Next Article in Journal
Patternable Poly(chloro-p-xylylene) Film with Tunable Surface Wettability Prepared by Temperature and Humidity Treatment on a Polydimethylsiloxane/Silica Coating
Previous Article in Journal
A Review of Mechanoluminescence in Inorganic Solids: Compounds, Mechanisms, Models and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism of Mercury Adsorption and Oxidation by Oxygen over the CeO2 (111) Surface: A DFT Study

1
National Engineering Laboratory for Biomass Power Generation Equipment, North China Electric Power University, Beijing 102206, China
2
Cain Department of Chemical Engineering, Louisiana State University, Baton Rouge, LA 70820, USA
*
Author to whom correspondence should be addressed.
Materials 2018, 11(4), 485; https://doi.org/10.3390/ma11040485
Submission received: 8 February 2018 / Revised: 20 March 2018 / Accepted: 20 March 2018 / Published: 23 March 2018
(This article belongs to the Section Catalytic Materials)

Abstract

:
CeO2 is a promising catalytic oxidation material for flue gas mercury removal. Density functional theory (DFT) calculations and periodic slab models are employed to investigate mercury adsorption and oxidation by oxygen over the CeO2 (111) surface. DFT calculations indicate that Hg0 is physically adsorbed on the CeO2 (111) surface and the Hg atom interacts strongly with the surface Ce atom according to the partial density of states (PDOS) analysis, whereas, HgO is adsorbed on the CeO2 (111) surface in a chemisorption manner, with its adsorption energy in the range of 69.9–198.37 kJ/mol. Depending on the adsorption methods of Hg0 and HgO, three reaction pathways (pathways I, II, and III) of Hg0 oxidation by oxygen are proposed. Pathway I is the most likely oxidation route on the CeO2 (111) surface due to it having the lowest energy barrier of 20.7 kJ/mol. The formation of the HgO molecule is the rate-determining step, which is also the only energy barrier of the entire process. Compared with energy barriers of Hg0 oxidation on the other catalytic materials, CeO2 is more efficient at mercury removal in flue gas owing to its low energy barrier.

Graphical Abstract

1. Introduction

Mercury, a pollutant harmful to human health and the environment, has become a global concern due to its toxicity, high volatility, and bioaccumulation [1,2,3,4]. In October 2013, the first legally-binding international treaty, called the Minamata Convention, was concluded to limit global emissions of mercury [5,6]. The convention entered into force on 16 August 2017. Coal-fired power plants are regarded as major sources of atmospheric mercury emissions [7,8]. The latest version of the Emission Standard of Air Pollutants for Thermal Power Plant (GB13223-2011), released by the Ministry of Environmental Protection of China, set the limit for the emission of mercury and its compounds from coal-fired boilers to 0.03 mg/m3. Therefore, mercury removal technology for coal-fired flue gas is urgently needed.
Mercury in coal-fired flue gas mainly exists in three forms, i.e., elemental mercury (Hg0), oxidized mercury (Hg2+), and particle mercury (HgP) [9]. Hg2+ is soluble in water and can be removed primarily in the form of HgCl2 by the wet desulfurization system (WFGD) in coal-fired power plants [10]. HgP can be removed with the fly ash in fabric filters (FF) and electrostatic precipitators (ESP) [11]. Hg0 is the main form of mercury released into the atmosphere, and is difficult to remove using existing pollutant control equipment in power plants due to its chemical inertness and insolubility in water [12]. The main solution for Hg0 removal is by the adsorption and oxidation method [13]. The injection of activated carbon or other adsorbents for Hg0 adsorption in flue gas greatly increases the cost of power plant because of the limited capacity and high cost of the adsorbent. On the other hand, mercury’s oxidation to readily captured Hg2+ by catalyst needs no installation of new equipment and can save significant transformation and operating costs compared to sorbents injection, which makes it easier to accept [14,15]. Hence, the development of efficient Hg0 oxidation catalysts is of great significance.
CeO2 is a promising catalytic oxidation material with low cost, no toxicity, and a large oxygen storage capacity because it contains the unique redox couple Ce3+/Ce4+, which can shift from CeO2 to Ce2O3 and vice versa, under oxidizing and reducing conditions, respectively [16,17]. Experimental results from Li et al. and Fan et al. showed that CeO2 had a satisfactory catalytic performance in the oxidation of Hg0 and maintains over 90% oxidation efficiency under simulated flue gas conditions [16,18]. Zhao et al. investigated the modification of a commercial SCR catalyst with a series of metal oxides and found that the SCR catalyst doped with CeO2 exhibited the highest Hg0 oxidation ability [19]. Theoretically, Liu et al. employed the DFT method to study the reaction mechanism of Hg0 oxidation by HCl on the CeO2 (111) surface, verifying the catalytic activity of CeO2 for Hg0 oxidation theoretically by analyzing energy barriers. It was found that the energy barrier of forming HgCl2 on the CeO2 surface was close to that on noble metals such as Au and Pd. The low energy barrier makes CeO2 an attractive catalyst for Hg0 oxidation, while HCl plays an important role in the oxidation process [20]. However, the content of Cl in Chinese coal (63–318 mg/kg) is very low, and thus, the even lower content of HCl in flue gas leads to catalysts’ lack of Hg0 oxidation activity [21]. Accordingly, it is necessary to study the oxidation of Hg0 in the absence of HCl and improve the oxidation ability of Hg0 when the catalyst is installed in flue gas conditions with insufficient HCl content.
As is shown in recent work, CeO2 can catalyze the oxidation of Hg0 in flue gas condition without HCl and obtain a high efficiency of Hg0 oxidation with the aid of oxygen [7,22,23]. Li et al. synthesized a series of CeO2–TiO2 catalysts by an ultrasound-assisted impregnation method. They found that the introduction of 4% O2 into the gas flow containing SO2 and NO resulted in an Hg0 oxidation efficiency as high as 99.9% [7]. He and his co-workers studied the reactivity of CeO2 catalyst based on pillared clay(PILC)–TiO2, and the results demonstrated that 15% Ce/Ti-PILC catalyst had the best Hg0 oxidation efficiency (88.6%) at 300 °C under a 5% O2 + N2 atmosphere [22]. Zhao et al. studied the catalytic effect of a CeO2-doped V2O5–WO3/TiO2 catalyst under flue gas conditions without HCl and found that the highest oxidation efficiency of Hg0 was 88.93% at the optimum temperature of 250 °C. XPS results showed that Hg0 and HgO are the main mercury species on the catalyst surface, indicating that Hg0 was oxidized to HgO by the oxygen atoms on the surface of the catalyst, and the oxidation process should follow the Mars–Maessen mechanism [23].
It is generally believed that there are four main mechanisms of Hg0 heterogeneous oxidation hitherto, i.e., the Deacon process, the Mars–Maessen mechanism, the Langmuir–Hinshewood mechanism, and the Eley–Rideal mechanism [24]. Recent DFT research indicates that the heterogeneous oxidation of Hg0 and HCl on the catalyst surface follows the Langmuir–Hinshewood mechanism [10,25,26,27,28]. Nevertheless, to the best of our knowledge, few theoretical calculations were reported to investigate the reaction mechanism between Hg0 and oxygen on the catalyst surface in detail, and there is no quantum chemistry study concerning the Hg + O2 → HgO progress on the CeO2 (111) surface to date. Based on previous work, density functional theory calculations are performed in this study to investigate the mechanism of mercury adsorption and oxidation by oxygen on the CeO2 (111) surface, and the reaction pathway, energy barriers, and transition state configurations of interaction between mercury and oxygen are also studied. The objective of this work is to provide theoretical guidance for the Hg0 catalytic oxidation without HCl.

2. Computational Details

2.1. Catalyst Models

The crystal structure of CeO2 is of a cubic fluorite structure with a space group Fm-3m. It contains four Ce atoms and eight O atoms in a primary cell, as shown in Figure 1.
Previous research indicates that CeO2 (111) is a typical low index surface with the most stable thermodynamic properties, and also has the closest physical and chemical properties to a real crystal surface [20]. Hence the CeO2 (111) surface was constructed by cleaving the optimized unit cell. In this work, p (2 × 2) and p (3 × 3) supercell periodic slab models with nine atomic layers were constructed. The bottom six layers were fixed in their bulk positions and the top three layers were fully relaxed for geometry optimization. A 12 Å-thick vacuum region was set so that the energy effect of interactions between slabs can be neglected. The energy effect of neighboring Hg atoms was tested by comparing the adsorption energies of Hg on p (2 × 2) and p (3 × 3) surfaces. The results indicate that the geometry parameters and adsorption energies of Hg adsorption are close (−5.71 kJ/mol vs. −6.76 kJ/mol), which is consistent with previous work [20]. Therefore, the p (2 × 2) model is used to simulate the CeO2 surface. The constructed CeO2 (111) surface model is depicted in Figure 2.

2.2. Computational Method

All density functional theory calculations in this study were performed using DMol3 code [29]. The exchange-correlation potential was calculated by the Perduw-Burke-Ernzerhof (PBE) function in a generalized gradient approximation (GGA) scheme [30,31]. The core electrons of Hg and Ce atoms were treated by the DFT semi-core pseudopotential (DSPP) method, in which the relativistic effect of core electrons was concerned, while the core electrons of O atoms were treated by the all-electron method [32]. A 4.5 Å global orbital cutoff was selected with a smearing value of 0.005 Ha to accelerate the convergence of calculations. The molecular orbitals are expanded by a double numerical basis set with polarization functions (DNP).
To properly account for the band structure of ceria, Hubbard U corrections to the f electrons have been applied in some calculations on the CeO2 surface [33,34]. However, previous results indicate that plain DFT calculations can provided a reasonable prediction of reduction energies, even better than that from DFT + U [34,35,36,37,38]. Moreover, the effect of U correction on reactions on stoichiometric CeO2 (111) surface is relatively not significant [39,40], which makes this approach less cost-effective. Hence, the DFT+U method was not considered in this study.
3 × 3 × 1 Monkhorst–Pack k-point sampling was selected for Brillouin zone integration during geometric optimization calculation. The convergence criteria of energy, force, and displacement are less than 10−5 Hartree, 0.002 Ha/atom and 0.005 Å, respectively.
To confirm the accuracy of the above parameters, a 4 × 4 × 4 k point mesh was first performed for geometry optimization of CeO2 unit cell. The optimized bulk lattice parameters (a = b = c = 5.485 Å) are in good agreement with the experimental values and calculated values reported in the literature [41]. The deviation is found to be minimal, which suggests that the calculations are reliable. The adsorption energy on CeO2 (111) surface (Eads) is defined as follows:
E a d s = E a d s o r b a t e s u b s t r a t e E a d s o r b a t e E s u b s t r a t e ,
where E a d s o r b a t e s u b s t r a t e represents the adsorption configuration on CeO2 (111) surface, E a d s o r b a t e and E s u b s t r a t e represent the total energy of adsorbate and CeO2 (111) surface, respectively. As shown in Equation (1), a negative value of adsorption energy refers to an endothermic reaction, while a positive value refers to an exothermic reaction. In addition, the more negative the adsorption energy, the more stable the adsorption configuration is.
The oxidation pathway of Hg0 includes intermediate (IM), transition state (TS) and final state (FS). All TSs were obtained by the complete linear synchronous transit and quadratic synchronous transit (LST/QST) method [42]. Vibrational frequencies were calculated at the optimized geometries to identify the nature of the stationary points (no imaginary frequency) and the transition state (only one imaginary frequency). The energy barriers of each reaction pathways are defined by the following equation:
E b a r r i e r = E t r a n s i t i o n   s t a t e E i n t e r m e d i a t e ,
where E t r a n s i t i o n   s t a t e and E i n t e r m e d i a t e denote the total energies of transition state and intermediate, respectively.

3. Results and Discussion

3.1. Adsorption of Hg0 on the CeO2 (111) Surface

Hg0 adsorption on the catalyst surface is the first step in heterogeneous mercury oxidation. All possible adsorption sites were considered on the CeO2 (111) surface and four stable configurations (Ce site, O-top site, O-sub site, and O-bridge site) are shown in Figure 3 as 1A, 1B, 1C, and 1D. The adsorption energy, Mulliken charge of Hg atom and corresponding geometry parameters of 1A, 1B, 1C, and 1D are listed in Table 1.
According to Figure 3 and Table 1, all the adsorption energy values are negative, indicating that the Hg0 adsorption on the CeO2 (111) surface is an exothermic process. In configurations 1B, 1C, and 1D, the adsorption of Hg atoms form Hg–O bonds with surface O atoms with lengths of 3.053 Å, 3.616/3.604 Å, and 4.360 Å, respectively. The adsorption energies are −1.411 kJ/mol, −5.30 kJ/mol, and −5.25 kJ/mol, respectively, suggesting a weak interaction. About 0.008 and 0.011 e charges of Hg atom are transferred to the surface in adsorption, suggesting that few electrons are transferred to the substrate, which is consistent with the results of adsorption energy. Regarding configuration 1A, the bond length between Hg0 and surface Ce atom is 3.797 Å, and the Mulliken charge of the Hg atom is about 0.012 e. Meanwhile, it has the most negative adsorption energy of −5.71 kJ/mol. Therefore, it is the most stable configuration for Hg0 adsorption and the adsorption mechanism is physisorption. The calculated results agree well with previous theoretical research [20].
To further investigate the interaction between the Hg atom and the CeO2 (111) surface during adsorption, the partial density of states (PDOS) analysis of Hg and Ce atoms are performed in the most stable adsorption configuration of 1A. The PDOS results of Hg pre-adsorption, Hg post-adsorption, Ce pre-adsorption, and Ce post-adsorption are presented in Figure 4. As is depicted in Figure 4, for pre-adsorption, the Hg s-orbital is occupied at 0.49 Ha and the Fermi level. The unoccupied p-orbital shows a single peak at 0.22 Ha, while the Hg d-orbital is occupied at −0.13 and 0.57 Ha. After Hg adsorption, all orbitals of the Hg atom are shifted to the lower energy level with s- and p-orbitals broadened and there is a decrease in energy due to the charge transfer from the Hg atom to the surface Ce atom. Meanwhile, the disappearance of the Hg d-orbital located at 0.57 Ha indicates a strong interaction between the Hg atom and the CeO2 (111) surface. As for the energy bands located between −0.7 Ha and 0.1 Ha of the surface Ce atom, an unapparent change occurs after adsorption in all orbitals. It can be concluded from the above analysis that Hg has a strong interaction with the CeO2 (111) surface after adsorption. Hg orbitals change significantly, while no apparent change occurs for the surface Ce atom, indicating that the CeO2 (111) surface can remain stable after Hg0 adsorption.

3.2. Adsorption of HgO on the CeO2 (111) Surface

Heterogeneous oxidation of Hg0 will form HgO on the catalyst surface. Similar to Hg0, the adsorption behavior of HgO was also investigated on the CeO2 (111) surface. All possible adsorption orientations, including parallel and perpendicular, were considered. After geometry optimization, four stable configurations, 2A, 2B, 2C, and 2D, are presented in Figure 5 and the corresponding adsorption energy and geometric parameters are given in Table 2. The configuration 2D with an adsorption energy of −198.37 kJ/mol is the most stable. HgO molecule is dissociatively adsorbed on the CeO2 (111) surface with a breakage of the Hg–O bond. The O atom of the HgO molecule forms a 1.415 Å O–O bond with the O-top atom on the CeO2 (111) surface. In the configuration 2A, the O-end of HgO is vertically adsorbed on the Ce atom of the CeO2 (111) surface, forming a 2.153 Å Ce–O bond. The length of Hg–O slightly decreases from 2.208 Å to 2.145 Å, and the adsorption energy of 2A is 69.90 kJ/mol. In terms of the configuration 2B, HgO molecules are adsorbed on the CeO2 (111) surface in an approximately parallel manner. The length of the Hg–O bond is 2.277 Å, while the distances between Hg, O, and the CeO2 (111) surface are 2.114 Å and 2.400 Å, respectively. The corresponding adsorption energy is −101.99 kJ/mol. As for the 2C configuration, the Hg end of HgO molecule is adsorbed perpendicularly to the O-top atom on the surface with an adsorption energy of −163.81 kJ/mol. The length of Hg–O decreases to 1.945 Å and the Hg–O-top distance is 2.077 Å.
Investigation of HgO adsorption on the CeO2 (111) surface indicates that the adsorption manner of HgO is chemisorption due to its strong adsorption energy. Hence, it is difficult for HgO to separate from the CeO2 (111) surface. In the most stable adsorption configuration, 2D, HgO dissociates on the surface with an adsorption energy of −198.37 kJ/mol; the Hg–O bond in the HgO molecule breaks and the O atom is bonded with the surface oxygen. Apart from 2D, 2C has the most negative adsorption energy of −163.81 kJ/mol. Based on the calculated results, the interaction of HgO with oxygen on the CeO2 (111) surface is relatively strong, which is instructive for the consideration of reaction sites on the surface.

3.3. Hg0 Oxidation Mechanism on the CeO2 (111) Surface

The surface oxygen on the catalyst plays a crucial role in the Hg0 heterogeneous oxidation process to form HgO. Its consumption can be replenished by gas phase O2 and Hg0 oxidation by surface oxygen is regarded as a Mars–Maessen process [43]. The Mars–Maessen mechanism can be illustrated by Equations (3)–(7) [24].
A ( g ) A ( a d s ) ,
A ( a d s ) + M x O y A O ( a d s ) + M x O y 1 ,
M x O y 1 + 1 2 O 2 M x O y ,
A O ( a d s ) A O ( g ) ,
A O ( a d s ) + M x O y A M x O y + 1 .
The Hg atom is initially adsorbed on the surface and then oxidized by the surface oxygen to form HgO. Based on the above calculations, the configurations of intermediate and final state can be determined according to the adsorption behaviors of Hg species. Furthermore, three possible reaction pathways are obtained by transition state search and frequency calculation verification. The proposed reaction pathways are given in Figure 6 with relative energy barriers. The optimized geometric configurations of intermediate (IM), transition state (TS) and final state (FS) in each reaction pathway are depicted in Figure 7.
In the case of pathway I, shown in Figure 6a, during the first stage, Hg0 and surface O atoms are firstl adsorbed on O-top site of the CeO2 (111) surface, forming IM1. The adsorption of Hg atom is a physisorption process and a barrierless exothermic reaction with exothermicity of −88.45 kJ/mol. In IM1, oxygen is adsorbed on the surface O-top atom with a 2.270 Å O–Otop bond, while the Hg atom is adsorbed in a physical manner, as mentioned above, and the length of the Hg–Otop bond is 2.339 Å. The distance between Hg0 and surface oxygen is 2.109 Å, which is larger than the Hg–O bond length of the gas-phase HgO molecule [43]. Then Hg0 moves toward the surface O atom and is oxidized by overcoming an energy barrier of 20.7 kJ/mol through transition state TS1. During this process, the distance between the surface O atom and Hg0 decreases gradually, i.e., 2.109 Å (IM1) → 2.019 Å (TS1) → 1.944 Å (FS). In TS1, Hg and O atoms move closer to each other, leading to the formation of an Hg–O bond with a length of 2.019 Å. The O atom migrates upward from the surface together with the Hg atom, and the distance of the Hg atom from the surface is 3.689 Å. In FS, Hg0 is oxidized to HgO with its Hg atom close to the CeO2 (111) surface, and is adsorbed in a vertical manner. The distance between Hg atoms and the surface is 2.079 Å, and the bond length of Hg–O further shortens to 1.944 Å. In this pathway, the reaction of Hg0 oxidation is an exothermic process and the reaction heat is −62.5 kJ/mol.
Figure 6b presents the energy profile of pathway II. Similar to pathway I, Hg0 is first adsorbed on the CeO2 (111) surface through a barrierless physical adsorption process. In this step, the reaction heat was −9.99 kJ/mol. IM2 formed after Hg0 was dissociatively adsorbed on CeO2 (111) surface. In IM2 configuration, adsorbed oxygen on the Otop site of CeO2 (111) surface forms a 1.409 Å O–O bond. The distance between Hg and surface O atom is 3.479 Å, which is much larger than the bond length of the gas-phase HgO molecule. Consequently, an oxidation reaction between the adsorbed Hg atom and the surface oxygen occurs in the IM2 configuration, during which the Hg atom approaches surface oxygen and HgO molecule forms on the surface O-site via IM2 → TS2 → FS. In TS2, the surface O atom is stripped by the Hg atom from the CeO2 (111) surface, resulting in the breakage of the surface O–O bond. The Hg atom also moves closer to the surface, from 3.526 Å to 2.826 Å. The distance between the surface O atom and Hg0 is similarly shortened gradually, i.e., 3.479 Å (IM2) → 2.051 Å (TS2) → 1.944 Å (FS), indicating the formation of an Hg–O bond. Subsequently, the HgO molecule in the TS2 configuration turns over and the horizontal adsorption configuration with the HgO molecule’s O atom near the surface transforms into a vertical adsorption configuration with the Hg atom near the CeO2 (111) surface. The whole reaction pathway is endothermic by 44.6 kJ/mol, with an energy barrier of 116.4 kJ/mol.
In addition, after the exothermic physisorption step with an exothermicity of −9.99 kJ/mol on the CeO2 (111) surface, IM2 configuration can be oxidized to FS’ via another potential pathway, which is noted as pathway III. The energy profile of reaction pathway III is presented in Figure 6c. In TS3, the surface oxygen and adsorbed Hg move toward each other, shortening the distance from 3.479 Å to 2.111 Å, which leads to the formation of the Hg–O bond. The distances between the Hg and O atoms and the CeO2 (111) surface are 2.332 Å and 2.275 Å, respectively. Finally, the FS’ is formed, in which HgO molecular is adsorbed in a parallel fashion on the surface. The length of the Hg–O bond is stabilized at 2.114 Å, while the distances between the Hg and O atoms and CeO2 (111) surface exhibit a tiny change and are 2.336 Å and 2.272 Å, respectively. This process is endothermic by 107.0 kJ/mol, with an energy barrier of 107.1 kJ/mol. It is noteworthy that since the coordinates of the saddle point on the potential energy surface are very close to the coordinates of the final state, the geometry configuration of TS3 is analogous to that of FS’, which makes the energy barrier of pathway III close to reaction heat.
In the above three pathways, since Hg0 adsorption on the surface is a barrierless process, the formation of HgO is the rate-determining step of the entire oxidation reaction. In addition, pathway I not only has the lowest energy barrier but also is an exothermic process. Hence, the Hg0 oxidation by oxygen on CeO2 (111) surface tends to occur through pathway I. In regard to other mercury heterogeneous oxidation mechanisms except the Mars–Maessen mechanism, Liu et al. studied the oxidation between Hg0 and HCl on the CeO2 (111) surface, identifying the rate-determining step as the formation of HgCl2 (Hg0 + HCl → HgCl2); this step followed the Langmuir–Hinshelwood mechanism [20]. Comparing the reaction pathways proposed in this work with the study by Liu et al. on the reaction between Hg0 and HCl, it is found that the energy barrier of pathway I, 20.7 kJ/mol, is lower than that of HgCl2 formation (56.92–78.05 kJ/mol). This suggests that the Mars–Maessen oxidation of Hg0 by oxygen on the CeO2 (111) surface can be of the same reactivity as the Langmuir–Hinshelwood oxidation of Hg0 by HCl.
Furthermore, the energy barrier of HgO forming on the CeO2 (111) surface is also compared with that on other materials, such as V2O5/TiO2 (001) and MnFe2O4 (100) [43,44]. In both systems, the oxidation of Hg0 by oxygen follows the Mars–Maessen mechanism. The barrierless adsorption of Hg0 is the first step and the formation of the HgO molecule is the rate-determining step, which is consistent with the results calculated in this study. It is found that CeO2 is superior at catalyzing the Hg0 oxidation by oxygen due to its lower energy barrier than the V2O5/TiO2 (001) system (143.9–151.2 kJ/mol) and the MnFe2O4 (100) surface (76.07–200.37 kJ/mol). Therefore, it can be concluded that, whether as a catalyst for Hg0 oxidation or as an adjuvant to modify the V–Ti-based catalyst, CeO2 exhibits excellent Hg0 oxidation performance, especially with insufficient HCl content in flue gas when burning low-rank coal. Based on the above results, it can be speculated that CeO2 can efficiently catalyze the oxidation of Hg0 by oxygen in flue gas and can be a promising catalytic material for flue gas mercury removal.

4. Conclusions

Mercury adsorption and oxidation mechanisms with oxygen on the CeO2 (111) surface were investigated using DFT calculations for the first time in this study. The adsorption manners of Hg species were studied first. Hg0 adsorbs on the CeO2 (111) surface in a physical adsorption manner with a −1~−5 kJ/mol adsorption energy, while HgO molecular is chemisorbed on the surface with its adsorption energy in the range of −69.90~−198.37 kJ/mol. Density of state analysis for Hg and Ce atoms indicates that Hg0 significantly interacts with the surface Ce atom and the adsorption configuration is stable. DFT calculations indicate that the reaction between Hg0 and surface oxygen on CeO2 (111) follows the Mars–Maessen mechanism. In the first step Hg0 is adsorbed on the surface through an exothermic process with no energy barriers; during the second stage, the adsorbed Hg atom is oxidized by the surface oxygen, which is subsequently replenished by the gas phase O2. Three possible reaction pathways are proposed for the abovementioned oxidation mechanism; among them the pathway I is exothermic and has the lowest energy barrier, while the others have relatively high energy barriers and are endothermic, which suggests that pathway I is the most favorable oxidation route. The rate-determining steps are the formation of the HgO molecule in each pathway.
To better evaluate the oxidation ability of the CeO2 + O2 system, energy barriers are compared with relative systems including CeO2 + HCl, V2O5 + O2, and MnFe2O4 + O2. The comparative analysis shows that the energy barrier of the oxidation reaction between Hg0 and oxygen on CeO2 (111) surface is lower than that of the reaction of Hg0 with HCl, and is also lower than those of other materials, which helps to prove that CeO2 is an active catalytic material with high Hg0 oxidation ability. In conclusion, the CeO2-based catalyst, which can enhance the oxidation rate of Hg0 in flue gas, may have promising application prospects for mercury removal, especially in low-rank coal-fired flue gas with a low HCl concentration.

Acknowledgments

The authors thank the National Basic Research Program of China (2015CB251501), the National Natural Science Foundation of China (51576064), the Beijing Nova Program (Z171100001117064), the Beijing Natural Science Foundation (3172030), the Fok Ying Tung Education Foundation (161051), and the Fundamental Research Funds for the Central Universities (2018ZD08, 2016YQ05) for financial support.

Author Contributions

Li Zhao designed the project and drafted the manuscript; Yangwen Wu and Jian Han performed the calculation and analyzed the data; Qiang Lu, Yongping Yang, and Laibao Zhang contributed to the analysis of data and discussion; all authors contributed to the revision of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, S.X.; Zhang, L.; Wang, L.; Wu, Q.R.; Wang, F.Y.; Hao, J.M. A review of atmospheric mercury emissions, pollution and control in China. Front. Environ. Sci. Eng. 2014, 8, 631–649. [Google Scholar] [CrossRef]
  2. Liu, R.H.; Xu, W.Q.; Tong, L.; Zhu, T.Y. Mechanism of Hg0 oxidation in the presence of HCl over a commercial V2O5-WO3/TiO2 SCR catalyst. J. Environ. Sci.-China 2015, 36, 76–83. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, W.H.; Lai, C.W.; Sharifah, B.A.H. In Situ Anodization of WO3-Decorated TiO2 Nanotube Arrays for Efficient Mercury Removal. Materials 2015, 8, 5702–5714. [Google Scholar] [CrossRef] [PubMed]
  4. Cao, R.S.; Fan, M.Y.; Hu, J.W.; Ruan, W.Q.; Xiong, K.N.; Wei, X.H. Optimizing Low-Concentration Mercury Removal from Aqueous Solutions by Reduced Graphene Oxide-Supported Fe3O4 Composites with the Aid of an Artificial Neural Network and Genetic Algorithm. Materials 2017, 10, 1279. [Google Scholar] [CrossRef] [PubMed]
  5. Streets, D.G.; Lu, Z.F.; Levin, L.; Ter Schure, A.; Sunderland, E.M. Historical releases of mercury to air, land, and water from coal combustion. Sci. Total Environ. 2017, 615, 131–140. [Google Scholar] [CrossRef] [PubMed]
  6. Ji, W.C.; Shen, Z.M.; Tang, Q.L.; Yang, B.W.; Fan, M.H. A DFT study of Hg0, adsorption on Co3O4 (110) surface. Chem. Eng. J. 2016, 289, 349–355. [Google Scholar] [CrossRef]
  7. Wu, Y.; Wang, S.X.; Streets, D.G.; Hao, J.M.; Chan, M.; Jiang, J.K. Trends in Anthropogenic Mercury Emissions in China from 1995 to 2003. Environ. Sci. Technol. 2006, 40, 5312–5318. [Google Scholar] [CrossRef] [PubMed]
  8. Li, H.L.; Zhu, L.; Wu, S.K.; Liu, Y.; Shih, K.M. Synergy of CuO and CeO2 combination for mercury oxidation under low-temperature selective catalytic reduction atmosphere. Int. J. Coal Geol. 2017, 170, 69–76. [Google Scholar] [CrossRef]
  9. Zhao, L.K.; Li, C.T.; Zhang, X.N.; Zeng, G.M.; Zhang, J.; Xie, Y.E. A review on oxidation of element mercury from coal-fired flue gas with selective catalytic reduction catalysts. Catal. Sci. Technol. 2015, 5, 3459–3472. [Google Scholar] [CrossRef]
  10. Zhou, Z.J.; Liu, X.W.; Zhao, B.; Chen, Z.G.; Shao, H.Z.; Wang, L.L.; Xu, M.H. Effects of existing energy saving and air pollution control devices on mercury removal in coal-fired power plants. Fuel Process. Technol. 2015, 131, 99–108. [Google Scholar] [CrossRef]
  11. Huang, W.J.; Xu, H.M.; Qu, Z.; Zhao, S.J.; Chen, W.M.; Yan, N.Q. Significance of Fe2O3, modified SCR catalyst for gas-phase elemental mercury oxidation in coal-fired flue gas. Fuel Process. Technol. 2016, 149, 23–28. [Google Scholar] [CrossRef]
  12. Zhao, B.; Liu, X.W.; Zhou, Z.J.; Shao, H.Z.; Xu, M.H. Catalytic oxidation of elemental mercury by Mn–Mo/CNT at low temperature. Chem. Eng. J. 2016, 284, 1233–1241. [Google Scholar] [CrossRef]
  13. Jung, J.E.; Geatches, D.; Lee, K.; Aboud, S.; Brown, G.E.; Wilcox, J. First-Principles Investigation of Mercury Adsorption on the α-Fe2O3 (1102) Surface. J. Phys. Chem. C 2015, 119, 26512–26518. [Google Scholar] [CrossRef]
  14. Zhao, L.K.; Li, C.T.; Zhang, X.N.; Zeng, G.M.; Zhang, J.; Xie, Y.E. Oxidation of elemental mercury by modified spent TiO2-based SCR-DeNOx catalysts in simulated coal-fired flue gas. Environ. Sci. Pollut. Res. 2016, 23, 1471–1481. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, X.N.; Li, C.T.; Zhao, L.K.; Zhang, J.; Zeng, G.M.; Xie, Y.E.; Yu, M.E. Simultaneous removal of elemental mercury and NO from flue gas by V2O5–CeO2/TiO2 catalysts. Appl. Surf. Sci. 2015, 347, 392–400. [Google Scholar] [CrossRef]
  16. Li, H.L.; 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] [PubMed]
  17. Phuong, T.M.P.; Thang, L.M.; Tien, T.N.; Isabel, V.D. CeO2 Based Catalysts for the Treatment of Propylene in Motorcycle’s Exhaust Gases. Materials 2014, 7, 7379–7397. [Google Scholar] [CrossRef]
  18. Fan, X.P.; Li, C.T.; Zeng, G.M.; Zhang, X.; Tao, S.S.; Lu, P.; Tan, Y.; Luo, D.Q. Hg0 Removal from Simulated Flue Gas over CeO2/HZSM-5. Energy Fuels 2012, 26, 2082–2089. [Google Scholar] [CrossRef]
  19. Zhao, L.; He, Q.S.; Li, L.; Lu, Q.; Dong, C.Q.; Yang, Y.P. Research on the catalytic oxidation of Hg0 by modified SCR catalysts. J. Fuel Chem. Technol. 2015, 43, 628–634. [Google Scholar] [CrossRef]
  20. Zhang, B.K.; Liu, J.; Shen, F.H. Heterogeneous Mercury Oxidation by HCl over CeO2 Catalyst: Density Functional Theory Study. J. Phys. Chem. C 2015, 119, 15047–15055. [Google Scholar] [CrossRef]
  21. Wang, S.X.; Zhang, L.; Li, G.H.; Wu, Y.; Hao, J.M.; Pirrone, N.; Sproviery, F.; Ancora, M.P. Mercury emission and speciation of coal-fired power plants in China. Atmos. Chem. Phys. 2010, 10, 1183–1192. [Google Scholar] [CrossRef]
  22. He, C.; Shen, B.X.; Chi, G.L.; Li, F.K. Elemental mercury removal by CeO2/TiO2-PILCs under simulated coal-fired flue gas. Chem. Eng. J. 2016, 300, 1–8. [Google Scholar] [CrossRef]
  23. Zhao, L.L.; Li, C.T.; Zhang, J.; Zhang, X.N.; Zhan, F.M.; Ma, J.F.; Xie, Y.E.; Zeng, G.M. 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]
  24. Presto, A.A.; Granite, E.J. Survey of catalysts for oxidation of mercury in flue gas. Environ. Sci. Technol. 2006, 40, 5601–5609. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, B.K.; Liu, J.; Zhang, J.Y.; Zheng, C.G.; Chang, M. Mercury oxidation mechanism on Pd (100) surface from first-principles calculations. Chem. Eng. J. 2014, 237, 344–351. [Google Scholar] [CrossRef]
  26. Zhang, B.K.; Liu, J.; Yang, Y.J.; Chang, M. Oxidation mechanism of elemental mercury by HCl over MnO2 catalyst: Insights from first principles. Chem. Eng. J. 2015, 280, 354–362. [Google Scholar] [CrossRef]
  27. Xu, Z.M.; Lv, X.J.; Chen, J.A.; Jiang, L.X.; Lai, Y.Q.; Li, J. First principles study of adsorption and oxidation mechanism of elemental mercury by HCl over MoS2 (100) surface. Chem. Eng. J. 2016, 308, 1225–1232. [Google Scholar] [CrossRef]
  28. Yang, Y.J.; Liu, J.; Zhang, B.K.; Liu, F. Density functional theory study on the heterogeneous reaction between Hg0, and HCl over spinel-type MnFe2O4. Chem. Eng. J. 2017, 308, 897–903. [Google Scholar] [CrossRef]
  29. Delley, B. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  30. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rec. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  31. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef]
  32. Delley, B. Hardness conserving semilocal pseudopotentials. Phys. Rev. B 2002, 66, 155125. [Google Scholar] [CrossRef]
  33. Song, Y.L.; Yin, L.L.; Zhang, J.; Hu, X.P.; Gong, X.Q. A DFT+U study of CO oxidation at CeO2 (110) and (111) surfaces with oxygen vacancies. Surf. Sci. 2013, 618, 140–147. [Google Scholar] [CrossRef]
  34. Huang, M.; Fabris, S. CO adsorption and oxidation on ceria surfaces from DFT+ U calculations. J. Phys. Chem. C 2008, 112, 8643–8648. [Google Scholar] [CrossRef]
  35. Ganduglia-Pirovano, M.V.; Da, S.J.; Sauer, J. Density-functional calculations of the structure of near-surface oxygen vacancies and electron localization on CeO2 (111). Phys. Rev. Lett. 2009, 102, 026101. [Google Scholar] [CrossRef] [PubMed]
  36. Fronzi, M.; Soon, A.; Delley, B.; Traversa, E.; Stampfl, C. Stability and morphology of cerium oxide surfaces in an oxidizing environment: A first-principles investigation. J. Chem. Phys. 2009, 131, 104701. [Google Scholar] [CrossRef]
  37. Fronzi, M.; Piccinin, S.; Delley, B.; Traversa, E.; Stampl, C. Water adsorption on the stoichiometric and reduced CeO2 (111) surface: A first-principles investigation. Phys. Chem. Chem. Phys. 2009, 11, 9188–9199. [Google Scholar] [CrossRef] [PubMed]
  38. Jiang, Y.; Adams, J.B.; Schilfgaarde, M. Density-functional calculation of CeO2 surfaces and prediction of effects of oxygen partial pressure and temperature on stabilities. J. Chem. Phys. 2005, 123, 064701. [Google Scholar] [CrossRef] [PubMed]
  39. Cheng, Z.; Sherman, B.J.; Lo, C.S. Carbon dioxide activation and dissociation on ceria (110): A density functional theory study. J. Chem. Phys. 2013, 138, 014702. [Google Scholar] [CrossRef] [PubMed]
  40. Kumari, N.; Haider, M.A.; Agarwal, M.; Shina, N.; Basu, S. Role of Reduced CeO2 (110) Surface for CO2 Reduction to CO and Methanol. J. Phys. Chem. C 2016, 120, 16626–16635. [Google Scholar] [CrossRef]
  41. Gerward, L.; Olsen, J.S.; Vaitheeswaran, G.; Kanchana, V.; Svane, A. Bulk modulus of CeO2 and PrO2—An experimental and theoretical study. J. Alloys Compd. 2005, 400, 56–61. [Google Scholar] [CrossRef]
  42. Halgren, T.A.; Lipscomb, W.N. The synchronous-transit method for determining reaction pathways and locating molecular transition states. Chem. Phys. Lett. 1977, 49, 225–232. [Google Scholar] [CrossRef]
  43. Yang, Y.J.; Liu, J.; Zhang, B.K.; Liu, F. Mechanistic studies of mercury adsorption and oxidation by oxygen over spinel-type MnFe2O4. J. Hazard. Mater. 2017, 321, 154–161. [Google Scholar] [CrossRef] [PubMed]
  44. Gao, Y.Y.; Li, Z.X. A DFT study of the Hg0 oxidation mechanism on the V2O5-TiO2 (001) surface. Mol. Catal. 2017, 433, 372–382. [Google Scholar] [CrossRef]
Figure 1. CeO2 unit cell: red ball = O atom; white ball = Ce atom.
Figure 1. CeO2 unit cell: red ball = O atom; white ball = Ce atom.
Materials 11 00485 g001
Figure 2. CeO2 (111) surface model: (a) Side view; (b) top view.
Figure 2. CeO2 (111) surface model: (a) Side view; (b) top view.
Materials 11 00485 g002
Figure 3. Adsorption configurations of Hg0 on the CeO2 (111) surface.
Figure 3. Adsorption configurations of Hg0 on the CeO2 (111) surface.
Materials 11 00485 g003
Figure 4. PDOS of Hg0 adsorption on the CeO2 (111) surface before and after adsorption.
Figure 4. PDOS of Hg0 adsorption on the CeO2 (111) surface before and after adsorption.
Materials 11 00485 g004
Figure 5. HgO adsorption configurations on the CeO2 (111) surface.
Figure 5. HgO adsorption configurations on the CeO2 (111) surface.
Materials 11 00485 g005
Figure 6. Reaction pathways and energy profiles of oxidation between Hg0 and O2 over the CeO2 (111) surface: (a) pathway I; (b) pathway II; (c) pathway III.
Figure 6. Reaction pathways and energy profiles of oxidation between Hg0 and O2 over the CeO2 (111) surface: (a) pathway I; (b) pathway II; (c) pathway III.
Materials 11 00485 g006
Figure 7. Configurations of intermediates, transition states and final states in Hg0 oxidation pathways on the CeO2 (111) surface.
Figure 7. Configurations of intermediates, transition states and final states in Hg0 oxidation pathways on the CeO2 (111) surface.
Materials 11 00485 g007
Table 1. The adsorption energy, geometry parameters and Mulliken charge for Hg0 adsorption on CeO2 (111) surface.
Table 1. The adsorption energy, geometry parameters and Mulliken charge for Hg0 adsorption on CeO2 (111) surface.
ConfigurationsEads (kJ/mol)RX-Hg (Å) *QHg (e)
1A−5.713.7000.012
1B−1.543.4170.008
1C−5.303.616/3.6040.009
1D−5.254.3600.011
* X denotes surface atom on the CeO2 (111) surface.
Table 2. The adsorption energy, geometry parameters, and Mulliken charge for Hg0 adsorption on the CeO2 (111) surface.
Table 2. The adsorption energy, geometry parameters, and Mulliken charge for Hg0 adsorption on the CeO2 (111) surface.
ConfigurationsEads (kJ/mol)RX-Hg (Å) *RO-Hg (Å)RX-O (Å)
2A−69.90-2.1452.153
2B−101.992.1142.2772.400
2C−163.812.0771.945-
2D−198.37-3.3541.415
* X denotes surface atom on the CeO2 (111) surface.

Share and Cite

MDPI and ACS Style

Zhao, L.; Wu, Y.; Han, J.; Lu, Q.; Yang, Y.; Zhang, L. Mechanism of Mercury Adsorption and Oxidation by Oxygen over the CeO2 (111) Surface: A DFT Study. Materials 2018, 11, 485. https://doi.org/10.3390/ma11040485

AMA Style

Zhao L, Wu Y, Han J, Lu Q, Yang Y, Zhang L. Mechanism of Mercury Adsorption and Oxidation by Oxygen over the CeO2 (111) Surface: A DFT Study. Materials. 2018; 11(4):485. https://doi.org/10.3390/ma11040485

Chicago/Turabian Style

Zhao, Li, Yangwen Wu, Jian Han, Qiang Lu, Yongping Yang, and Laibao Zhang. 2018. "Mechanism of Mercury Adsorption and Oxidation by Oxygen over the CeO2 (111) Surface: A DFT Study" Materials 11, no. 4: 485. https://doi.org/10.3390/ma11040485

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