DOI:
10.1039/C6RA20457K
(Paper)
RSC Adv., 2016,
6, 104294-104302
Removal of gas-phase Hg0 by Mn/montmorillonite K 10
Received
13th August 2016
, Accepted 26th October 2016
First published on 27th October 2016
Abstract
Mn/montmorillonite K 10 (Mn/MK10) prepared by impregnation method was studied to remove Hg0 in simulated coal-fired flue gas. The samples were characterized by Brunauer–Emmett–Teller (BET), X-ray diffraction (XRD), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS) and temperature-programmed desorption (TPD). Effects of manganese loading value, reaction temperature and flue gas components on Hg0 removal efficiency were investigated. The results indicated that 4% Mn/MK10 was the optimal sample with outstanding Hg0 removal efficiency over the temperature range of 100–400 °C. The characteristic analysis demonstrated that amorphous MnO2 and active oxygen were crucial for Hg0 removal. Besides, NO had a promoting effect due to the formation of Hg(NO3)2. The addition of only 5 ppm HCl led to excellent Hg0 removal performance as HCl enhanced Hg0 conversion to HgClx. The inhibition effect of SO2 could be counteracted in the presence of NO and/or HCl. The Hg0 removal capacity showed a relative decrease when H2O (g) was added to simulated flue gas. Moreover, Hg0 removal performance was maintained at 80–99% (NO/SO2 = 0.26–1.71) in simulated flue gas without HCl, which appeared to be promising in industrial application.
1. Introduction
Mercury emissions have become a major environmental issue that attracts considerable attention owing to its extreme toxicity, volatility, persistence, and bioaccumulation.1 As coal is a predominant energy source in China, coal combustion is a primary source of emitted mercury. There are three major forms of mercury emitted during coal combustion: oxidized mercury (Hg2+), particle-bound mercury (HgP) and elemental mercury (Hg0).2 Hg2+ is water-soluble, enabling its removal through wet flue gas desulfurization devices (WFGD). Hgp can be easily captured by electrostatic precipitators or fabric filters. However, Hg0 is much harder to remove due to its low water solubility and high vapor pressure.3 Thus, the control of Hg0 in coal-fired flue gas is particularly difficult and challenging.4
Numerous technologies have been developed for removing Hg0, and one of the most effective is activated carbon injection. However, some shortcomings limit its development, such as high cost, poor capacity, low temperature, high carbon-to-mercury ratios, and slow regeneration.5,6 Transition metal oxides, proposed as one of the most promising alternatives to activated carbon sorbents, have been extensively examined in recent years.7–17 In particular, manganese oxide (MnOx) has been shown to have a good mercury capture potential through effective catalytic oxidation.8,9 Manganese, which has multiple oxidation states, offers the ability for Hg0 oxidation to be achieved. According to the literature, Mn/MgO is effective for Hg0 removal at low temperatures, with amorphous MnO2 and chemisorbed O2 playing a crucial role in Hg0 oxidation.10 Xu et al.11 found that MnOx/graphene exhibits excellent Hg0 removal efficiency at 150 °C. Not only is MnOx considered the main active site, but also it is often an additive for producing a synergistic effect for Hg0 removal. In Li's study,12 Mn–Ce/Ti catalysts were highly active for Hg0 removal because of the coexistence of Mn4+ and Ce3+. Consequently, most previous studies have shown the positive effects of manganese for Hg0 removal.
Materials, like TiO2 and Al2O3, were often employed as effective carries for Hg0 removal.13–17 However, previous studies revealed some significant drawbacks for Hg0 removal over these catalysts, such as the inefficiency at high temperature,12,13 the low resistance to SO2 poisoning14,15 and the dependence on high HCl concentrations.16 In recent years, scientists have shown much interest in minerals, such as magnetite,18 zeolite,19,20 and clay,21–23 principally to take advantage of their natural, abundant, and unique structural features. Montmorillonite, a type of natural Si–Al-based mineral, is widely used in adsorption and catalysis fields such as the removal of VOCs,24,25 CO2
26,27 and SO2.28,29 Up to now, however, it has hardly been used in gas-phase Hg0 removal.
As an effective adsorbent and catalyst, the montmorillonite owns several advantages in Hg0 removal. Firstly, its layer-structure and abundance of porosity are good for Hg0 adsorption.30 Secondly, the good thermal stability of montmorillonite makes it possible to remove Hg0 at high temperature.31 Thirdly, its high surface acidity benefits the catalytic activity, especially for heavy metal removal.32 Finally, its cost is only 1/10–1/20 of that of activated carbon.33 Therefore, it is interesting and promising to combine the advantages of Mn and montmorillonite for synthesizing a novel sample to effectively remove Hg0.
In this work, we synthesized a novel sample based on manganese oxide impregnated on montmorillonite K 10 (MK10: montmorillonite prepared by acid activation), and the Hg0 removal performance in simulated flue gas was investigated over the temperature range of 100–400 °C. Brunauer–Emmett–Teller (BET), X-ray diffraction (XRD), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), and temperature-programmed desorption (TPD) analysis were used to characterize the samples. The effects of flue gas components on Hg0 removal were examined, especially SO2 resistance and the effect of low/zero HCl concentration. Further, a reaction mechanism of Hg0 removal over the prepared samples was proposed. This work also provides information on the industrial application of Mn/MK10 samples.
2. Experimental
2.1. Preparation of samples
The Mn/MK10 samples were prepared by an impregnation method. In a typical procedure, requisite quantities of MK10 and Mn(NO3)2 precursors were mixed in deionized water and stirred for 4 h at 25 °C. Afterwards, a rotary evaporator at 55 °C was used to evaporate excess water from the mixtures, and the obtained solid was dried at 110 °C and calcined at 500 °C for 4 h. The derived samples were crushed and sieved to 60–80 mesh (180–250 μm) particles for the sample activity tests. The samples are denoted as x% Mn/MK10, where x represents the weight ratio of Mn to MK10.
2.2. Characterization of samples
Nitrogen adsorption–desorption isotherms were recorded with a nitrogen adsorption apparatus (Autosorb-iQ, Quantachrome, USA). The specific surface areas and pore volumes were calculated using the BET and Barrett–Joyner–Halenda (BJH) methods, respectively. XRD patterns were determined using a Rigaku D/Max-RA powder diffractometer with a Ni-filtered Cu Kα radiation source (40 kV and 150 mA) and a scanning range from 4° to 70°. The physical morphology of the samples was observed using SEM (JSM 6700F, JEOF) with a field emission gun. XPS was carried out on an ESCALAB 250Xi employing Al Kα radiation (hν = 1486.6 eV) as the excitation source. The binding energies were corrected using the binding energy of adventitious carbon (284.8 eV).
Before the Hg temperature-programmed desorption (Hg-TPD) tests, about 50 mg samples were first treated with a gas flow of Hg0 balanced in N2 for 3 h at 150 °C and then purged with N2 at a flow rate of 500 mL min−1 until the outlet of the Hg0 concentration decreasing to zero. The sample was heated from 100 °C to 700 °C at a heating rate of 10 °C min−1 in pure N2.
2.3. Sample activity test
The sample activity test was performed in a fixed-bed quartz tube reactor (4 mm i.d.) containing 0.05 g of sample. A gas comprising 120 μg m−3 Hg0 with a balance of N2 was generated using an Hg0 permeation tube. The simulated flue gas comprising 0–5% O2, 0–500 ppm NO, 0–1000 ppm SO2, 0–5 ppm HCl, 0–5% H2O and balanced N2 was introduced into the reactor at the steady state. All lines that Hg0 passed through were heated to 90 °C to prevent mercury deposition on the inner surface. The Hg0 concentration was continuously monitored using a Lumex RA915M Zeeman mercury analyzer. In all tests, the total gas flow rate was 500 mL min−1, the space velocity was 478
000 h−1, and the steady reaction time was more than 3 h. A mercury concentration mass balance was conducted before each test. The Hg0 removal efficiency (η) is defined as| |
 | (1) |
where Hg0inlet (μg m−3) and Hg0outlet (μg m−3) are the concentrations of Hg0 measured at the inlet and the outlet of the reactor, respectively.
3. Results and discussion
3.1. Activity test for Hg0 removal
A series of experiments were conducted to examine the effect of the Mn contents on Hg0 removal over virgin MK10 and Mn/MK10 at various temperatures. Fig. 1 reveals no apparent Hg0 removal efficiency is observed over pure MK10, while the efficiency of Hg0 removal is gradually enhanced to over 90% with increasing Mn loading and temperature. When the Mn content was below 4%, the Hg0 removal performance at low temperature (<200 °C) was relatively poor, but the efficiency significantly increased with increasing temperature. However, at high temperature (>200 °C), the performance was comparably high and stable. The 4% Mn/MK10 sample showed the highest Hg0 removal efficiency and stability across a wide temperature window from 100 to 400 °C. Unlike other Mn-loaded materials, the outstanding performance of 4% Mn/MK10 at high temperatures is partly a result of the thermal and chemical stability of MK10.32 Further loading with Mn did not obviously heighten the sample efficiency. Thus, all subsequent tests were conducted utilizing the 4% Mn/MK10 samples.
 |
| | Fig. 1 Hg0 removal efficiency over various Mn/MK10 samples at different temperatures (reaction conditions: O2 = 5%, balance N2). | |
3.2. Characterization of samples
3.2.1. BET analysis. The textural properties, including specific surface areas and pore volumes, of virgin MK10 and the impregnated MK10 (x% Mn/MK10) samples are summarized in Table 1. Compared with virgin MK10, the MK10 samples impregnated with MnOx possess larger BET surface areas and pore volumes, similar to results found in other literature.34,35 For one thing, as shown in Fig. 2, the impregnation and calcination process generated some new pores and increased the number of small pores (pore size < 4 nm) due to the powerful oxidizing property of Mn. For another, MnOx was relatively well dispersed on the surface of the samples.35 With an increase of the Mn content from 1% to 8%, the BET surface area first increased from 69.90 to 88.39 m2 g−1 and then decreased to 81.16 m2 g−1. This tendency was likely due to the formation and aggregation of MnOx crystal particles on the sample surface, which caused pore blockages at high Mn loadings. Moreover, this phenomenon partly accounts for the trend observed in Fig. 1, in which the Hg0 removal efficiency did not increase further at high Mn loadings.
Table 1 Textural properties of various Mn/MK10 samples
| Sample |
SBET (m2 g−1) |
Vtotal (cm3 g−1) |
| Virgin MK10 |
64.16 |
0.108 |
| 1% Mn/MK10 |
69.93 |
0.115 |
| 2% Mn/MK10 |
76.59 |
0.122 |
| 3% Mn/MK10 |
80.47 |
0.130 |
| 4% Mn/MK10 |
88.39 |
0.141 |
| 6% Mn/MK10 |
86.40 |
0.139 |
| 8% Mn/MK10 |
81.16 |
0.132 |
 |
| | Fig. 2 Pore size distribution of MK10 and 4% Mn/MK10. | |
3.2.2. XRD analysis. To identify the structure of Mn species on the sample surface, XRD measurements were obtained for virgin MK10 and various Mn-loaded MK10 samples. The XRD patterns in Fig. 3 show characteristics peaks of MK10 at 2θ values of 5.9°, 19.8°, and 62.3° for all of the samples. However, the intensity of these peaks gradually weakened with increasing MnOx content, indicating that MnOx strongly interacted with MK10 in these samples. No obvious characteristic peaks of MnOx were observed when Mn loading was below 6%, which indicated that MnOx might be highly dispersed or present as an amorphous phase on sample surface, which can promote the catalytic oxidation activity efficiently.36 However, on further increasing the Mn content, peaks were observed at the 2θ values of 28.7°, 37.4°, and 56.8°, corresponding to MnO2. This result confirmed the BET results, in which too high a loading led to the formation and aggregation of crystalline phases. In addition, no Mn2O3 or MnO characteristic peaks were detected in any of these samples, signifying that no or little Mn3+ and Mn2+ were generated during the impregnation process.
 |
| | Fig. 3 XRD pattern of various Mn/MK1O samples. | |
3.2.3. SEM and EDS analysis. SEM was used to analyze the alteration of the sample prior to and after impregnation. As shown in Fig. 4a (MK10) and 4c (4% Mn/MK10), the differences in the shape and size of the whole structure under low magnification (×2000) were very minute, illustrating that the addition of 4% Mn during the synthetic process did not substantially change the particle size. Moreover, the increase in the amount of small pores, as shown in Fig. 2, could not be observed under low magnification. Notably, under high magnification (×40
000), the “flower-like” smooth layer structure clearly observed in Fig. 4b (MK10) was destroyed after Mn impregnation, as seen in Fig. 4d (4% Mn/MK10), which could indicate a reaction between MK10 and the active components as well as MnOx is likely highly dispersed on the surface of 4% Mn/MK10, consistent with the XRD results. Similarly, the EDS spectrum shown in Fig. 4e confirmed that the pure MK10 mainly contained elements of Si, Al, O. And Fig. 4f clearly shows the existence of Mn on the 4% Mn/MK10, which demonstrates the successful impregnating manganese oxide into MK10.
 |
| | Fig. 4 SEM images (low and high magnification) and EDS patterns of MK10 (a, b, e) and 4% Mn/MK10 (c, d, f). | |
3.2.4. XPS analysis. To determine the chemical state and the relative proportion of the main elements on the surface of the samples, XPS was used to investigate the 4% Mn/MK10 sample before and after Hg0 adsorption under pure N2, as seen in Fig. 5 and Table 2.
 |
| | Fig. 5 O1s and Mn2p XPS spectra of fresh and spent 4% Mn/MK10 samples. | |
Table 2 The XPS data of O1s and Mn2p species in 4% Mn/MK10 samples
| Sample |
Oα (%) |
Oβ (%) |
Oγ (%) |
Mn4+ (%) |
Mn3+ (%) |
| Fresh 4% Mn/MK10 |
42.24 |
55.20 |
2.56 |
89.97 |
10.03 |
| Spent 4% Mn/MK10 |
41.36 |
44.07 |
14.57 |
58.57 |
41.43 |
The O1s XPS spectra revealed three types of oxygen on the 4% Mn/MK10 sample, with characteristic peaks at 530.1–530.3 eV, 531.1–531.3 eV, and 532.3–532.7 eV ascribed to lattice oxygen (denoted as Oα), chemisorbed oxygen/OH groups (denoted as Oβ), and molecular water (denoted as Oγ), respectively.37,38 Clearly, Oα and Oβ were the primary oxygen species in the fresh sample. However, both of these species were diminished after Hg0 absorption, while the amouqf Oγ increased from 2.56% to 14.57%, which suggested that both lattice oxygen and chemisorbed oxygen or OH groups participated in the Hg0 oxidation process. The reduction of Oβ was much more obvious than that of Oα (Table 2); chemisorbed oxygen has a highly active and important role in oxidation reactions because it has higher mobility than lattice oxygen.39 Hence, the consumption of more chemisorbed oxygen than lattice oxygen after Hg0 adsorption was reasonable.
Similar to other Mn-containing samples,38,40 the Mn2p region of the samples in this work included Mn2p3/2 with a binding energy of 641–644 eV and Mn2p1/2 with a binding energy of 654.4 eV. The Mn2p3/2 peak could be further fitted into two peaks around 643.0 eV and 641.7 eV, corresponding to Mn4+ and Mn3+, respectively. For the fresh sample, Mn4+ was the dominant Mn valence state, accounting for 89.97% (Table 2), which demonstrated that MnO2 is the main active species, consistent with the XRD results. However, the amount of Mn3+ increased from 10.03% to 41.43% after Hg0 adsorption, indicating that Mn4+ took part in Hg0 oxidation and was converted to Mn3+.41
3.3. Effect of individual gas components
To better understand Hg0 removal capacity over 4% Mn/MK10, it is necessary to explore the roles of individual flue gas components on Hg0 removal efficiencies, especially NO, HCl and SO2. Experiments were conducted by mixing Hg0 with individual flue gas components, with or without O2, and the results are shown in Fig. 6. To explain the proposed reaction mechanism of each flue gas components, the TPD tests were carried out and the results are shown in Fig. 7.42,43
 |
| | Fig. 6 Hg0 removal efficiency of 4% Mn/MK10 over individual gas components at 150 °C (reaction conditions: O2 = 0/5%, NO = 500 ppm, HCl = 5 ppm, SO2 = 1000 ppm, balance). | |
 |
| | Fig. 7 TPD profiles of 4% Mn/MK10 under N2 with various gas components. | |
3.3.1. Effect of O2. As seen in Fig. 6, the Hg0 removal efficiency on 4% Mn/MK10 was 36.3% under pure N2 atmosphere, and the TPD profiles in Fig. 7a indicate the primary Hg0 desorption peaks appeared at approximately 210 °C was attributed to HgO.44,45 Therefore, the loss of mercury could be explained by Mars–Maessen mechanism, where Hg0 bonded with lattice oxygen to form weakly bonded Hg–O–Mn–Ox−1 or reacted with surface oxygen (including chemisorbed oxygen and lattice oxygen) to form HgO directly. Apparently, surface oxygen played a key role in Hg0 conversion under pure N2 flue gas. When 5% O2 was added to the gas flow, the performance rapidly improved to 90.9%. Serving as the Hg0 oxidant, O2 could regenerate the lattice oxygen and replenish the consumed chemisorbed oxygen, and hence improve Hg0 oxidation.
3.3.2. Effect of NO. To determine the effect of NO on Hg0 removal over Mn/MK10, 500 ppm NO was added into pure N2 gas flow, the change of Hg0 removal efficiency is shown in Fig. 6. A prominent promotional effect was observed with the removal efficiency rapidly increased to 96.6%. This might partly be due to the addition of MnO2 in the sample which can adsorb and oxidize NO.46 Moreover, as shown in Fig. 7b, the TPD peaks appearing at around 213 °C and 435 °C indicated the production of Hg(NO3)2.42 It has been reported that,47 NO would be weakly adsorbed on the surface of the metal oxide samples in the absence of O2, and a fraction of it would react with the surface oxygen to generate NO2, which are more active for Hg0 oxidation. Adding 5% O2 resulted in 99.6% Hg0 removal efficiency, demonstrating that gas-phase O2 regenerates and replenishes surface oxygen to facilitate the conversion of NO into NO2 and Hg(NO3)2.
3.3.3. Effect of HCl. Chlorine, the major halogen specie in coals, is considered to be present chiefly as HCl in coal derived flus gases. As the main oxidized mercury specie in coal combustion flue gas is HgCl2, HCl is regarded as the most important flue gas component affecting Hg0 removal. Therefore, the effect of HCl on Hg0 removal was investigated. In accordance with the low concentration of HCl in actual flue gas in China, only 5 ppm HCl was added to pure N2 gas flow.48 As shown in Fig. 6, the Hg0 removal efficiency increased to 98.9%, which is much higher than that observed under N2 atmosphere. Lattice oxygen and chemisorbed oxygen supported the transformation of HCl to either Cl2 or other chlorine species,49,50 which are responsible for Hg0 oxidation in the presence of HCl. Additionally, the TPD results in Fig. 7c showed two desorption peaks at 220 °C and 370 °C. Because the Hg0 removal performance under HCl atmosphere is evidently superior to that of the N2 atmosphere, these two peaks are most likely related to the decomposition of HgClx species.51 This is consistence with the result of other literatures that HCl could enhance Hg0 conversion to HgCl2 over manganese based catalysts.52
3.3.4. Effect of SO2. Promotional,53 negligible54 or inhibitive38,55 effects of SO2 have all been reported. There was not consistent conclusion concerning the effect of SO2 on Hg0 removal in flue gas over metal oxide samples. In this study, the results in Fig. 6 showed the efficiency decreased from 36.3% to 20.7% when 1000 ppm SO2 was added into pure N2, which revealed a suppressive effect of SO2 on Hg0 removal. Two possible reasons are responsible for this result. On the one hand, SO2 could compete with Hg0 for active sites on Mn/MK10. On the other hand, the adsorbed SO2 could react with the surface oxygen to form SO3 and thus consumed reactive oxygen which is beneficial for Hg0 removal.53 If the deactivation of SO2 was mainly attributed to the consumption of reactive oxygen, the addition of O2 would apparently relieve the inhibitive effect. However, the presence of 5% O2 only increased the removal efficiency from 20.7% to 29.1%, which still greatly inhibited Hg0 removal. As the TPD results shown in Fig. 7d, the slightly increased efficiency was probably due to the formation of HgxSO4 in the presence of O2.42 That is, the addition of O2 only resulted in slight remission, with the removal efficiency still significantly suppressed compared with the efficiency under pure N2. Therefore, it was not the consumption of surface oxygen but the competition of SO2 and Hg0, the dominant factor in the inhibition by SO2.
3.4. Effect of multiple gas components
3.4.1. Effect of SO2 in the presence of NO and/or HCl. As discussed above, SO2 had an inhibitive effect on Hg0 removal, regardless of the presence or absence of O2. Thus, it is important to investigate whether the suppressive effect of SO2 on Mn/MK10 sample can be offset in the presence of NO and/or HCl. As shown in Fig. 8, the efficiencies significantly increase from 29.1% to over 97% only if NO and/or HCl are added to the gas flow, which means the presence of NO and/or HCl can effectively counteract the inhibiting effect of SO2. This phenomenon was probably due to the different active sites between SO2 and NO/HCl, and Hg0 could be preferentially absorbed on NO/HCl. Thus, the promoting role of NO and/or HCl could offset the inhibitory effect of SO2 on Hg0 removal. In general, NO always exists in actual coal-fired flue gases, and there is often a low concentration of HCl in flue gases in China. Therefore, it was reasonable and acceptable to conclude that the SO2-poisoning phenomenon would be avoided for Mn/MK10 samples in simulated flue gases, even in the absence of HCl. Notably, under a typical-simulated flue gas including 5% O2, 1000 ppm SO2, 500 ppm NO, and 5 ppm HCl, the Hg0 removal performance of 4% Mn/MK10 was maintained at >97% for more than 140 h, which shows the excellent stability of 4% Mn/MK10 for Hg0 removal.
 |
| | Fig. 8 Hg0 removal efficiency of 4% Mn/MK10 over multiple gas components at 150 °C (reaction conditions: O2 = 5%, SO2 = 1000 ppm, NO = 500 ppm, HCl = 5 ppm, H2O = 5%, balance N2). | |
3.4.2. Effect of H2O in simulated flue gas. Water vapor, which is unavoidable in coal-fired flue gases, has been reported to suppress Hg0 oxidation and removal over metal oxide catalysts due to competitive adsorption.35,55 Experiments were conducted to determine the effect of H2O (g) on Hg0 removal over 4% Mn/MK10, and the results in Fig. 8 indicate a suppression of the removal efficiency from 99.1% to 78.6% on introduction of 5% H2O (g) into the dry simulated flue gas. The effect is likely the result of the competition of water vapor on active sites inhibited the adsorption of reactive species that have promotional effect on Hg0 removal such as HCl and NO.54 Additionally, H2O (g) might react with SO2 to form H2SO4 and further react with MnO2 to form Mn(SO4)2,56 which could cover the sample surface and prevent contact between Hg0 and the active components.Nevertheless, it should be noted that only 0.05 g of sample (with the high GHSV of 487
000 h−1) was used in these experiments to highlight the effect of H2O (g) on Hg0 removal. According to the literature,55 the suppressant effect of H2O (g) should be mitigated by using a larger amount of sample. In this work, the inhibition by H2O (g) was lessened significantly when 0.30 g of the 4% Mn/MK10 sample was used under the same wet flue gas conditions (Fig. 8). Thus, the industrial application of the Mn/MK10 sample is advantageous and promising for Hg0 removal in coal-fired flue gases.
3.5. Effect of NO/SO2 in the absence of HCl
For coal-fired flue gas in china, it is more useful to investigate Hg0 removal efficiency of the samples without the assistance of HCl. Therefore, in this work, experiments were carried out to study the effect of SO2 on Hg0 removal when adding various concentration of NO into the simulated flue gases. The effect of the NO/SO2 ratio on Hg0 removal efficiency is shown in Fig. 9. In comparison with the efficiency of 29.1% under O2 and SO2, Hg0 removal performance was enhanced with increasing NO/SO2 ratios. And when the NO/SO2 ratio was beyond 0.33, the efficiency was maintained at 99%. There are three possible reasons accounting for this phenomenon. First, the promotion role of NO was much remarkable than the suppression effect of SO2. In contrast to the situation under SO2 and O2, even at a low concentration, NO would react with Hg0 to promote oxidation, and when the concentration of NO exceeded a threshold value, the facilitation effect would play the dominant role. Second, intermediate products derived from the reaction between NO and SO2 could be formed, which would consume some SO2 to alleviate its inhibitory effect.57 Thirdly, as discussed above, the active sites of SO2 and NO might be different, and Hg0 could be preferentially absorbed on NO. Generally, the concentrations of SO2 and NOx in coal-fired flue gases are 340–1120 ppm and 290–580 ppm,57 respectively, with actual NO/SO2 volume ratios of 0.26–1.71, as shown in the shaded area in Fig. 9. The removal efficiency of Hg0 in this region is 80–99%. Thus, the efficiency of the 4% Mn/MK10 sample will not be notably affected in simulated coal-fired flue gases without the assistance of HCl and this sample has great potential for industrial applications.
 |
| | Fig. 9 Hg0 removal efficiency of 4% Mn/MK10 over various NO/SO2 ratios at 150 °C (reaction conditions: NO/SO2 < 0.5, SO2 = 1000 ppm, NO = 125–500 ppm; NO/SO2 > 0.5, NO = 500 ppm, SO2 = 250–1000 ppm. O2 = 5%, balance N2). | |
4. Conclusions
4% Mn/MK10 sample exhibited the highest activity (>90%, 100–400 °C) and stability (>140 h, 150 °C) for Hg0 removal in simulated coal-fired flue gas. Characterization of the sample indicated that amorphous MnO2 and active oxygen were crucial for Hg0 removal. The effects of various gas components revealed that NO plays a promoting role in Hg0 removal and the addition of O2 provided enough surface oxygen for formation of Hg(NO3)2. Moreover, as little as 5 ppm HCl prominently improved the Hg0 removal performance as HCl enhanced Hg0 conversion to HgClx. However, SO2 had a strong inhibitory effect on Hg0 removal. The addition of O2 only resulted in slight remission of the SO2 inhibition effect, but NO and HCl effectively counteracted the observed suppression. H2O also slightly inhibited Hg0 removal, mostly due to competitive adsorption. The Hg0 removal performance was maintained at 80–99% (NO/SO2 = 0.26–1.71) in simulated flue gas without HCl, which appeared to be promising in industrial application.
Acknowledgements
This work was supported by the National Basic Research Program (973) of China (no. 2013CB430005) and the Special Research Funding for Public Benefit Industries from the National Ministry of Environmental Protection (201309018).
References
- M. Brauer, M. Amann, R. T. Burnett, A. Cohen, F. Dentener, M. Ezzati, S. B. Henderson, M. Krzyzanowski, R. V. Martin, R. Van Dingenen, A. van Donkelaar and G. D. Thurston, Environ. Sci. Technol., 2011, 46, 652–660 CrossRef PubMed.
- Y. Liu, C. Tian, B. Yan, Q. Lu, Y. Xie, J. Chen, R. Gupta, Z. Xu, S. M. Kuznicki, Q. Liu and H. Zeng, RSC Adv., 2015, 5, 15634–15640 RSC.
- D. Jampaiah, S. J. Ippolito, Y. M. Sabri, B. M. Reddy and S. K. Bhargava, Catal. Sci. Technol., 2015, 5, 2913–2924 CAS.
- Y. Yang, W. Xu, Y. Wu, J. Xiong, T. Zhu, X. Zhou and L. Tong, RSC Adv., 2016, 6, 59009–59015 RSC.
- Y. Gao, Z. Zhang, J. Wu, L. Duan, A. Umar, L. Sun, Z. Guo and Q. Wang, Environ. Sci. Technol., 2013, 47, 10813–10823 CrossRef CAS PubMed.
- W. Xu, L. Tong, H. Qi, X. Zhou, J. Wang and T. Zhu, Ind. Eng. Chem. Res., 2014, 54, 146–152 CrossRef.
- D. Jampaiah, K. M. Tur, S. J. Ippolito, Y. M. Sabri, J. Tardio, S. K. Bhargava and B. M. Reddy, RSC Adv., 2013, 3, 12963–12974 RSC.
- Y. Shao, J. Li, H. Chang, Y. Peng and Y. Deng, Catal. Sci. Technol., 2015, 5, 3536–3544 CAS.
- S. Cavallaro, N. Bertuccio, P. Antonucci, N. Giordano and J. C. J. Bart, J. Catal., 1982, 73, 337–348 CrossRef CAS.
- Y. Xu, Q. Zhong and X. Liu, J. Hazard. Mater., 2015, 283, 252–259 CrossRef CAS PubMed.
- H. Xu, Z. Qu, C. Zong, W. Huang, F. Quan and N. Yan, Environ. Sci. Technol., 2015, 49, 6823–6830 CrossRef CAS PubMed.
- H. Li, C. Y. Wu, Y. Li and J. Zhang, Appl. Catal., B, 2012, 111, 381–388 CrossRef.
- S. Zhao, Z. Qu, N. Yan, Z. Li, W. Zhu, J. Xu and M. Li, RSC Adv., 2015, 5, 30841–30850 RSC.
- S. Zhao, Z. Qu, N. Yan, Z. Li, H. Xu, J. Mei and F. Quan, Environ. Sci. Technol., 2015, 5, 2985–2993 CAS.
- J. He, G. K. Reddy, S. W. Thiel, P. G. Smirniotis and N. G. Pinto, Energy Fuels, 2013, 27, 4832–4839 CrossRef CAS.
- W. Du, L. Yin, Y. Zhuo, Q. Xu, L. Zhang and C. Chen, Fuel Process. Technol., 2015, 131, 403–408 CrossRef CAS.
- X. Li, Z. Liu, J. Kim and J. Y. Lee, Appl. Catal., B, 2013, 132, 401–407 CrossRef.
- S. Yang, Y. Guo, N. Yan, D. Wu, H. He, Z. Qu and J. Jia, Ind. Eng. Chem. Res., 2011, 50, 9650–9656 CrossRef CAS.
- Z. Wei, Y. Luo, B. Li, Z. Cheng, J. Wang and Q. Ye, Atmos. Pollut. Res., 2015, 6, 45–51 CrossRef CAS.
- C. H. Chiu, H. C. Hsi and H. P. Lin, Fuel Process. Technol., 2015, 126, 138–144 CrossRef.
- D. Shi, Y. Lu, Z. Tang, F. Han, R. Chen and Q. Xu, Korean J. Chem. Eng., 2014, 31, 1405–1412 CrossRef CAS.
- C. He, B. Shen and F. Li, J. Hazard. Mater., 2016, 304, 10–17 CrossRef CAS PubMed.
- F. Ding, Y. Zhao, L. Mi, H. Li, Y. Li and J. Zhang, Ind. Eng. Chem. Res., 2012, 5, 3039–3047 CrossRef.
- I. Fatimah and T. Huda, Appl. Clay Sci., 2013, 74, 115–120 CrossRef CAS.
- S. Zuo, F. Liu, R. Zhou and C. Qi, Catal. Commun., 2012, 22, 1–5 CrossRef CAS.
- M. Tahir and N. S. Amin, Appl. Catal., B, 2013, 142, 512–522 CrossRef.
- L. Stevens, K. Williams, W. Y. Han, T. Drage, J. Wood and J. Wang, Chem. Eng. J., 2013, 215–216, 699–708 CrossRef CAS.
- D. Olszewska, Fuel Process. Technol., 2011, 92, 2412–2419 CrossRef CAS.
- C. Volzone and J. Ortiga, Appl. Clay Sci., 2009, 44, 251–254 CrossRef CAS.
- T. Munusamy, J. Ya-Ting and L. Jiunn-Fwu, RSC Adv., 2015, 5, 23957 RSC.
- R. Zhu, Q. Chen, Q. Zhou, Y. Xi, J. Zhu and H. He, Appl. Clay Sci., 2016, 123, 239–258 CrossRef CAS.
- B. S. Kumar, A. Dhakshinamoorthy and K. Pitchumani, Catal. Sci. Technol., 2014, 4, 2378–2396 CAS.
- J. Qiu, S. Dong, H. Wang, X. Cheng and Z. Du, RSC Adv., 2015, 5, 58284–58291 RSC.
- L. Tian, C. Li, Q. Li, G. Zeng, Z. Gao, S. Li and X. Fan, Fuel, 2009, 88, 1687–1691 CrossRef CAS.
- J. Yang, Y. Zhao, J. Zhang and C. Zheng, Environ. Sci. Technol., 2014, 48, 14837–14843 CrossRef CAS PubMed.
- Z. Wu, B. Jiang and Y. Liu, Appl. Catal., B, 2008, 79, 347–355 CrossRef CAS.
- J. Xie, Z. Qu, N. Yan, S. Yang, W. Chen, L. Hu and W. H. P. Liu, J. Hazard. Mater., 2013, 261, 206–213 CrossRef CAS PubMed.
- Y. Xie, C. Li, L. Zhao, J. Zhang, G. Zeng, X. Zhang, W. Zhang and S. Tao, Appl. Surf. Sci., 2015, 333, 59–67 CrossRef CAS.
- H. Li, C. Y. Wu, Y. Li and J. Zhang, Environ. Sci. Technol., 2011, 45, 7394–7400 CrossRef CAS PubMed.
- Y. Guo, N. Yan, S. Yang, P. Liu, J. Wang, Z. Qu and J. Jia, J. Hazard. Mater., 2012, 213, 62–70 CrossRef PubMed.
- S. Cimino and F. Scala, Ind. Eng. Chem. Res., 2016, 55, 5133–5138 CrossRef CAS.
- M. Rumayor, M. Diaz-Somoano, M. A. Lopez-Anton and M. R. Martinez Tarazona, Chemosphere, 2015, 119, 459–465 CrossRef CAS PubMed.
- A. Murakami, M. A. Uddin, R. Ochiai, E. Sasaoka and S. Wu, Energy Fuels, 2010, 24, 4241–4249 CrossRef CAS.
- S. Wu, R. Katayama, M. A. Uddin, E. Sasaoka and Z. Xie, Energy Fuels, 2015, 29, 6598–6604 CrossRef CAS.
- R. Liu, W. Xu, L. Tong and T. Zhu, J. Environ. Sci., 2015, 36, 76–83 CrossRef PubMed.
- L. Zhao, C. Li, J. Zhang, X. Zhang, F. Zhan, J. Ma, Y. Xie and G. Zeng, Fuel, 2015, 153, 361–369 CrossRef CAS.
- G. Q. And and R. T. Yang, J. Phys. Chem. B, 2004, 108, 15738–15747 CrossRef.
- W. Xu, H. Wang, X. Zhou and T. Zhu, Chem. Eng. J., 2014, 243, 380–385 CrossRef CAS.
- L. Tong, W. Q. Xu, X. Zhou, R. H. Liu and T. Y. Zhu, Energy Fuels, 2015, 29, 5231–5236 CrossRef CAS.
- Y. Zhuang, J. Laumb, R. Liggett, M. Holmes and J. Pavlish, Fuel Process. Technol., 2007, 88, 929–934 CrossRef CAS.
- J. Yang, Y. Zhao, J. Zhang and C. Zheng, Fuel, 2015, 167, 366–374 CrossRef.
- P. Wang, S. Su, J. Xiang, H. You, F. Cao, L. Sun, S. Hu and Y. Zhang, Chemosphere, 2014, 101, 49–54 CrossRef CAS PubMed.
- S. Eswaran and H. G. Stenger, Energy Fuels, 2005, 19, 2328–2334 CrossRef CAS.
- Y. Li, P. D. Murphy, C. Y. Wu, K. W. Powers and J. C. Bonzongo, Environ. Sci. Technol., 2008, 42, 5304–5309 CrossRef CAS PubMed.
- H. Li, C. Y. Wu, Y. Li, L. Li, Y. Zhao and J. Zhang, J. Hazard. Mater., 2012, 243, 117–123 CrossRef CAS PubMed.
- S. Tao, C. Li, X. Fan, G. Zeng, P. Lu, X. Zhang, Q. Wen, W. Zhao, D. Luo and C. Fan, Chem. Eng. J., 2012, 210, 547–556 CrossRef CAS.
- Y. Guo, Y. Li, T. Zhu and M. Ye, Energy Fuels, 2012, 27, 360–366 CrossRef.
|
| This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.