Next Article in Journal
Research Progress and Reaction Mechanism of CO2 Methanation over Ni-Based Catalysts at Low Temperature: A Review
Next Article in Special Issue
Ordered Mesoporous MnAlOx Oxides Dominated by Calcination Temperature for the Selective Catalytic Reduction of NOx with NH3 at Low Temperature
Previous Article in Journal
The Inorganic Perovskite-Catalyzed Transfer Hydrogenation of Cinnamaldehyde Using Glycerol as a Hydrogen Donor
Previous Article in Special Issue
Study on the Mechanism of SO2 Poisoning of MnOx/PG for Lower Temperature SCR by Simple Washing Regeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hg0 Removal by V2O5 Modified Palygorskite in Simulated Flue Gas at Low Temperature

1
College of Chemistry and Chemical Engineering, Anqing Normal University, Anqing 246011, China
2
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(2), 243; https://doi.org/10.3390/catal12020243
Submission received: 20 January 2022 / Revised: 17 February 2022 / Accepted: 17 February 2022 / Published: 21 February 2022
(This article belongs to the Special Issue Frontiers in Catalytic Emission Control)

Abstract

:
The V2O5-modified palygorskite (V2O5/PG catalysts) were prepared and used for Hg0 removal in simulated flue gas at low temperature. It was found that the V2O5/PG catalyst had excellent performance for Hg0 removal at 150 °C. O2 exhibited a positive effect on Hg0 removal over V2O5/PG, while SO2 and H2O showed an inhibiting effect. However, Hg0 removal efficiency showed a promotion trend in the presence of H2O, SO2, and O2. The Brunauer–Emmett–Teller (BET) method, scanning electron microscopy (SEM), and X-ray photoelectron spectroscopy (XPS) were applied to characterize the physicochemical properties of the V2O5/PG catalyst. Mercury temperature-programmed desorption (Hg-TPD) experiments were also conducted to identify the mercury species adsorbed on the V2O5/PG catalyst, and the pathway of Hg0 removal over V2O5/PG was also discussed. The used V2O5/PG catalyst after Hg0 removal was regenerated, and its capability for Hg0 removal can be completely recovered. The V2O5/PG-Re-300 °C catalyst showed excellent performance and good stability for Hg0 removal after regeneration.

1. Introduction

Mercury, as a trace toxic chemical, is harmful and threatening to people’s health and the environment [1,2,3]. In China, about 38% of annual mercury emissions are from coal-fired power plants, which are regarded as the main anthropogenic source of mercury emissions [4,5,6]. Therefore, more and more attention has been paid to mercury emission control from coal-fired flue gas. To control mercury release, a series of environmental regulations and laws have been promulgated. The Minamata Convention on Mercury has been in effect since 16 August 2017, which means that reducing atmospheric mercury emissions has been a consensus-driven and compulsive goal of many countries around the world [7]. Moreover, the new Emission Standard of Air Pollutants for Thermal Power Plants in China has been executed, and explicitly stipulates that the emission limitation of mercury and its compounds in coal-fired flue gas is 30 μg/m3 [8]. Therefore, the control situation of mercury in coal-fired flue gas is very serious. The urgent task is to develop and research effective and economical technology for mercury emissions in coal-fired power plants.
Generally, coal-fired flue gas mainly contains several components, including O2 (3–7%) (v/v), moisture (8–10%) (v/v), SO2 (100–3000 ppm), etc. [9,10]. Mercury in coal-fired flue gas mainly exists in three forms: oxidized mercury (Hg2+), particulate mercury (Hgp), and elemental mercury (Hg0) [11,12]. Among them, wet flue gas desulfurization (WFGD) can effectively remove Hg2+ [13,14], and electrostatic precipitators (ESPs) or fabric filters (FFs) can easily capture Hgp [15]. However, due to its high volatility, stability, and insolubility, Hg0 is difficult to remove by existing air pollution control devices (APCDs) [16,17]. Therefore, researchers focus their efforts on controlling the emission of Hg0 from flue gas.
Various methods have been studied and used to control Hg0 emission, such as activated carbon injection (ACI) [18,19], catalytic oxidation, photochemical oxidation, and adsorption [20,21,22]. Among them, ACI technology is considered as the most effective Hg0 control technology [23]. However, the large-scale industrial application of ACI technology is limited due to its high operation cost [24,25,26].
In recent years, the natural mineral has attracted the attention of researchers since the raw materials are abundant in reserves and are cost-effective [27,28]. Palygorskite (noted as PG), as a kind of natural magnesium aluminum silicate clay mineral, has a unique crystal structure, well-developed pores, high specific surface area, high adsorption ability, good cohesiveness and mechanical properties, low price, and can be used as a good catalyst carrier. However, few studies have been reported on Hg0 removal by PG from flue gas at low temperature.
It is well known that V2O5-based selective catalytic reduction (SCR) catalysts are widely used for NOx removal in coal-fired flue gas [29,30]. Meanwhile, it was also found that V2O5 exhibited high catalytic oxidation activity for Hg0 and can effectively oxidize Hg0 to Hg2+ [31,32,33,34]. However, V2O5-based SCR catalysts themselves have limited ability to adsorb the formed Hg2+ at high operating temperatures (300–450 °C), and the formed Hg2+ must be removed by the downstream WFGD unit [13]. Therefore, it is necessary to develop cost-effective and environmentally friendly technologies for Hg0 control in coal-fired flue gas at low temperature.
In this work, palygorskite-supported V2O5 catalysts (V2O5/PG) were prepared and used to remove Hg0 in simulated flue gas at low temperature (120–210 °C). The V2O5/PG catalysts were characterized by BET, SEM, XPS, and Hg-TPD. Moreover, the effects of V2O5 loading, temperature, flue gas components (SO2, H2O, and O2), space velocity (GHSV), and the reaction pathways were also studied in the presence of SO2. The regeneration of the used V2O5/PG catalyst after Hg0 removal and reuse for Hg0 removal were also studied.

2. Results and Discussion

2.1. Effect of Reaction Temperature

Figure 1 shows the effect of reaction temperature (120 °C, 150 °C, 180 °C, and 210 °C) on Hg0 removal over 5V2O5/PG catalyst. It can be seen that the initial EHg of 5V2O5/PG reached 98% at 120 °C, and that it decreased gradually to 73.8% at 420 min. As the temperature increased from 150 °C to 180 °C and 210 °C, EHg decreased from 83.1%, 78.4%, and 52.5% at 420 min, respectively. Generally, adsorption becomes weaker at high temperature while oxidation will be enhanced. The difference of EHg at different temperatures may be due to the different effect extent of adsorption and oxidation at different temperatures [35]. The 5V2O5/PG catalyst had the highest Hg0 removal capability at 150 °C, which was much lower than the operating temperature of the V2O5-based SCR catalyst. Meanwhile, the V2O5/PG catalyst had much higher adsorption ability than that of V2O5-based SCR catalyst. Furthermore, the optimum reaction temperature (150 °C) was very close to the actual exhaust gas temperature of flue gas in a power plant, indicating that the V2O5/PG catalyst can be used to remove Hg0 from flue gas without an extra heat supply.

2.2. Effect of V2O5 Loading

Figure 2 shows the effect of V2O5 loading on Hg0 removal over V2O5/PG at 150 °C. Since it was found that a V2O5 loading range of 1% to 5% was favorable in our previous research, V2O5 loading (1% to 5%) was selected in this paper [36]. It can be seen that EHg of the PG carrier was very low, and that it decreased rapidly from 50% to 8.3% in 300 min. The EHg of the V2O5/PG was obviously higher than that of the PG, which was mainly due to the contribution of V2O5. V2O5, as the active site for Hg0 oxidation, was of great importance in Hg0 removal by V2O5/PG. It was found that lattice oxygen of V2O5 played a critical role in Hg0 oxidation in our previous research, which was consumed in Hg0 oxidation and can be subsequently replenished by gas-phase O2 [35]. With the increase of V2O5 loading (1% to 5%), the EHg increased from 38.6% to 86.0% at 420 min. Since the 5V2O5/PG catalyst showed the best Hg0 removal performance, 5V2O5/PG was used in the following experiments.
The specific surface areas and other pore parameters of the PG and V2O5/PG catalysts are summarized in Table 1. It can be seen that the BET surface area and total pore volume of the V2O5/PG catalyst are lower than those of the pure PG carrier. With the increase of the V2O5 loading, the BET area of the catalyst decreased obviously (from 151.4 m2/g to 109.3 m2/g). This may be due to the pores of the PG carrier being blocked by the active component of V2O5 and/or the sloughing of the PG skeleton during the calcination process, which led to the decrease of the BET area of the V2O5/PG catalyst. It is well known that a larger BET area can provide more reaction sites for the adsorption of Hg0. However, the specific surface area of the V2O5/PG catalyst showed little effect on Hg0 removal, i.e., the BET area was not the decisive factor affecting Hg0 removal, while the oxidation activity of V2O5 played a critical role in Hg0 removal. This was similar to the results reported in the literature [35].
Figure 3A,B shows the SEM morphology of 5V2O5/PG. It can be seen that the V2O5/PG catalyst had a porous structure, which was conducive to the distribution of V2O5 on the surface of the PG carrier and the adsorption of Hg0 on the surface of the V2O5/PG catalyst. The Lewis acid–base properties of the PG and V2O5/PG catalysts were analyzed and the detailed results are shown in the Supporting Information. The results showed that there were several different acid sites on the V2O5/PG catalyst surface, which were beneficial for Hg0 adsorption on V2O5/PG.

2.3. Effect of Flue Gas Components on Hg0 Removal

Figure 4A shows the effects of flue gas components on Hg0 removal over 5V2O5/PG. It can be seen that the 5V2O5/PG catalyst exhibited excellent Hg0 removal capability (EHg = 90.9%) within 400 min in a N2 atmosphere. However, after 0.15% SO2 was introduced, an obviously inhibitive effect was observed, and the EHg decreased to 43.5%. A similar negative effect (EHg = 53.48%) was also observed by adding 5% H2O into N2. These may be due to the competitive adsorption of SO2 (or H2O) with Hg0 on the surface of 5V2O5/PG in a N2 atmosphere, as well as the reaction of SO2, H2O, and V2O5 [18,35]. Adding 8% O2 into the N2 atmosphere promoted Hg0 removal over V2O5/PG, and the EHg increased to 96.8%, indicating that O2 played a positive role in Hg0 removal. Since no Hg0 oxidation by O2 was measured in the gas phase, the effect of O2 should be on V2O5, i.e., V2O5 was reduced to V4+ in Hg0 oxidation and lost its oxidation activity, while O2 can replenish O to the used V2O5 sites and resume its oxidation activity. This was consistent with SO2 removal and Hg0 oxidation over V2O5/AC catalysts in our previous research [37], and is similar to Hg0 oxidation on metal oxide catalysts [38,39].
Figure 4B shows Hg0 removal over 5V2O5/PG in a complex atmosphere containing more components. As mentioned above (Figure 4A), EHg was 43.5% in N2 + SO2, and it decreased to 34.2% after adding 5% H2O into N2 + SO2. However, after 8% O2 was added into N2 + SO2 or N2 + H2O, the EHg increased to 72.4% and 77.4%, respectively. This suggested that O2 played a critical role and could offset the negative effect of H2O (or SO2) on Hg0 removal to a certain extent [18,40,41,42]. The EHg increased to 86.9% after 5% H2O was added into N2 + O2 + SO2, which may be due to the formation of SO 4 2 - , which could react with Hg0 [43,44,45]. HgSO4 was formed, as shown in the following TPD experiment results.

2.4. Effects of GHSV on Hg0 Removal

The GHSV, as an important industrial parameter, can directly affect the contact time between the flue gas components and catalyst bed, and it has an important influence on the Hg0 removal process. Figure 5 shows the results of Hg0 removal by 5V2O5/PG at different GHSV (6000 h−1, 10,000 h−1, and 15,000 h−1) at 150 °C. It can be seen that GHSV had an obvious effect on Hg0 removal over 5V2O5/PG. As the GHSV increased from 6000 h−1 to 15,000 h−1, the EHg decreased from 86.9% to 57.9% in 400 min. The lower GHSV (6000 h−1) is more beneficial for Hg0 removal over 5V2O5/PG. This may be due to the lower GHSV increasing the contact time of Hg0 and 5V2O5/PG [35]. The 5V2O5/PG catalyst exhibited good Hg0 removal activity at 6000 h−1, which can match the actual GHSV of flue gas in a power plant and was suitable for mercury removal in power plants.

2.5. The Pathway of Hg0 Removal over V2O5/PG

The speciation of Hg adsorbed on the surface of V2O5/PG was characterized by XPS, and the results are shown in Figure 6. It can be seen that there was only one peak, at 102.90 eV, for the fresh 5V2O5/PG, which could be attributed to Si 2p in the PG carrier. As for the spent 5V2O5/PG after Hg0 removal, two peaks were observed at 103.1 eV and 107.15 eV, respectively. The peak at 103.1 eV can be assigned to Si 2p of the PG carrier, while the peak at 107.15 eV was attributed to Hg2+ [24,28].
To further investigate the mercury species adsorbed over V2O5/PG, the experiments of Hg temperature-programmed desorption (Hg-TPD) were performed for the fresh and spent 5V2O5/PG. As shown in Figure 7, it can be seen that there was no Hg release peak during the whole Hg-TPD process for the fresh 5V2O5/PG sample. However, as for the spent 5V2O5/PG sample, two Hg release peaks appeared at 240 °C and 495 °C, respectively. This indicated that there were two forms of mercury on the surface of the spent 5V2O5/PG, which can be ascribed to HgO and HgSO4 [46,47]; the relevant reactions may be described as follows, and the enthalpy change of adsorption and reactions were calculated in the Vienna ab initio simulation package (VASP 5.4.4) [48]:
Hg0 (gas) → Hg0 (ad)      (△H = −18.337 kJ/mol)
O2 (gas) → O2 (ad)       (△H = −12.589 kJ/mol)
SO2 (gas) → SO2 (ad)      (△H = −33.545 kJ/mol)
2Hg0 (ad) + O2 (ad) → 2HgO (ad)    (△H = 147.448 kJ/mol)
SO2 (ad) + O2 (ad) + Hg0 (ad) → HgSO4 (ad) (△H = −161.557 kJ/mol)
Additionally, it can be seen that the Hg adsorbed on 5V2O5/PG began to release at 150 °C, which could well explain the result of Figure 2 that the EHg decreased as the temperature was higher than 150 °C.

2.6. Regeneration of V2O5/PG Catalyst after Hg0 Removal

The above results showed that the 5V2O5/PG catalyst had an excellent Hg0 removal capability at low temperature. To investigate the reusability of the 5V2O5/PG catalyst, the 5V2O5/PG after Hg0 removal was regenerated and reused for Hg0 removal again. Since the V2O5/PG catalyst was prepared by calcining at 300 °C in air, a regeneration temperature was chosen from 300 °C to 500 °C to resume its catalytic activity. The results of the PG, 5V2O5/PG, 5V2O5/PG-Re-300 °C, 5V2O5/PG-Re-400 °C, and 5V2O5/PG-Re-500 °C for Hg0 removal are shown in Figure 8. It can be seen that, with the rise of regeneration temperature (from 300 °C to 500 °C), the EHg of 5V2O5/PG-Re-x decreased obviously. 5V2O5/PG-Re-300 °C showed the highest activity (EHg = 90.3%) at 420 min, even higher than that of the fresh 5V2O5/PG catalyst (EHg = 83.1%). The EHg of 5V2O5/PG-Re-400 °C and 5V2O5/PG-Re-500 °C were 66.7% and 46.8%, respectively. This may be due to the change of structure and chemical properties of 5V2O5/PG during regeneration, which caused partial active sites loss on the surface of the 5V2O5/PG catalyst at higher regeneration temperatures (400 °C and 500 °C) [36].

3. Materials and Methods

3.1. Catalysts Preparation

The PG was mixed with distilled water in a certain proportion (1 g: 4–8 mL), dried at 110 °C for 6 h, then calcined at 300 °C for 6 h in air. The obtained PG samples were crushed and screened into a 40–60 mesh.
The xV2O5/PG catalysts (x is the mass fraction of V2O5 in the V2O5/PG catalyst, and x = 0–5 wt%) were prepared by impregnating PG particles with different V2O5 loadings, which was similar to the V2O5/AC catalysts in our previous research [36]., Briefly, the PG particles are impregnated in ammonium metavanadate solution with the required concentration for 1 h, dried at 60 °C for 5 h, then 110 °C for 8 h. Finally, the samples were calcined at 300 °C for 4 h in air.

3.2. Catalytic Activity Evaluation of V2O5/PG

The catalytic activity of the V2O5/PG catalyst for Hg0 removal was tested in a bench-scale fixed-bed reactor; a detailed description of the experimental schematic is in Figure 9. The Hg0 removal activity evaluation device system mainly includes simulated flue gas, a programmed temperature controller, a fixed-bed reactor, an online mercury analyzer and a tail gas treatment cleaner. A total of 0.5 g V2O5/PG was loaded into the quartz tube. The simulated flue gas includes 8% O2, 0.15% SO2, 5% H2O and balance N2, and was controlled by a mass flowmeter. The GHSV was approximately 6000 h−1. The temperature (120–210 °C) was controlled by a digital temperature controller. The Hg0 vapor was produced from a mercury permeation tube (VICI Metronics) and carried out by N2 with a constant flow rate. Hg0 concentration was detected by an online RA-915M Mercury Analyzer (Lumex Co, Ltd., St. Petersburg, Russia). Finally, the tail gas was treated by a device equipped with activated carbon.
The Hg0 removal efficiency (EHg) was defined as follows:
E Hg ( % ) = C 0   C 1 C 0 × 100   %
where C0 and C1 represent the Hg0 concentration (μg/m3) at the inlet and outlet of the reactor, respectively.

3.3. Characterization

The N2 adsorption–desorption tests were carried out by an Autosorb-iQ analyzer (Quantachrome, Boynton Beach, FL, USA). The specific surface areas were calculated by the Brunauer–Emmett–Teller (BET) method, the pore structure parameters were analyzed by the Barrett–Joyner–Halenda (BJH) method.
The morphologies of the V2O5/PG samples were observed, which was performed on a scanning electron microscope (SEM) (JSM-7001F, JEOL, Akishima City, Japan).
The X-ray photoelectron spectroscopy (XPS) analysis was performed on an ESCALAB 250Xi spectrometer (Thermo Fisher, Waltham, MA, USA) using an Al Kα X-ray source at room temperature. All binding energies (BE) were adjusted with the C 1s binding energy value of 284.6 eV.
The Hg temperature-programmed desorption (Hg-TPD) experiments were conducted in a quartz tube reactor using a 0.1 g sample in N2 (200 mL/min). The sample was firstly used to remove Hg0 at 150 °C for 6 h, then swept with N2 at 150 °C for 2 h, and finally, heated from 30 °C to 600 °C with a heating rate of 5 °C/min. The outlet gas from the reactor was introduced into a KBH4 solution to reduce the possibly existing Hg2+ to Hg0. The Hg0 concentration in the effluent gas after the KBH4 solution was continuously measured by an on-line mercury analyzer (RA-915M, Lumex, St. Petersburg, Russia).

4. Conclusions

The V2O5/PG catalyst had excellent Hg0 removal capability, which was mainly due to the V2O5/PG catalyst combined with the adsorption ability of PG and the catalytic oxidation activity of V2O5. Hg0 was oxidized to form HgO and HgSO4, and then adsorbed on the V2O5/PG catalyst. The EHg of the V2O5/PG catalyst increased with the increase of V2O5 loading, and the 5V2O5/PG catalyst showed the highest EHg at 150 °C. O2 exhibited a promoting effect on Hg0 removal, while SO2 and H2O showed an obvious inhibitory effect. However, when O2, H2O, and SO2 were added together, the EHg showed a promoting trend. The used V2O5/PG catalyst after Hg0 removal can be regenerated and its capability for Hg0 removal can be completely recovered, and the V2O5/PG-Re-300 °C catalyst showed excellent performance and good stability for Hg0 removal after regeneration.

Author Contributions

Conceptualization, J.W. and X.W.; methodology, J.W.; validation, H.W., C.J. and J.W.; formal analysis, H.W.; investigation, H.W.; data curation, C.J.; writing—original draft preparation, J.W.; writing—review and editing, J.Z.; visualization, H.W.; supervision, J.W.; project administration, X.W.; funding acquisition, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (21203003, 51404014), Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (2020-KF-28) and Anhui Provincial Discipline (Professional) Top Talent Academic Funding Project (gxbjZD2021062).

Data Availability Statement

Not available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Horowitz, H.M.; Jacob, D.J.; Amos, H.M.; Streets, D.G.; Sunderland, E.M. Historical mercury releases from commercial products: Global environmental implications. Environ. Sci. Technol. 2014, 48, 10242–10250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nriagu, J.O.; Pacyna, J.M. Quantitative assessment of worldwide contamination of air, water and soils by trace metals. Nature 1988, 333, 134–139. [Google Scholar] [CrossRef] [PubMed]
  3. Gao, Y.S.; Zhang, Z.; Wu, J.W.; Duan, L.H.; Umar, A.; Sun, L.H. A critical review on the heterogeneous catalytic oxidation of elemental mercury in flue gases. Environ. Sci. Technol. 2013, 47, 10813–10823. [Google Scholar] [CrossRef] [PubMed]
  4. Kocman, D.; Horvat, M.; Pirrone, N.; Cinnirella, S. Contribution of contaminated sites to the global mercury budget. Environ. Res. 2013, 125, 160–170. [Google Scholar] [CrossRef] [PubMed]
  5. Li, H.L.; Wu, C.Y.; Li, Y.; Li, L.Q.; Zhao, Y.C.; Zhang, J.Y. Impact of SO2 on elemental mercury oxidation over CeO2-TiO2 catalyst. Chem. Eng. J. 2013, 219, 319–326. [Google Scholar] [CrossRef]
  6. Streets, D.G.; Hao, J.M.; Wu, Y.; Jiang, J.K.; Chan, M.; Tian, H.Z. Anthropogenic mercury emissions in China. Atmos. Environ. 2005, 39, 7789–7806. [Google Scholar] [CrossRef] [Green Version]
  7. Wu, Q.R.; Wang, S.X.; Liu, K.Y.; Li, G.L.; Hao, J.M. Emission-limit-oriented strategy to control atmospheric mercury emissions in coal-fired power plants toward the implementation of the Minamata Convention. Environ. Sci. Technol. 2018, 52, 11087–11093. [Google Scholar] [CrossRef]
  8. Zhao, Y.; Ma, X.Y.; Xu, P.Y.; Wang, H.; Liu, Y.C.; He, A.E. Elemental mercury removal from flue gas by CoFe2O4 catalyzed peroxymonosulfate. J. Hazard. Mater. 2018, 341, 228–237. [Google Scholar] [CrossRef]
  9. 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]
  10. Luo, J.J.; Zhang, L.D.; Huang, H.W.; Zhang, J.R. Effects of flue gas components and fly ash on mercury oxidation. J. Univ. Sci. Technol. Beijing 2011, 33, 771–776. [Google Scholar]
  11. Yang, Y.J.; Liu, J.; Liu, F.; Wang, Z.; Miao, S. Molecular-level insights into mercury removal mechanism by pyrite. J. Hazard. Mater. 2018, 344, 104–112. [Google Scholar] [CrossRef] [PubMed]
  12. Zhou, Z.J.; Liu, X.W.; Zhao, B.; Shao, H.Z.; Xu, Y.S.; Xu, M.H. Elemental mercury oxidation over manganese-based perovskite-type catalyst at low temperature. Chem. Eng. J. 2016, 288, 701–710. [Google Scholar] [CrossRef]
  13. Dίaz-Somoano, M.; Unterberger, S.; Hein, K.R. Mercury emission control in coal-fired plants: The role of wet scrubbers. Fuel Process. Technol. 2007, 88, 259–263. [Google Scholar] [CrossRef]
  14. Li, B.; Wang, H.L. Effect of flue gas purification facilities of coal-fired power plant on mercury emission. Energy Rep. 2021, 7, 190–196. [Google Scholar] [CrossRef]
  15. Li, H.; Wu, C.Y.; Li, Y.; Zhang, J.Y. CeO2-TiO2 catalysts for catalytic oxidation of elemental mercury in low-rank coal combustion flue gas. Environ. Sci. Technol. 2011, 45, 7394–7400. [Google Scholar] [CrossRef]
  16. Fukuda, N.; Takaoka, M.; Doumoto, S.; Oshita, K.; Morisawa, S.; Mizuno, T. Mercury emission and behavior in primary ferrous metal production. Atmos. Environ. 2011, 45, 3685–3691. [Google Scholar] [CrossRef]
  17. Zhao, S.J.; Mei, J.; Xu, H.M.; Liu, W.; Qu, Z.; Cui, Y. Research of mercury removal from sintering flue gas of iron and steel by the open metal site of Mil-101(Cr). J. Hazard. Mater. 2018, 351, 301–307. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, H.Q.; Xu, Z.H.; Fan, M.H.; Bland, A.E.; Judkins, R.R. Adcatalyst for capturing mercury in coal-fired boiler flue gas. J. Hazard. Mater. 2007, 146, 1–11. [Google Scholar] [CrossRef]
  19. Scala, F. Simulation of mercury capture by activated carbon injection in incinerator flue gas in-duct removal. Environ. Sci. Technol. 2001, 35, 4367–4372. [Google Scholar] [CrossRef]
  20. Chen, W.M.; Pei, Y.; Huang, W.J.; Qu, Z.; Hu, X.F.; Yan, N.Q. Novel effective catalyst for elemental mercury removal from coal-fired flue gas and the mechanism investigation. Environ. Sci. Technol. 2016, 50, 2564–2572. [Google Scholar] [CrossRef]
  21. Yuan, Y.; Zhang, J.Y.; Li, H.L.; Li, Y.; Zhao, Y.C.; Zheng, C.G. Simultaneous removal of SO2, NO and mercury using TiO2-aluminum silicate fiber by photocatalysis. Chem. Eng. J. 2012, 192, 21–28. [Google Scholar] [CrossRef]
  22. Wu, J.; Xu, K.; Liu, Q.Z.; Ji, Z.; Qu, C.; Qi, X.M. Controlling dominantly reactive (010) facets and impurity level by in-situ reduction of BiOIO3 for enhancing photocatalytic activity. Appl. Catal. B 2018, 232, 135–145. [Google Scholar] [CrossRef]
  23. Pavlish, J.H.; Sondreal, E.A.; Mann, M.D.; Olson, E.S.; Galbreath, K.C.; Laudal, D.L. Status review of mercury control options for coal-fired power plants. Fuel Process. Technol. 2003, 82, 89–165. [Google Scholar] [CrossRef]
  24. Xu, H.M.; Qu, Z.; Zong, C.X.; Quan, F.Q.; Mei, J.; Yan, N.Q. 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]
  25. Cai, J.; Shen, B.X.; Li, Z.; Chen, J.H.; He, C. Removal of elemental mercury by clays impregnated with KI and KBr. Chem. Eng. J. 2014, 241, 19–27. [Google Scholar] [CrossRef]
  26. He, C.; Shen, B.X.; Chen, J.H.; Cai, J. Adsorption and oxidation of elemental mercury over Ce-MnOx/Ti-PILCs. Environ. Sci. Technol. 2014, 48, 7891–7898. [Google Scholar] [CrossRef]
  27. Ding, F.; Zhao, Y.C.; Mi, L.L.; Li, H.L.; Li, Y.; Zhang, J.Y. Removal of gas-phase elemental mercury in flue gas by inorganic chemically promoted natural mineral catalyst. Ind. Eng. Chem. Res. 2012, 51, 3039–3047. [Google Scholar] [CrossRef]
  28. Li, H.L.; Zhang, M.G.; Zhu, L.; Yang, J.P. Stability of mercury on a novel mineral sulfide sorbent used for efficient mercury removal from coal combustion flue gas. Environ. Sci. Pollut. Res. 2018, 25, 28583–28593. [Google Scholar] [CrossRef]
  29. Takagi, M.; Kawai, T.; Soma, M.; Onishi, T.; Tamaru, K. The mechanism of the reaction between NOx and NH3 on V2O5 in the presence of oxygen. J. Catal. 1977, 50, 441–446. [Google Scholar] [CrossRef]
  30. Zhao, L.K.; Li, C.T.; Zhang, J.; Zhang, X.N.; Zhan, F.M.; Ma, J.F. Promotional effect of CeO2 modified support on V2O5–WO3/TiO2 catalyst for elemental mercury oxidation in simulated coal-fired flue gas. Fuel 2015, 153, 361–369. [Google Scholar] [CrossRef]
  31. Lee, W.; Bae, G.N. Removal of elemental mercury (Hg0) by nanosized V2O5/TiO2 catalysts. Environ. Sci. Technol. 2009, 43, 1522–1527. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, X.P.; Li, Z.F.; Wang, J.X.; Tan, B.J.; Cui, Y.Z.; He, G.H. Reaction mechanism for the influence of SO2 on Hg0 adsorption and oxidation with Ce-0.1-Zr-MnO2. Fuel 2017, 203, 308–315. [Google Scholar] [CrossRef]
  33. Zhang, A.C.; Xing, W.B.; Zhang, Z.H.; Meng, F.M.; Liu, Z.C.; Xiang, J. Promotional effect of SO2 on CeO2-TiO2 material for elemental mercury removal at low temperature. Atmos. Pollut. Res. 2016, 7, 895–902. [Google Scholar] [CrossRef]
  34. Yang, Z.Q.; Li, H.L.; Liu, X.; Li, P.; Yang, J.P.; Lee, P.H. 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]
  35. Wang, J.W.; Shen, Y.Y.; Dong, Y.J.; Qin, W.; Zhang, Q.P.; Lu, L.; Zhang, Y.G. Oxidation and adsorption of gas-phase Hg0 over a V2O5/AC catalyst. RSC Adv. 2016, 6, 77553–77557. [Google Scholar] [CrossRef]
  36. Zhang, Q.P. Removal of Hg0 from Simulated Flue Gas by CuO/PG and V2O5/PG. Master’s Thesis, Anqing Normal University, Anqing, China, 2018. (In Chinese). [Google Scholar]
  37. Zhu, Y.C.; Han, X.J.; Huang, Z.G.; Hou, Y.Q.; Guo, Y.; Wu, M.H. Superior activity of CeO2 modified V2O5/AC catalyst for mercury removal at low temperature. Chem. Eng. J. 2018, 337, 741–749. [Google Scholar] [CrossRef]
  38. Ma, Y.P.; Mu, B.L.; Yuan, D.L.; Zhang, H.Z.; Xu, H.M. Design of MnO2/CeO2-MnO2 hierarchical binary oxides for elemental elemental mercury removal from coal-fired flue gas. J. Hazard. Mater. 2017, 333, 186–193. [Google Scholar] [CrossRef]
  39. Ye, D.; Wang, X.X.; Wang, R.X.; Wang, S.Y.; Liu, H.; Wang, H.N. Recent advances in MnO2-based adsorbents for mercury removal from coal-fired flue gas. J. Environ. Chem. Eng. 2021, 9, 105993. [Google Scholar] [CrossRef]
  40. Zhu, Y.C.; Hou, Y.Q.; Wang, J.W.; Guo, Y.P.; Huang, Z.G.; Han, X.J. Effect of SCR atmosphere on the removal of Hg0 by a V2O5–CeO2/AC catalyst at low temperature. Environ. Sci. Technol. 2019, 53, 5521–5527. [Google Scholar] [CrossRef]
  41. Li, H.L.; Wu, C.Y.; Li, Y.; Li, L.Q.; Zhao, Y.C.; Zhang, J.Y. Role of flue gas components in mercury oxidation over TiO2 supported MnOx-CeO2 mixed-oxide at low temperature. J. Hazard. Mater. 2012, 243, 117–123. [Google Scholar] [CrossRef]
  42. He, C.; Shen, B.X.; Li, F.K. Effects of flue gas components on removal of elemental mercury over Ce–MnOx/Ti-PILCs. J. Hazard. Mater. 2016, 304, 10–17. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, Y.Y.; Shen, B.X.; He, C.; Yue, S.J.; Wang, F.M. Simultaneous removal of NO and Hg0 from flue gas over Mn-Ce/Ti-PILCs. Environ. Sci. Technol. 2015, 49, 9355–9363. [Google Scholar] [CrossRef]
  44. Zhao, L.K.; Li, C.T.; Li, S.H.; Wang, Y.; Zhang, J.Y.; Wang, T. Simultaneous removal of elemental mercury and NO in simulated flue gas over V2O5/ZrO2-CeO2 catalyst. Appl. Catal. B 2016, 198, 420–430. [Google Scholar] [CrossRef]
  45. Morris, E.A.; Kirk, D.W.; Jia, C.Q.; Morita, K. Roles of sulfuric acid in elemental mercury removal by activated carbon and sulfur-impregnated activated carbon. Environ. Sci. Technol. 2012, 46, 7905–7912. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, Q.; Duan, Y.F.; Chen, M.M.; Liu, M.; Lu, P. Studies on mercury adsorption species and equilibrium on activated carbon surface. Energy Fuels 2017, 31, 14211–14218. [Google Scholar] [CrossRef]
  47. Rumayor, M.; Diaz-Somoano, M.; López-Antón, M.A. Temperature programmed desorption as a tool for the identification of mercury fate in wet-desulphurization systems. Fuel 2015, 148, 98–103. [Google Scholar] [CrossRef] [Green Version]
  48. Ranea, V.A. A DFT + U study of H2O adsorption on the V2O5(0 0 1) surface including van der Waals interactions. Chem. Phys. Lett. 2019, 730, 171–178. [Google Scholar] [CrossRef]
Figure 1. Effect of temperature on Hg0 removal over 5V2O5/PG (reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 120–210 °C).
Figure 1. Effect of temperature on Hg0 removal over 5V2O5/PG (reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 120–210 °C).
Catalysts 12 00243 g001
Figure 2. Effect of V2O5 loading on Hg0 removal (reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 150 °C).
Figure 2. Effect of V2O5 loading on Hg0 removal (reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 150 °C).
Catalysts 12 00243 g002
Figure 3. SEM morphology of 5V2O5/PG (A): ×10,000; (B): ×25,000.
Figure 3. SEM morphology of 5V2O5/PG (A): ×10,000; (B): ×25,000.
Catalysts 12 00243 g003
Figure 4. Effect of flue gas components on Hg0 removal over 5V2O5/PG ((A) single component; (B) multi-components. reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 150 °C).
Figure 4. Effect of flue gas components on Hg0 removal over 5V2O5/PG ((A) single component; (B) multi-components. reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 150 °C).
Catalysts 12 00243 g004
Figure 5. Effect of GHSV on Hg0 removal over 5V2O5/PG (reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000–15,000 h−1, T = 150 °C).
Figure 5. Effect of GHSV on Hg0 removal over 5V2O5/PG (reaction conditions: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000–15,000 h−1, T = 150 °C).
Catalysts 12 00243 g005
Figure 6. XPS of Hg 4f for fresh and spent V2O5/PG catalyst.
Figure 6. XPS of Hg 4f for fresh and spent V2O5/PG catalyst.
Catalysts 12 00243 g006
Figure 7. TPD profiles of fresh and spent 5V2O5/PG (reaction conditions: T = 30–600 °C, N2 = 200 mL/min, heating rate = 5 °C/min).
Figure 7. TPD profiles of fresh and spent 5V2O5/PG (reaction conditions: T = 30–600 °C, N2 = 200 mL/min, heating rate = 5 °C/min).
Catalysts 12 00243 g007
Figure 8. Comparison of Hg0 removal over PG, 5V2O5/PG, 5V2O5/PG-Re-300 °C, 5V2O5/PG-Re-400 °C, and 5V2O5/PG-Re-500 °C (reaction condition: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 150 °C).
Figure 8. Comparison of Hg0 removal over PG, 5V2O5/PG, 5V2O5/PG-Re-300 °C, 5V2O5/PG-Re-400 °C, and 5V2O5/PG-Re-500 °C (reaction condition: O2 = 8%, H2O = 5%, SO2 = 0.15%, N2 as balance, CHg0 = 240 μg/m3, GHSV = 6000 h−1, T = 150 °C).
Catalysts 12 00243 g008
Figure 9. Schematic diagram of the fixed-bed reactor for Hg0 removal by V2O5/PG.
Figure 9. Schematic diagram of the fixed-bed reactor for Hg0 removal by V2O5/PG.
Catalysts 12 00243 g009
Table 1. Properties of the PG and V2O5/PG catalysts.
Table 1. Properties of the PG and V2O5/PG catalysts.
SampleBET Surface Area
(m2/g)
Average Pore Size
(nm)
Total Pore Volume
(cm3/g)
PG151.4310.510.32
1V2O5/PG139.9110.670.30
3V2O5/PG130.2410.850.26
5V2O5/PG109.3111.040.21
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, J.; Wang, H.; Jiang, C.; Wang, X.; Zhang, J. Hg0 Removal by V2O5 Modified Palygorskite in Simulated Flue Gas at Low Temperature. Catalysts 2022, 12, 243. https://doi.org/10.3390/catal12020243

AMA Style

Wang J, Wang H, Jiang C, Wang X, Zhang J. Hg0 Removal by V2O5 Modified Palygorskite in Simulated Flue Gas at Low Temperature. Catalysts. 2022; 12(2):243. https://doi.org/10.3390/catal12020243

Chicago/Turabian Style

Wang, Junwei, Huan Wang, Caihong Jiang, Xie Wang, and Jianli Zhang. 2022. "Hg0 Removal by V2O5 Modified Palygorskite in Simulated Flue Gas at Low Temperature" Catalysts 12, no. 2: 243. https://doi.org/10.3390/catal12020243

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