suggesting a more stable removal activity. These results indicate that TiO2 shell does not inhibit the Hg0 diffusion from gas phase to the surface of αMnO2-NR.
Figure 5 shows the effects of SO2 on Hg0 adsorption performance. For αMnO2-NR, Hg0 removal efficiency sharply declines from 55% to 14% during the 35 min reaction, when SO2 is injected into flue gas. However, for αMnO2-NR@TiO2, the downward trend of Hg0 removal efficiency is much slower and decreases from 76% to 43% in a 30 min test, and still has a Hg0 removal efficiency of 25% after 80 min. These results confirm that TiO2 shell can inhibit the direct interaction between SO2 and MnO2 surface, which will efficiently protect MnO2 core from SO2 poisoning.
αMnO2-NR@TiO2 was used to investigate reusability for Hg0 removal. The results are shown in Figure 6. After 10 h Hg0 adsorption test, αMnO2-NR@TiO2 reaches a Hg0 adsorption equilibrium. And then, the sorbent was heated at 450 °C for 2 h to release the HgO on sorbent surface. It can be found that, after heated treatment, the Hg0 adsorption efficiency and capacity of αMnO2-NR@TiO2 recovers to its original level. After two recycling, it still shows a good Hg0 adsorption efficiency. Furthermore, SEM results of the fresh and used αMnO2-NR@TiO2 (shown in Figure S2) show that recycle have no effect on the microstructure. These results suggest an outstanding reusability of αMnO2-NR@TiO2. The Hg0 adsorption capacity of αMnO2-NR@TiO2 is 0.11 mg/g, it is good enough compared to other sorbents (shown in Table S1). The surface areas of the sorbents in the present work are relatively low thereby lowering the available surface active sites. αMnO2-NR@TiO2 with higher surface area will be studied in our following works, and may give a better Hg0 adsorption capacity.
Hg0 adsorption test of αMnO2-NR@TiO2 at different Hg0 concentration was also investigated and the results are shown in Figure S3. With a doubled Hg0 concentration, the breakthrough curve gets steep suggesting that αMnO2-NR@TiO2 will easily reach Hg0 adsorption equilibrium at a higher Hg0 concentration.

2.2.2. Structure-Activity Relationship

Fourier Transform Infrared Spectrometer (FTIR) was used to confirm the kind of surface active site for Hg0 adsorption. As can be seen in Figure 7, the peaks at 429, 503, and 700 cm−1 correspond to Mn-O vibration [26], which becomes much weaker after reaction. It suggests that Mn-O group participates in Hg0 adsorption process. According to previous work, the surface active oxygen species in Mn-O group should be the active sites for Hg0 adsorption.

2.3. Models of Adsorption Kinetics

In order to better illustrate the Hg0 adsorption mechanisms of αMnO2-NR and αMnO2-NR@TiO2, two popular models of pseudo-first order and pseudo-second order kinetic models, which have been widely used to investigate the adsorption process [27], were employed to fit the above experimental data. These two kinetic equations are displayed as follows [28]:
lg ( q e q t ) = lg q e k 1 2.303 t   pseudo first   order
t q t = 1 k 2 q e 2 + 1 q e   pseudo sec ond   order   kinetic
where qe and qt are the adsorption capacity of Hg0 on the sorbents at equilibrium, and at reaction time t (min), respectively. The parameters k1 (min−1) and k2 (g/(μg·min)) are the rate constants of the pseudo-first order, and pseudo-second order models, respectively.
The fitting results are shown in Figure 8, and the obtained values of correlation coefficient (R2) are summarized in Table 2. The values of R2 of the pseudo-second order model for αMnO2-NR and αMnO2-NR@TiO2 are 0.991, and 0.995, respectively, which are higher than those of pseudo-first order kinetic model (0.944 and 0.938 for αMnO2-NR and αMnO2-NR@TiO2). It indicates that the pseudo-second order model can better fit the experimental data and Hg0 removal process are dominantly controlled by chemisorption. After SO2 introduction, the values of R2 of the pseudo-second order model for αMnO2-NR and αMnO2-NR@TiO2 are 0.997 and 0.992, which are still much higher than those of the pseudo-first order model. These results show that Hg0 adsorption process in the presence of SO2 atmosphere are also dominantly controlled by chemisorption.

2.4. The Mechanism of SO2 Effects on Hg0 Adsorption

XPS analysis was employed to explore the relative proportion of elements on the sample surface. The XPS spectra of Mn 2p, O 1s and S 2p for the fresh and used samples are shown in Figure 9. The surface atomic concentrations and surface atomic ratios are summarized in Table 3.
Figure 9a shows the XPS spectra of Mn 2p. A doublet due to spin orbital coupling can be detected which corresponds to Mn 2p3/2 (around 641.24 eV) and Mn 2p1/2 (around 652.82 eV). Due to the high intensity of Mn 2p3/2, it was fitted to give detail information of valence state of Mn and it can be separated into three peaks at 640.2–641.2 eV, 641.2–642.1 eV, and 642.2–643.4 eV corresponding to Mn2+, Mn3+, and Mn4+, respectively [29,30]. As shown in Table 3, the ratio of Mn4+/Mn is about 37.8% for the fresh αMnO2-NR and it decreases to 33.4% after the SO2 resistance test. Compared to αMnO2-NR, Mn4+ content is almost constant for αMnO2-NR@TiO2 before, and after, SO2 resistance test. These results indicate that, for αMnO2-NR, Mn4+ is easily reduced to Mn2+ during SO2 resistance process via the reaction between SO2 and MnO2 [31]. For αMnO2-NR@TiO2, the interaction between SO2 and MnO2 is inhibited by the TiO2 shell structure, which can efficiently protect active component Mn4+ in the core.
Figure 9b shows O 1s XPS spectra. For the fresh catalysts, O 1s bands can be split into two peaks, corresponding to lattice oxygen (peak at 529.5 eV, denoted as Oα) and chemisorbed oxygen (peak at 530.8 eV, denoted as Oβ), respectively [32]. Whereas, a new peak appears around 532.3 eV after SO2 treatment, which corresponds to SO42− (denoted as Oγ) [33]. The intensity of the peak around 532.3 eV for αMnO2-NR@TiO2 is weaker than that for αMnO2-NR suggesting a lower amount of SO42− on the used αMnO2-NR@TiO2 surface. Furthermore, the peaks of Oα and Oβ in αMnO2-NR have an obvious slight shift to higher binding energy after SO2 treatment. It might be due to the formation of sulfate salts during the sulfating process [34].
To determine the above deduction, S 2p bands was further investigated and the results are shown in Figure 9c. For the fresh αMnO2-NR and αMnO2-NR@TiO2, two peaks around 162.2 eV and 163.2 eV attributed to S2− and S22− can be detected [35,36], which may come from MnSO4 (the precursor of MnO2). But for the used αMnO2-NR and αMnO2-NR@TiO2, two new peaks at about 168.8 eV and 170.0 eV are observed, which may be assigned to SO42−, and HSO4, respectively [37,38]. The peak intensity of the used αMnO2-NR is much higher than that of αMnO2-NR@TiO2. As shown in Table 3, for αMnO2-NR, the surface atomic concentrations of S increases from 3.17% to 4.97% after SO2 teatment while it increases from 2.27% to 2.66% for αMnO2-NR@TiO2. These results confirm that TiO2 shell can inhibit the S accumulation on catalyst surface.
To obtain more information about the SO2 poisoning mechanism, Thermo-gravimetric-differential thermos-gravimetry (TG-DTG) was performed to investigate the weight loss of αMnO2-NR and αMnO2-NR@TiO2 after SO2 treatment, and the results are presented in Figure 10. It can be seen that the used αMnO2-NR has an obvious weight loss step in the temperature range of 680−780 °C with a weight loss of about 2.4%, which can be attributed to manganese sulfate decomposition [39,40,41]. There is no weight loss step between 680−780 °C with respect to αMnO2-NR@TiO2, but there is a new weak step around 780–850 °C can be detected, and it may be due to the decomposition of Ti(SO4)O [42]. This result demonstrates that SO2 tends to react with titanium oxides instead of manganese oxides over αMnO2-NR@TiO2. Based on these results, TiO2 shell can lead to the preferential adsorption of SO2 on Ti surrounding forming Ti(SO4)O to protect Mn active component from being deactivated.

3. Materials and Methods

3.1. Catalysts Preparation

The αMnO2 nanorods were synthesized through a hydrothermal method [43]. KMnO4 (2.5 g, AR) and MnSO4·H2O (1.05 g, AR) were dissolved in 80 mL distilled water. The mixed solution was transferred into a Teflon-line stainless steel autoclave, sealed, and kept in an oven at 160 °C for 12 h. After cooling to room temperature, the precipitates were filtered off, washed several times using deionized water and dried at 110 °C overnight. Finally, the product was calcined at 400 °C in a muffle furnace for 4 h and the obtained sample is denoted as αMnO2-NR.
MnO2@TiO2 core-shell nanorods were synthesized through a versatile kinetics-controlled coating method [44]. αMnO2-NR (0.075 g) and aqueous ammonia (0.28 mL, 28 wt.%) were dispersed in 100 mL absolute ethanol under ultrasound for 30 min. Afterwards, titanium tetrabutoxide (TBOT) (0.75 mL) was added drop-wise into the mixture and then kept at 45 °C for 24 h. The mixed solution was filtered, washed and dried at 60 °C for 12 h. Finally, the solid was calcined under flow air at 500 °C for 2 h to obtain the sample (denoted as αMnO2-NR@TiO2).
ALL reagents are from Aladdin company, Shanghai, China.

3.2. Hg0 Adsorption Experiments

The Hg0 removal test has been described in detail in our previous work [45]. The experimental reactor contains a gas distribution system, a Hg0 vapor generating device, a fixed-bed quartz reactor (ID = 8 mm), an online mercury analyzer and a tail gas treating unit. The mercury permeation tube was placed in a U-shape glass tube, which was immersed in a water bath at a constant-temperature (38 °C) to ensure a constant Hg0 permeation rate. The total gas flow was 600 mL/min, and the sorbent volume was generally 0.2 mL, resulting in a GHSV of 1.8 × 105 h−1. The concentrations of Hg0 and SO2 were monitored by a VM-3000 online mercury analyzer (Mercury Instruments, München, German), and flue gas analyzer (KM950, Kane International Ltd., London, United Kingdom), respectively.
During each test, the Hg0 gas first bypassed the fixed-bed reactor, and then introduced into the reactor for 2 h to obtain a stable Hg0 concentration. Hg0 breakthrough ratio was quantified by the following formula,
B r e a k t h r o u g h   r a t i o ( % ) = C C 0 × 100 %
where C and C0 represent the inlet and outlet Hg0 concentrations (μg/Nm3) in the fixed-bed reactor.

3.3. Characterization

The morphology and microstructure of the samples were observed using SEM (Nova NanoSEM 450, FEI) and TEM (Tecnai G2 F30 S-Twin, FEI). The surface areas and pore parameters of the samples were determined by Nitrogen adsorption/desorption method at liquid nitrogen temperature at −196 °C on an automated gas sorption analyzer (Autosorb-iQ-C, Quantachrome Instruments, Boynton Beach, FL, USA). The pore size and pore volume were derived from the desorption branches using the Barrette-Joynere-Halenda (BJH) model. The crystal structures of the samples were characterized by an XRD (XRD-7000S, SHIMADZU Corporation, Kyoto, Japan) operating at 40 kV and 100 mA using a Cu Kα radiation. The scanning range (2θ) was from 10° to 90° with a scan speed of 5°/min. The element (Mn, O, and Hg) valence state was analyzed by XPS (ESCALAB250 Thermo Fisher Scientific, Wilmington, DE, USA) with a monochromatic Al Kα source. The C 1s binding energy value of 284.8 eV was used to calibrate the observed spectra. TG was performed on TGA/DSC1 analyser (METTLER TOLEDO, Schwerzenbach, Switzerland), under a nitrogen flow of 20 mL/min, using a heating rate of 10 °C/min from room temperature to 900 °C (NETZSCH Corporation, Selb, Germany). DTG analysis was obtained based on residual weight of the sample with respect to time. FTIR spectra were obtained on a Nicolet Magana-IR 750 spectrometer to measure the surface groups of the samples (Thermo Nicolet Corporation, Madison, WI, USA).

4. Conclusions

αMnO2-NR@TiO2 was prepared by versatile kinetics-controlled coating method to compare with αMnO2-NR in the Hg0 removal process. SEM, BET, and XRD results showed that TiO2 shell did not change the structure of αMnO2-NR. Therefore, the two sorbents had similar Hg0 removal performance in N2 atmosphere. When SO2 was introduced, αMnO2-NR@TiO2 had a much better performance than αMnO2-NR. XPS and TG-DTG results showed that αMnO2-NR@TiO2 had lower surface S concentration after treatment of SO2, and no manganese sulfate could be detected in αMnO2-NR@TiO2. It suggests that the TiO2 shell can effectively protect MnO2 from being deactivated by SO2. Adsorption kinetic results showed that Hg0 adsorption process over both the two sorbents obeys pseudo-second order model with, or without, SO2.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/1/72/s1, Figure S1: Outlet Hg0 concentration without sorbent, Figure S2: The image of αMnO2-NR@TiO2 after adsorption, Figure S3: Breakthrough curve of αMnO2-NR@TiO2 with different Hg0 feed concentration, Table S1: Comparison of the adsorption capacities of the sorbents.

Author Contributions

X.S. and H.Z. designed the experiments; X.H. and C.L. performed the experiments and analyzed the data; X.Z. wrote the paper; J.B., N.Z. and G.H. contributed reagents/materials/analysis tools. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [National Natural Science Foundation of China] grant number [51978124], [Liaoning Provincial Natural Science Foundation of China] grant number [20180510054].

Acknowledgments

We gratefully acknowledge the financial support of the National Natural Science Foundation of China (51978124), Liaoning Provincial Natural Science Foundation of China (20180510054), Shandong Provincial Natural Science Foundation PhD Programme (ZR2016EEB33), Foundation of China the Program for Changjiang Scholars (T2012049), Education Department of the Liaoning Province of China (LT2015007), the Fundamental Research Funds for the Central Universities (DUT18JC45).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhao, H.; Mu, X.; Yang, G.; George, M.; Cao, P.; Fanady, B.; Rong, S.; Gao, X.; Wu, T. Graphene-like MoS2 containing adsorbents for Hg0 capture at coal-fired power plants. Appl. Energy 2017, 207, 254–264. [Google Scholar] [CrossRef]
  2. Zhang, J.; Duan, Y.; Zhou, Q.; Zhu, C.; She, M.; Ding, W. Adsorptive removal of gas-phase mercury by oxygen non-thermal plasma modified activated carbon. Chem. Eng. J. 2016, 294, 281–289. [Google Scholar] [CrossRef]
  3. Zhang, B.; Zeng, X.; Xu, P.; Chen, J.; Xu, Y.; Luo, G.; Xu, M.; Yao, H. Using the Novel Method of Nonthermal Plasma To Add Cl Active Sites on Activated Carbon for Removal of Mercury from Flue Gas. Environ. Sci. Technol. 2016, 50, 11837–11843. [Google Scholar] [CrossRef] [PubMed]
  4. Zhou, Q.; Duan, Y.F.; Hong, Y.G.; Zhu, C.; She, M.; Zhang, J.; Wei, H.Q. Experimental and kinetic studies of gas-phase mercury adsorption by raw and bromine modified activated carbon. Fuel Process. Technol. 2015, 134, 325–332. [Google Scholar] [CrossRef]
  5. Liu, W.; Vidic, R.D.; Brown, T.D. Impact of Flue Gas Conditions on Mercury Uptake by Sulfur-Impregnated Activated Carbon. Environ. Sci. Technol. 2000, 34, 154–159. [Google Scholar] [CrossRef]
  6. Bisson, T.M.; Xu, Z. Potential Hazards of Brominated Carbon Sorbents for Mercury Emission Control. Environ. Sci. Technol. 2015, 49, 2496–2502. [Google Scholar] [CrossRef]
  7. Chen, C.; Jia, W.; Liu, S.; Cao, Y. The enhancement of CuO modified V2O5-WO3/TiO2 based SCR catalyst for Hg0 oxidation in simulated flue gas. Appl. Surf. Sci. 2018, 436, 1022–1029. [Google Scholar] [CrossRef]
  8. Li, H.; Zhang, W.; Wang, J.; Yang, Z.; Li, L.; Shih, K. Coexistence of enhanced Hg0 oxidation and induced Hg2+ reduction on CuO/TiO2 catalyst in the presence of NO and NH3. Chem. Eng. J. 2017, 330, 1248–1254. [Google Scholar] [CrossRef]
  9. Liu, T.; Xue, L.; Guo, X.; Liu, J.; Huang, Y.; Zheng, C. Mechanisms of Elemental Mercury Transformation on α-Fe2O3(001) Surface from Experimental and Theoretical Study: Influences of HCl, O2, and SO2. Environ. Sci. Technol. 2016, 50, 13585–13591. [Google Scholar] [CrossRef]
  10. Zarei, S.; Niad, M.; Raanaei, H. The removal of mercury ion pollution by using Fe3O4-nanocellulose: Synthesis, characterizations and DFT studies. J. Hazard. Mater. 2018, 344, 258–273. [Google Scholar] [CrossRef]
  11. He, C.; Shen, B.; Chi, G.; Li, F. Elemental mercury removal by CeO2/TiO2-PILCs under simulated coal-fired flue gas. Chem. Eng. J. 2016, 300, 1–8. [Google Scholar] [CrossRef]
  12. Zhu, Y.; Han, X.; Huang, Z.; Hou, Y.; Guo, Y.; Wu, M. Superior activity of CeO2 modified V2O5/AC catalyst for mercury removal at low temperature. Chem. Eng. J. 2018, 337, 741–749. [Google Scholar] [CrossRef]
  13. Xu, H.; Jia, J.; Guo, Y.; Qu, Z.; Liao, Y.; Xie, J.; Shangguan, W.; Yan, N. Design of 3D MnO2/Carbon sphere composite for the catalytic oxidation and adsorption of elemental mercury. J. Hazard. Mater. 2018, 342, 69–76. [Google Scholar] [CrossRef] [PubMed]
  14. Cimino, S.; Scala, F. Removal of Elemental Mercury by MnOx Catalysts Supported on TiO2 or Al2O3. Ind. Eng. Chem. Res. 2016, 55, 5133–5138. [Google Scholar] [CrossRef]
  15. Zhang, S.; Zhao, Y.; Yang, J.; Zhang, J.; Zheng, C. Fe-modified MnOx/TiO2 as the SCR catalyst for simultaneous removal of NO and mercury from coal combustion flue gas. Chem. Eng. J. 2018, 348, 618–629. [Google Scholar] [CrossRef]
  16. Yao, T.; Duan, Y.; Zhu, C.; Zhou, Q.; Xu, J.; Liu, M.; Wei, H. Investigation of mercury adsorption and cyclic mercury retention over MnOx/γ-Al2O3 sorbent. Chemosphere 2018, 202, 358–365. [Google Scholar] [CrossRef]
  17. Xie, J.K.; Qu, Z.; Yan, N.Q.; Yang, S.J.; Chen, W.M.; Hu, L.G.; Huang, W.J.; Liu, P. Novel regenerable sorbent based on Zr–Mn binary metal oxides for flue gas mercury retention and recovery. J. Hazard. Mater. 2013, 261, 206–213. [Google Scholar] [CrossRef]
  18. Li, H.; Wang, Y.; Wang, S.; Wang, X.; Hu, J. Removal of elemental mercury in flue gas at lower temperatures over Mn-Ce based materials prepared by co-precipitation. Fuel 2017, 208, 576–586. [Google Scholar] [CrossRef]
  19. Yi, Y.; Li, C.; Zhao, L.; Du, X.; Gao, L.; Chen, J.; Zhai, Y.; Zeng, G. The synthetic evaluation of CuO-MnOx-modified pinecone biochar for simultaneous removal formaldehyde and elemental mercury from simulated flue gas. Environ. Sci. Pollut. Res. 2017, 25, 4761–4775. [Google Scholar] [CrossRef]
  20. Xu, H.; Qu, Z.; Zhao, S.; Mei, J.; Quan, F.; Yan, N. Different crystal-forms of one-dimensional MnO2 nanomaterials for the catalytic oxidation and adsorption of elemental mercury. J. Hazard. Mater. 2015, 299, 86–93. [Google Scholar] [CrossRef]
  21. Chalkidis, A.; Jampaiah, D.; Hartley, P.G.; Sabri, Y.M.; Bhargava, S.K. Regenerable α-MnO2 nanotubes for elemental mercury removal from natural gas. Fuel Process. Technol. 2019, 193, 317–327. [Google Scholar] [CrossRef]
  22. Zhang, X.; Li, Z.; Wang, J.; Tan, B.; Cui, Y.; He, G. Reaction mechanism for the influence of SO2 on Hg0 adsorption and oxidation with Ce0.1-Zr-MnO2. Fuel 2017, 203, 308–315. [Google Scholar] [CrossRef]
  23. Yang, Z.; Li, H.; Liu, X.; Li, P.; Yang, J.; Lee, P.H.; Shih, K. Promotional effect of CuO loading on the catalytic activity and SO2 resistance of MnOx/TiO2 catalyst for simultaneous NO reduction and Hg0 oxidation. Fuel 2018, 227, 79–88. [Google Scholar] [CrossRef]
  24. Li, H.; Wu, C.Y.; Li, Y.; Zhang, J. Superior activity of MnOx-CeO2/TiO2 catalyst for catalytic oxidation of elemental mercury at low flue gas temperatures. Appl. Catal. B 2012, 111, 381–388. [Google Scholar] [CrossRef]
  25. Mitsudome, T.; Yamamoto, M.; Maeno, Z.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. One-step Synthesis of Core-Gold/Shell-Ceria Nanomaterial and Its Catalysis for Highly Selective Semihydrogenation of Alkynes. J. Am. Chem. Soc. 2015, 137, 13452–13455. [Google Scholar] [CrossRef] [PubMed]
  26. Wu, J.; Zhou, T.; Wang, Q.; Umar, A. Morphology and chemical composition dependent synthesis and electrochemical properties of MnO2-based nanostructures for efficient hydrazine detection. Sens. Actuators B 2016, 224, 878–884. [Google Scholar] [CrossRef]
  27. Tabish, T.A.; Memon, F.A.; Gomez, D.E.; Horsell, D.W.; Zhang, S. A facile synthesis of porous graphene for efficient water and wastewater treatment. Sci. Rep. 2018, 8, 1817. [Google Scholar] [CrossRef]
  28. McKay, G.; Ho, Y.S. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar]
  29. Zhang, D.; Hou, L.A.; Chen, G.; Zhang, A.; Wang, F.; Wang, R.; Li, C. Cr Doping MnOx Adsorbent Significantly Improving Hg0 Removal and SO2 Resistance from Coal-Fired Flue Gas and the Mechanism Investigation. Ind. Eng. Chem. Res. 2018, 57, 17245–17258. [Google Scholar] [CrossRef]
  30. Hu, H.; Cai, S.; Li, H.; Huang, L.; Shi, L.; Zhang, D. Mechanistic Aspects of deNOx Processing over TiO2 Supported Co–Mn Oxide Catalysts: Structure–Activity Relationships and In Situ DRIFTs Analysis. ACS Catal. 2015, 5, 6069–6077. [Google Scholar] [CrossRef]
  31. Xu, H.; Qu, Z.; Zong, C.; Quan, F.; Mei, J.; Yan, N. Catalytic oxidation and adsorption of Hg0 over low-temperature NH3-SCR LaMnO3 perovskite oxide from flue gas. Appl. Catal. B 2016, 186, 30–40. [Google Scholar] [CrossRef]
  32. Peng, Y.; Qu, R.; Zhang, X.; Li, J. The relationship between structure and activity of MoO3-CeO2 catalysts for NO removal: Influences of acidity and reducibility. Chem. Commun. 2013, 49, 6215–6217. [Google Scholar] [CrossRef] [PubMed]
  33. Ma, Y.; Mu, B.; Zhang, X.; Yuan, D.; Ma, C.; Xu, H.; Qu, Z.; Fang, S. Graphene enhanced Mn-Ce binary metal oxides for catalytic oxidation and adsorption of elemental mercury from coal-fired flue gas. Chem. Eng. J. 2019, 358, 1499–1506. [Google Scholar] [CrossRef]
  34. Reddy, B.M.; Sreekanth, P.M.; Yamada, Y.; Xu, Q.; Kobayashi, T. Surface characterization of sulfate, molybdate, and tungstate promoted TiO2-ZrO2 solid acid catalysts by XPS and other techniques. Appl. Catal. A 2002, 228, 269–278. [Google Scholar] [CrossRef]
  35. Li, H.; Zhu, L.; Wang, J.; Li, L.; Shih, K. Development of Nano-Sulfide Sorbent for Efficient Removal of Elemental Mercury from Coal Combustion Fuel Gas. Environ. Sci. Technol. 2016, 50, 9551–9557. [Google Scholar] [CrossRef] [PubMed]
  36. Liao, Y.; Chen, D.; Zou, S.; Xiong, S.; Xiao, X.; Dang, H.; Chen, T.; Yang, S. Recyclable Naturally Derived Magnetic Pyrrhotite for Elemental Mercury Recovery from Flue Gas. Environ. Sci. Technol. 2016, 50, 10562–10569. [Google Scholar] [CrossRef]
  37. Yang, S.; Guo, Y.; Yan, N.; Wu, D.; He, H.; Xie, J.; Qu, Z.; Jia, J. Remarkable effect of the incorporation of titanium on the catalytic activity and SO2 poisoning resistance of magnetic Mn–Fe spinel for elemental mercury capture. Appl. Catal. B Environ. 2011, 101, 698–708. [Google Scholar] [CrossRef]
  38. Xu, W.; He, H.; Yu, Y. Deactivation of a Ce/TiO2 Catalyst by SO2 in the Selective Catalytic Reduction of NO by NH3. J. Phys. Chem. C 2009, 113, 4426–4432. [Google Scholar] [CrossRef]
  39. Yu, J.; Guo, F.; Wang, Y.; Zhu, J.; Liu, Y.; Su, F.; Gao, S.; Xu, G. Sulfur poisoning resistant mesoporous Mn-base catalyst for low-temperature SCR of NO with NH3. Appl. Catal. B Environ. 2010, 95, 160–168. [Google Scholar] [CrossRef]
  40. Zhang, A.; Zhang, Z.; Lu, H.; Liu, Z.; Xiang, J.; Zhou, C.; Xing, W.; Sun, L. Effect of Promotion with Ru Addition on the Activity and SO2 Resistance of MnOx–TiO2 Adsorbent for Hg0 Removal. Ind. Eng. Chem. Res. 2015, 54, 2930–2939. [Google Scholar] [CrossRef]
  41. Wu, Z.; Jin, R.; Wang, H.; Liu, Y. Effect of ceria doping on SO2 resistance of Mn/TiO2 for selective catalytic reduction of NO with NH3 at low temperature. Catal. Commun. 2009, 10, 935–939. [Google Scholar] [CrossRef]
  42. Gardy, J.; Hassanpour, A.; Lai, X.; Ahmed, M.H. Synthesis of Ti(SO4)O solid acid nano-catalyst and its application for biodiesel production from used cooking oil. Appl. Catal. A 2016, 527, 81–95. [Google Scholar] [CrossRef]
  43. Liang, S.; Teng, F.; Bulgan, G.; Zong, R.; Zhu, Y. Effect of phase structure of MnO2 nanorod catalyst on the activity for CO oxidation. J. Phys. Chem. C 2008, 112, 5307–5315. [Google Scholar] [CrossRef]
  44. Li, W.; Yang, J.; Wu, Z.; Wang, J.; Li, B.; Feng, S.; Deng, Y.; Zhang, F.; Zhao, D. A versatile kinetics-controlled coating method to construct uniform porous TiO2 shells for multifunctional core–shell structures. J. Am. Chem. Soc. 2012, 134, 11864–11867. [Google Scholar] [CrossRef]
  45. Zhang, X.; Cui, Y.; Wang, J.; Tan, B.; Li, C.; Zhang, H.; He, G. Simultaneous removal of Hg0 and NO from flue gas by Co0.3-Ce0.35-Zr0.35O2 impregnated with MnOx. Chem. Eng. J. 2017, 326, 1210–1222. [Google Scholar] [CrossRef]
Figure 1. Scanning electron microscopy (SEM), and transmission electron microscopy (TEM) images of (a,a’) αMnO2-NR; (b,b’), and (c) αMnO2-NR@TiO2.
Figure 1. Scanning electron microscopy (SEM), and transmission electron microscopy (TEM) images of (a,a’) αMnO2-NR; (b,b’), and (c) αMnO2-NR@TiO2.
Catalysts 10 00072 g001
Figure 2. N2 sorption-desorption isotherms for the sorbents.
Figure 2. N2 sorption-desorption isotherms for the sorbents.
Catalysts 10 00072 g002
Figure 3. X-ray diffractometer (XRD) patterns of αMnO2-NR and αMnO2-NR@TiO2.
Figure 3. X-ray diffractometer (XRD) patterns of αMnO2-NR and αMnO2-NR@TiO2.
Catalysts 10 00072 g003
Figure 4. Hg0 breakthrough curves of αMnO2-NR and αMnO2-NR@TiO2 under pure N2 atmosphere. Reaction condition: 150 °C, GHSV = 180,000 h−1.
Figure 4. Hg0 breakthrough curves of αMnO2-NR and αMnO2-NR@TiO2 under pure N2 atmosphere. Reaction condition: 150 °C, GHSV = 180,000 h−1.
Catalysts 10 00072 g004
Figure 5. Hg0 breakthrough curves of αMnO2-NR and αMnO2-NR@TiO2 under 100 ppm SO2, N2 balanced. Reaction condition: 150 °C, GHSV = 180,000 h−1.
Figure 5. Hg0 breakthrough curves of αMnO2-NR and αMnO2-NR@TiO2 under 100 ppm SO2, N2 balanced. Reaction condition: 150 °C, GHSV = 180,000 h−1.
Catalysts 10 00072 g005
Figure 6. Hg0 breakthrough curves of αMnO2-NR@TiO2 under pure N2 atmosphere. Reaction condition: 150 °C, GHSV = 180,000 h−1.
Figure 6. Hg0 breakthrough curves of αMnO2-NR@TiO2 under pure N2 atmosphere. Reaction condition: 150 °C, GHSV = 180,000 h−1.
Catalysts 10 00072 g006
Figure 7. FTIR spectrum of αMnO2-NR@TiO2 before and after test.
Figure 7. FTIR spectrum of αMnO2-NR@TiO2 before and after test.
Catalysts 10 00072 g007
Figure 8. Kinetic analysis of Hg0 adsorption on αMnO2-NR and αMnO2-NR@TiO2 (a) pseudo-first order kinetic model without SO2, (b) pseudo-second order kinetic model without SO2, (c) pseudo-first order kinetic model with SO2, (d) pseudo-second order kinetic model with SO2.
Figure 8. Kinetic analysis of Hg0 adsorption on αMnO2-NR and αMnO2-NR@TiO2 (a) pseudo-first order kinetic model without SO2, (b) pseudo-second order kinetic model without SO2, (c) pseudo-first order kinetic model with SO2, (d) pseudo-second order kinetic model with SO2.
Catalysts 10 00072 g008
Figure 9. XPS spectra of (a) Mn 2p, (b) O 1s and (c) S 2p.
Figure 9. XPS spectra of (a) Mn 2p, (b) O 1s and (c) S 2p.
Catalysts 10 00072 g009aCatalysts 10 00072 g009b
Figure 10. Thermo-gravimetric (TG) and differential thermos-gravimetry (DTG) of spectras of αMnO2-NR and αMnO2-NR@TiO2 after SO2 treatment.
Figure 10. Thermo-gravimetric (TG) and differential thermos-gravimetry (DTG) of spectras of αMnO2-NR and αMnO2-NR@TiO2 after SO2 treatment.
Catalysts 10 00072 g010
Table 1. Pore structure analysis of the sorbents.
Table 1. Pore structure analysis of the sorbents.
SamplesBET Surface Area (m2/g)Pore Volume (cm3/g)Average Pore Diameter (nm)
αMnO2-NR29.1030.1925.428
αMnO2-NR@TiO232.9850.2074.186
Table 2. Kinetic parameters (R2) of pseudo-first order and pseudo-second order models.
Table 2. Kinetic parameters (R2) of pseudo-first order and pseudo-second order models.
Kinetic ModelsαMnO2-NR without SO2αMnO2-NR@TiO2 without SO2αMnO2-NR with SO2αMnO2-NR@TiO2 with SO2
Pseudo-first (R2)0.9440.9380.9540.941
Pseudo-second (R2)0.9910.9950.9970.992
Table 3. The surface atomic concentrations and the relative concentration ratios of samples based on XPS.
Table 3. The surface atomic concentrations and the relative concentration ratios of samples based on XPS.
SamplesSMn4+/MnOβ/O
αMnO2-NR (fresh)3.1737.826.0
αMnO2-NR@TiO2 (fresh)2.2733.424.7
αMnO2-NR (used)4.9734.022.8
αMnO2-NR@TiO2 (used)2.6633.020.0

Share and Cite

MDPI and ACS Style

Zhang, X.; Han, X.; Li, C.; Song, X.; Zhu, H.; Bao, J.; Zhang, N.; He, G. Promoting Effect of the Core-Shell Structure of MnO2@TiO2 Nanorods on SO2 Resistance in Hg0 Removal Process. Catalysts 2020, 10, 72. https://doi.org/10.3390/catal10010072

AMA Style

Zhang X, Han X, Li C, Song X, Zhu H, Bao J, Zhang N, He G. Promoting Effect of the Core-Shell Structure of MnO2@TiO2 Nanorods on SO2 Resistance in Hg0 Removal Process. Catalysts. 2020; 10(1):72. https://doi.org/10.3390/catal10010072

Chicago/Turabian Style

Zhang, Xiaopeng, Xiangkai Han, Chengfeng Li, Xinxin Song, Hongda Zhu, Junjiang Bao, Ning Zhang, and Gaohong He. 2020. "Promoting Effect of the Core-Shell Structure of MnO2@TiO2 Nanorods on SO2 Resistance in Hg0 Removal Process" Catalysts 10, no. 1: 72. https://doi.org/10.3390/catal10010072

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop