Next Article in Journal
Continuous Estimation of Finger and Wrist Joint Angles Using a Muscle Synergy Based Musculoskeletal Model
Previous Article in Journal
Influence of Unburned Carbon on Environmental-Technical Behaviour of Coal Fly Ash Fired Clay Bricks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Study on the Elemental Mercury Removal Performance and Regeneration Ability of CoOx–FeOx-Modified ZSM-5 Adsorbents

1
College of Quality & Safety Engineering, China Jiliang University, Hangzhou 310018, China
2
School of Resource & Environmental Engineering, Jiangxi University of Science & Technology, Ganzhou 341000, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(8), 3769; https://doi.org/10.3390/app12083769
Submission received: 11 March 2022 / Revised: 27 March 2022 / Accepted: 1 April 2022 / Published: 8 April 2022

Abstract

:
Herein, a series of Co-Fe mixed oxide modified ZSM-5 adsorbents were synthesized using the ultrasonic-assisted impregnation method for the capture of elemental mercury. In comparison with other samples, Co4Fe1-ZSM-5 produced a relatively better performance, with the removal efficiency of around 96.6% Hg0 and the adsorption capacity of around 901.63 ug/mg Hg0 achieved at 120 °C. The interaction between CoOx and FeOx improved the reducibility of oxygen species, thus promoting the oxidation of Hg0. Among a variety of other gas components, O2 and NO exerted a positive effect on Hg0, which improved its removal to a certain extent. By contrast, SO2 caused an adverse effect on the capture of Hg0, which could be reversed to some degree by the introduction of 5% O2. After five cycles, the mercury removal efficiency of Co4Fe1-ZSM-5 remained above 90%, suggesting excellent recyclability. Finally, XPS analysis was conducted to reveal that Mars–Maessen mechanisms are dominant in the process of mercury adsorption.

1. Introduction

Mercury is a universal contaminant due to its toxicity, worldwide migration, and bioaccumulation in the food chain and the environment. Among various sources of contamination, coal-fired power plants have been found responsible for the large majority of atmospheric mercury emission [1]. In general, the mercury released from coal-fired flue gas can be classified into three categories: elemental mercury (Hg0), oxidized mercury (Hg2+), and particulate-bonded mercury (HgP) [2,3]. As for HgP and Hg2+, they can be easily removed by the existing air pollution control devices (APCDs) [4]. Nevertheless, it is practically difficult to control Hg0 due to its low solubility and high volatility [5]. Therefore, it is necessary to develop an efficient and environmentally friendly technology that is applicable to remove Hg0 from flue gases.
Up until now, there have been various technologies tested to reduce Hg0 emissions, such as adsorbent injection, plasma, catalytic oxidation, photochemical oxidation, and so on [6,7,8,9,10]. Among them, adsorbent injection, as represented by activated carbon injection (ACI), has been widely recognized as the most promising method of Hg0 removal [11]. In the meantime, however, it is also disadvantaged by drawbacks such as low Hg0 removal efficiency, high operational cost, and the deteriorating quality and reusability of fly ash, all of which impede the extensive application of this technology in practice [12,13,14]. Therefore, it is imperative to develop a less costly but more efficient kind of mercury adsorbent. Recently, ZSM-5 has been proposed as an extraordinary candidate adsorbent for mercury removal due to its high specific surface area and adjustable pore size [15,16]. However, considering the lack of surface-active sites for virgin ZSM-5, it is necessary to improve Hg0 removal performance by modifying ZSM-5 [17]. As reported in the relevant literature, halides, selenides, and metal oxides can be applied as the active additives for Hg0 capture [18,19,20,21,22]. Due to the ease of operation, environmental friendliness, and low cost, metal oxide additives are deemed more attractive than others.
Cobalt oxides, with their unique redox couple Co2+/Co3+, have now attracted widespread attention for Hg0 removal [23,24,25,26,27,28]. In the research conducted by Zhang et al., it was suggested that the excellent Hg0 removal performance could be achieved at 180 °C, which synthesized a Co3O4-based catalyst by one-step calcination method for Hg0 oxidation [29]. Xu et al. reported that Co3O4-modified fly ash by the wet impregnation method showed good Hg0 removal performance [30]. Using a single-step hydrothermal method to produce the cobalt-doped CeO2 catalyst for Hg0 removal, Yang et al. achieved a Hg0 removal efficiency of over 90% under a simulated low-rank coal-burning atmosphere [31]. Moreover, Fe-based catalysts performed equally well in the Hg0 removal process. As indicated by Xu et al., Fe-modified HZSM-5 produced a remarkable Hg0 removal performance [32]. Yang et al. proposed that the synthesized magnetic Fe-based adsorbent achieved a satisfactory performance in mercury removal [33]. In spite of this, there is still limited research conducted on the efficiency of Co-Fe-modified ZSM-5 adsorbent in Hg0 removal. Given the above-mentioned advantages of ZSM-5, CoOx, and FeOx, it is believed that the efficiency of Hg0 removal could be further enhanced for Co-Fe-modified ZSM-5 materials.
To achieve this purpose, Co-Fe-decorated ZSM-5 adsorbents were produced in this study to explore their potential in the abatement of Hg0 in flue gases. To be specific, the physicochemical properties of the adsorbents were characterized through a combination of X-ray diffraction (XRD), N2 adsorption, scanning electron microscope (SEM), and X-ray photoelectron spectroscopy (XPS). Then, it was investigated as to how the Hg0 removal performance of the adsorbents would be affected by Co/Fe mass ratio, reaction temperature, and flue gas components (O2, NO, SO2). Finally, the level of recyclability and mechanisms of Hg0 adsorption were examined.

2. Experimental

2.1. Sample Preparation

Commercial Zeolite Socony Mobil-5 (ZSM-5 for short), the Si/Al molar ratio (n(SiO2)/n(Al2O3)) of which is around 140, was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China)
The mercury adsorbent was synthesized using the ultrasound-assisted impregnation method. Firstly, ZSM-5 was calcined at 500 °C for 2 h to remove all impurities, thus enhancing the validity and reliability of the research result. Then, the cobalt nitrate and ferric nitrate of calculated amount were dissolved in deionized water to obtain a transparent solution, before being mixed with a certain amount of ZSM-5. After 12 h of stirring and 45 min of ultrasonic treatment, the sample was dried at 105 °C for 6 h and finally calcined at 450 °C in a muffle furnace in the air for 4 h. The obtained sample was labeled CoxFe1-ZSM-5, where x denotes the Co/Fe mass ratio. The Co/Fe mass ratio of the prepared samples ranged from 1 to 4. A similar method was used to synthesize Co-ZSM-5 and Fe-ZSM-5 with the mixed solution replaced by ferric nitrate aqueous solution and cobalt nitrate aqueous solution. The (Co + Fe) mass ratio of all the samples was set to 8 wt%.

2.2. Sample Characterization

IR spectra were obtained through a FTIR spectrometer (GanDong SCI.&TECH. Co., Ltd., Tianjing, China) ranging from 400 to 4000 cm−1 at a spectral resolution of 4 cm−1.
An N2 adsorption experiment was carried out for several adsorbents on an Autosorb-iQ (Quantachrome Instruments, Boynton Beach, Florida) instrument. Prior to the experiment, the tested sample was degassed at 200 °C for 6 h. The specific surface areas of the tested samples were calculated using the Brunauer–Emmett–Teller (BET) equation, while the density functional theory (DFT) method was used to determine pore size distributions.
The crystal structure of a series of adsorbents was determined using an Empyrean (PANalytical B.V., Almelo, The Nederlands) X-ray diffraction device equipped with Cu Ka as the radian source in the range of 5–80° at a scanning rate of 2 °/min. Micro-structure and morphology of the modified samples were investigated on an EVO 18 (ZEISS, Oberkochen, Germany) coupled with energy dispersive X-ray spectroscopy to examine the chemical composition.
H2 temperature-programmed reduction (H2-TPR) was analyzed by means of AutoChem II 2920 (Micromeritics Instruments, Norcross, GA, USA). Before the experiment was conducted, the tested sample was pretreated at 300 °C for 30 min. After cooling to room temperature, its temperature was raised again to 800 °C at a heating rate of 10 °/min. The H2 signal was continuously recorded by a thermal conductivity detector (TCD).
The element chemical state and chemical composition of the samples were examined using a Thermo ESCALAB 250XI (Thermo Fisher Scientific Inc., Waltham, MA, USA) equipped with Al Kα X-ray source. All the binding energies were calibrated with the C1s line at 284.6 eV.
The metal contents of modified samples were analyzed by an Inductively Coupled Plasma-Atomic Emission Spectrometry (SPECTRO Analytical Instruments, Kleve, Germany).

2.3. Experimental Setup and Procedure

The experimental configuration intended for Hg0 adsorption is illustrated in Figure 1. The entire experimental system consists of a mercury generation system, an electric tube furnace reactor, an online mercury detector, an online data acquisition system, and an exhaust gas absorption device.
The performance test was carried out using a 0.4 g sample. Hg0 vapour (≈200 μg/m3) was generated through a mercury permeation tube placed in a U-shaped quartz tube. The temperature of the water bath was maintained at 50 °C. The total flow rate of the flue gas was set to 600 mL/min, with 5% O2 500 ppm NO (when used), 800 ppm SO2 (when used), and N2 contained as the balanced gas. Both inlet and outlet Hg0 concentrations were monitored by an online mercury analyzer RA-915M (LUMEX Ltd., St. Petersburg, Russia). Hg0 removal efficiency was calculated using the following equation [33]:
η = C in C out C in × 100 %
where ƞ represents the Hg0 removal efficiency (%), while Cin and Cout (µg/m3) refer to the inlet and outlet mercury concentrations.
Hg0 uptake capacity qt (μg/g) was determined for the sample using Equation (2) [33], where M represents the mass (g) of samples, f indicates the total flow rate (mL/min) of flue gas, and t denotes the adsorption time (min) required for Hg0 capture. Each test lasted for 90 min to calculate Hg0 removal efficiency and Hg0 adsorption capacity.
q t = 1 M t 1 t 2 C in C out   f d t
The pseudo-first-order model and pseudo-second-order model were applied to analyze the process of Hg0 capture for the estimate of Hg0 adsorption capacity. The pseudo-first-order model is expressed as follows [23,34]:
q t = q e × 1 e k 1 t
The pseudo-second-order model is expressed as the following equation [34]:
q t = q e 2 k 2 t 1 + q e k 2 t
where qt and qe represent the level of mercury adsorption capacity at time t (μg/g) and the equilibrium point, respectively, while k1 and k2 refer to the rate constant of pseudo-first-order-model and pseudo-second-order-model, respectively.
Recyclability of the adsorbents was carried out in the above-mentioned fixed-bed micro-reactor. Briefly, after the adsorption process, the spent Co4Fe1-ZSM-5 was treated in certain flue gases at different temperatures for 2 h. Then, the sample was cooled down to the desired temperature point for the next Hg0 removal performance evaluation cycle.

3. Results and Discussion

3.1. Performance of a Series Samples

Figure 2 shows the performance of the adsorbents at different Co/Fe mass ratios in O2-containing flue gas at a temperature ranging between 60 and 180 °C. The pure ZSM-5 sample exhibited a relatively low level of efficiency in Hg0 removal. Given the modification to CoOx or FeOx, there was an obvious improvement detected in the efficiency of Hg0 removal. The mercury removal efficiency of CoxFe1-ZSM-5 sorbents was enhanced with the increase of Co mass ratio. Compared with other samples, Co4Fe1-ZSM-5 and Co-ZSM-5 achieved a satisfactory performance in mercury removal. Hg0 removal efficiency reached up to around 96%. In order to identify the variation in performance between Co4Fe1-ZSM-5 and Co-ZSM-5, the theoretical capacity of Hg0 adsorption by these samples was calculated (Figure S1). It was found out that Co4Fe1-ZSM-5 performed better in Hg0 removal, as evidenced by the higher theoretical capacity of Hg0 adsorption. According to these results, there must be a significant synergistic effect at play between cobalt oxides and iron oxide, thus promoting mercury removal. Moreover, Co4Fe1-ZSM-5 is applicable as an optimal sample for further experimentation. The theoretical capacity of Hg0 capture by Fe-ZSM-5 is extremely low, which is consistent with the result of Hg0 removal efficiency, as shown in Figure 2a. Additionally, it is worth noting that the Hg0 removal efficiency profiles of CoxFe1-ZSM-5 resemble a volcano, which is attributable to the increasingly active reactant with the rise of temperature and the desorption of mercury species from the adsorbent surfaces at an excessively high reaction temperature [33,35]. Compared with the reported adsorbents (Table S1), it seems that Co4Fe1ZSM-5 can be a potential adsorbent for the practical application.

3.2. Physicochemical Properties of the Samples

3.2.1. FTIR Analysis

FTIR was involved in exploring the surface nature of various samples. As shown in Figure 3, all of the samples exhibited six characteristic peaks in the range of 400 and 2000 cm−1, suggesting no significant effect of loading Co-Fe mixed oxides on the surface functional groups of ZSM-5. As demonstrated in the previous research, the bands at 796 and 1233 cm−1 are related to the symmetric stretching vibrations of Si-O-Si [36]. Moreover, the peaks at 448 and 1102 cm−1 corresponded to the Si-O-Si asymmetric stretching vibrations. Conversely, the characteristic peak at 543 cm−1 was assigned to the double six-membered ring, which indicates the existence of micropore structure in ZSM-5 [17,36]. Moreover, the bending vibration of hydroxyl species accounted for the appearance of the peak at 1632 cm−1 [37].

3.2.2. XRD Analysis

Figure 4 shows the XRD patterns of these four samples under study. All of the samples exhibited typical ZSM-5 peaks (2θ = 23.1°, 23.3°, and 24.4° PDF-ICDD 44-0003), suggesting the presence of MFI structure. Moreover, it can be seen that there were no evident variations in the peaks of XRD diffraction that could be detected for these three modified samples, indicating the limited impact of doping active components on the crystal structure of the support [38]. As for Co-ZSM-5 and Fe-ZSM-5 samples, the peaks at 36.8° indexed to Co3O4 (PDF-ICDD 42-1467) and 33.1° attributed to Fe2O3 (PDF-ICDD 33-0664) were found distinct, suggesting the loading of Co and Fe beyond the theoretical monolayer coverage on the ZSM-5 surface. With regard to Co4Fe1-ZSM-5, the appearance of diffraction peaks was attributed to the disappearance of iron oxides, suggesting the potential existence of iron oxides in a highly dispersed state on the surface of ZSM-5 [39,40]. Moreover, compared with Co4Fe1-ZSM-5 and Co-ZSM-5, the peaks indexed to cobalt oxides were weakened and widened, implying that the introduction of Fe2O3 exerts an inhibitory effect on the growth of Co3O4, thus reducing the size of cobalt oxide particles. As a consequence, there might be more active sites provided for the reactions [41].

3.2.3. BET Analysis

N2 adsorption experiments were conducted to investigate the structural properties of these samples. As shown in Figure 5a, all of these samples exhibited typical type IV isotherms and type H3 hysteresis loops, implying the existence of mesopores with slit-shaped structures [2]. In addition, it is worth noting that all these samples displayed a sharp rise of nitrogen uptake in low relative pressure regions, signaling the presence of abundant micropore structures [42]. Figure 5b shows the corresponding pore size distributions of various samples, which confirms the co-existence between micropores and mesopores [41,43]. The N2 adsorption data are listed in Table 1, from which it can be found out that the adsorbent micropore surface area shrunk upon the introduction of the active species, but with no obvious variations detected in the external surface area. This result indicates that most of the added metal oxides entered the micropore channels to obstruct them. In addition, the Co4Fe1-ZSM-5 sample had the largest external surface area, which means the existence of iron oxide contributes to a satisfactory outcome of cobalt oxide distribution. This is consistent with the result of XRD measurement. Moreover, the relatively large specific surface area might provide an increased number of active sites for Hg0 capture, which can partially explain the higher Hg0 removal efficiency of Co4Fe1-ZSM-5 than the other samples.

3.2.4. SEM Analysis

Figure 6 shows the SEM images of virgin ZSM-5 and Co4Fe1-ZSM-5. As can be seen from Figure 6a,b, the surface of ZSM-5 showed smoothness. Differently, there were some tiny particles emerging on the surface of ZSM-5 for Co4Fe1-ZSM-5, suggesting the successful introduction of active species. Meanwhile, there were no significant variations detected in the structure of the modified samples, suggesting that morphology is not the main reason for the variations in the Hg0 removal efficiency of the serial adsorbents.
EDX was used to characterize the distributions of Fe and Co species over ZSM-5. As shown in Figure 7b, Co3O4 with bright contrast was dispersed over the entire surface of ZSM-5. The corresponding elemental maps of ferric also imply that Fe2O3 was well distributed throughout the ZSM-5. EDX analysis (Figure 7d) displays that the Fe/Co mass ratio of Co4Fe1-ZSM-5 was 3.9, which matched the ICP result well (Table S2).

3.2.5. H2-TPR Analysis

The redox properties of ZSM-5 and modified ZSM-5 were studied using H2-TPR, the results of which are shown in Figure 8. Regarding virgin ZSM-5, there were a few signs of hydrogen consumption detected, suggesting the absence of reducible species available for the capture of Hg0. For all the modified ZSM-5, there were a number of evident peaks of H2 reduction. The reduction of Fe-ZSM-5 after deconvolution can be divided into two stages. Located at 346 °C, the first sub-peak is possibly attributed to the reduction of Fe2O3 to Fe3O4, whereas the reduction of Fe3O4 to FeO can account for the appearance of the peak at 485 °C [44]. In respect of Co-ZSM-5, the peaks centered at 328 and 357 °C were associated with the reduction of Co3O4 to CoO and CoO to metallic cobalt, respectively [45]. Co4Fe-ZSM-5 manifests four peaks at 316, 361, 404, and 436 °C. The first one is suspected to result from the reduction of Co3O4 to CoO, the second and third ones are probably attributable to the reduction of CoO and Fe2O3, and the last one is ascribed to the reduction of Fe3O4. Notably, there appears to be a decline of temperature for the reduction peaks, suggesting the improved reducibility of Co4Fe1-ZSM-5 due to the synergistic effect between CoOx and FeOx [23]. This would facilitate the oxidation of Hg0, constituting another reason for the elevated Hg0 removal efficiency of Co4Fe1-ZSM-5.

3.2.6. XPS Analysis

To demonstrate the chemical properties of the adsorbent surface, the XPS method was used to explore the atom environment of the samples. As shown in Figure 9, for Co 2p spectra, there were two distinct peaks detected at around 780 and 795 eV, which can be attributed to Co 2p3/2 and Co 2p1/2, respectively. Following deconvolution, there were two spin-orbit doublets (D1 and D2) emerging, along with three satellite peaks observed (S1, S2, and S3). Located at 780.0~780.1 eV and 795.0~795.1 eV, D1 bands were attributable to Co3+ species. By comparison, the D2 bands located at 781.7 and 797.4 eV corresponded to Co2+ species, suggesting the co-existence between Co2+ and Co3+ [2]. A similar conclusion can be drawn for Fe atoms that both Fe3+ and Fe2+ were observable in the samples [46]. Upon calculation (Table 2), it can be found out that the addition of Fe2O3 into Co-ZSM-5 increased the ratio Co3+/Co2+ from 0.77 to 1.37 and caused the ratio Fe3+/Fe2+ to drop sharply from 1.62 to 1.29 with the Fe-containing samples as modified by CoOx, which indicates that the conversion of Co2+ to Co3+ can be accelerated by the interaction between Co species and Fe species. This has been proved as highly active in catalytic reactions, thus enhancing the activity of the Co4Fe1-ZSM-5 sample.

3.3. Effects of Gas Components

3.3.1. Effects of O2

As shown in Figure 10, the mercury removal efficiency of Co4Fe1-ZSM-5 in N2 reached 73.4% at 120 °C. In this process, the lattice oxygen of the adsorbents played an important role. With O2 concentration rising to 5%, the mercury removal efficiency showed a significant increase to 96.6%. This was because O2 could compensate for the lattice oxygen consumed in the reaction, thus improving the outcome of Hg0 capture [47].

3.3.2. Effects of NO

As shown in Figure 10, the introduction of 500 ppm NO into pure N2 flue gas led to a significant improvement of Hg0 removal efficiency, which can be accounted for by the capability of NO to react with the surface oxygen species of CoOx and FeOx from reactive NOx species. This promotes the Hg0 heterogeneous oxidation occurring on the adsorbent surface [48]. However, in the context of O2- and NO-containing flue gas, the impressive removal efficiency of Hg0, close to 100%, was achievable. It is suggested that NO is conducive to the removal of Hg0, regardless of the presence of O2, which is supposed to result from O2 accelerating the formation of reactive NOx species. Thus, the enhancement of Hg0 removal efficiency is achieved [35].

3.3.3. Effects of SO2

With 600 ppm SO2 added into the reaction system, Hg0 removal efficiency declined from 73.4% to 56.1% in the absence of O2. Notably, the level of Hg0 removal efficiency in the presence of SO2 and O2 remained lower than in the flue gas without SO2, which implies that SO2 negatively affects Hg0 removal. Immediately after the increase of concentration of SO2, such an inhibitory effect was enhanced, for which there were two reasons. On the one hand, SO2 and Hg share the same adsorption sites on the adsorbent surface site, with competitive adsorption occurring between Hg0 and SO2 [49]. On the other hand, SO2 reacts with surface-active components to form metal sulfate salts, which reduces the number of available surface-active sites, thus interfering with the Hg0 adsorption process [50]. After the addition of 5% O2, this inhibitory effect diminishes, which is accompanied by a slight increase in the level of Hg0 removal efficiency. This is suspected to result from the capacity of O2 to compensate for the consumed lattice oxygen. Eventually, the oxidation of Hg0 to HgO as deposited onto adsorbent surface subsequently is enhanced [42].

3.4. Regeneration Ability

The mercury temperature-programmed desorption method was applied to further analyze the mercury species on Co4Fe1-ZSM-5, which is also suited to determining the conditions required for regeneration of the adsorbents. As shown in Figure 11a, the Hg-TPD profile of Co4Fe1-ZSM-5 fit well into three peaks, which was ascribed to the desorption of physically adsorbed Hg0 and the decomposition of weakly adsorbed HgO (Hg-O) and strong bond HgO (Hg≡O) [51,52,53]. It suggests the dominance of chemisorption over Co4Fe1-ZSM-5 in the Hg0 adsorption process, which means the possibility to obtain a clean adsorbent surface after treatment at 500 °C and achieve a satisfactory performance in Hg0 capture. To demonstrate this point, a test was conducted on the Hg0 capture efficiency of the adsorbents regenerated at different temperatures, which revealed that the efficiency of the sample regenerated at 500 °C was higher than that regenerated at 300 and 400 °C, confirming that 500 °C is the optimal temperature for the regeneration of the spent adsorbents. Apart from that, the regeneration atmosphere is another important influencing factor in the recyclability of the adsorbents.
As shown in Figure 11c, the Hg0 removal efficiency of adsorbent was just around 85%. With the addition of 5% O2, however, it improved to around 95%, suggesting that O2 is conducive to improving the performance of the regenerated sample in Hg0 adsorption. Despite a further increase in the level of O2 concentration, there were no evident variations observed in Hg0 removal efficiency, indicating that 5% O2 is sufficient for restoring the Hg0 capture performance of the adsorbent. After five cycles of recycling (Figure 11d), adsorbent Hg0 removal efficiency remained basically unchanged above 90%, which could be accounted for by the insignificant variation in the physicochemical properties of the adsorbents after the adsorption and regeneration process (Figure S2, Table S3). It demonstrates the excellent recyclability of the Co4Fe1-ZSM-5 sample and its applicability as an adsorbent.

3.5. Mercury Capture Mechanism

Finally, XPS analysis was conducted to explore the adsorption mechanisms of Co4Fe1-ZSM-5. Figure 12a,b shows the Co 2p spectra of the fresh and spent samples. It can be discovered from the figures that both ratios of Co3+/Co2+ (from 1.37 to 0.84) and Fe3+/Fe2+ (from 1.29 to 0.93) declined after Hg0 adsorption, implying that Co3+ and Fe3+ play a crucial role in the removal of Hg0 and undergo a redox-reduction reaction to a lower valance state [2].
As can be seen from Figure 12c, the asymmetry of O 1s spectra suggested the existence of various oxygen species. After curve-fitting, those peaks in the range of 530.2–530.5 eV could be indexed to the lattice oxygen of metal oxides (denoted as Oα), while the peaks at 532.9–533.2 eV were associated with the lattice oxygen of ZSM-5 (denoted as Oβ), as shown in Figure 12c [17]. According to Table 2, the concentration of Oα dropped to 5.9%, suggesting that the lattice oxygen of metal oxides was primarily responsible for the oxidation of Hg0.
As for Hg 4f spectra, there were two photoelectron peaks detected at 103.2 and 104.0 eV after deconvolution (Figure 12d). The former can be assigned to Si 2p for SiO2, while the latter was attributable to Hg 4f5/2 for HgO, being consistent with the Hg-TPD result shown in Figure 11a that the captured mercury species existed mainly in the form of HgO on the adsorbent surfaces. Moreover, it can be concluded that the Mars–Maessen reaction routine is dominant in the Hg0 adsorption process over Co4Fe1-ZSM-5 (Equations (5)–(12)) [45,54,55]. More specifically, Hg0 is first physically adsorbed on the adsorbent surface, which is followed by the combination with lattice oxygen of metal oxides to form HgO. Meanwhile, Co3+ and Fe3+ are reduced. Finally, gaseous O2 acts as an oxidant to re-oxidize the reduced cobalt and iron species into Co3+ and Fe3+.
Reaction condition: N2 + 5% O2 + 200 μg/m3 Hg0.
Hg 0 g Hg 0 ads
Hg 0 ads + C o x O y Hg-O C o x O y 1
Hg 0 ads + F e x O y Hg-O F e x O y 1
Hg 0 ads + C o x O y Hg C o x O y
Hg 0 ads + F e x O y Hg F e x O y
HgO   ads HgO   g
Co x O y 1 + 1 / 2 O 2 Co x O y
Fe x O y 1 + 1 / 2 O 2 Fe x O y
On the basis of this result, it can be concluded that adsorbent oxidative ability contributed to the capture of Hg0 for the existence of the adsorbed mercury species in the form of Hg-O-M compounds. On the basis of the H2-TPR profiles shown in Figure 7, Co4Fe1-ZSM-5 adsorbent exhibited a relative high oxidative ability, which, to some extent, had a positive effect on the oxidation of Hg0 and thus stimulated the capture of Hg0. Meanwhile, the optimized textural structures could be beneficial for the mass transfer of the reactant, enabling Co4Fe1-ZSM-5 with a satisfactory Hg0 removal performance. In addition, Co4Fe1-ZSM-5 was found to effectively capture Hg0 in SO2- or NO-containing flue gas. It possesses a superior recycle attribute, hindering that this material can be a potential adsorbent for the practical application.

4. Conclusions

This study explored a series of Co-Fe mixed oxides supported on ZSM-5, as synthesized using an ultrasound-assisted impregnation method for the abatement of Hg0. The conclusions of this study are presented as follows:
  • Compared with other samples, Co4Fe1-ZSM-5 performed relatively better, with a roughly 96.6% Hg0 removal efficiency and a roughly 901.63 μg/g Hg0 adsorption capacity achieved at 120 °C. The interaction between CoOx and FeOx enhanced the reducibility of oxygen species, which promoted the oxidation of Hg0.
  • Among various gas components, O2 and NO had a positive impact the removal of Hg0, whereas SO2 inhibited the adsorption of Hg0 to a certain extent.
  • The Mars–Masson mechanism dominated the Hg0 adsorption process over Co4Fe1-ZSM-5. Co3+ and Fe3+ provided the active sites required to oxidize Hg0 into Hg0 for the subsequent deposition onto the adsorbent surface.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app12083769/s1, Figure S1: (a) simulated adsorption capacity of Co4Fe1-ZSM-5, Co-ZSM-5 and Fe-ZSM-5; (b) breakthrough curve of Co4Fe1-ZSM-5, Co-ZSM-5 and Fe-ZSM-5 under N2 + 5%O2 + 200 μg/m3, t = 540 min; Figure S2: (a) N2 adsorption-desorption isotherms and (b) pore distributions of the fresh and regeneration samples; (c) XRD pattern of the fresh and regeneration sample; Table S1: Comparison of the Hg0 removal performances onto various adsorbents; Table S2: The content of metal in samples obtained by ICP-AES; Table S3: N2 adsorption results of the fresh and regeneration samples.

Author Contributions

Conceptualization, W.M. and D.Y.; methodology, W.M. and D.Y.; validation, W.M., D.Y. and H.W.; formal analysis, W.M.; investigation, W.M.; resources, H.W.; data curation, W.M.; writ-ing—original draft preparation, W.M.; writing—review and editing, D.Y.; visualization, W.M.; supervision, H.W.; project administration, H.W.; funding acquisition, H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was supported by the National Natural Science Foundation of China (51568024).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, W.; Hussain, A.; Liu, Y. A review on modification methods of adsorbents for elemental mercury from flue gas. Chem. Eng. J. 2018, 346, 692–711. [Google Scholar] [CrossRef]
  2. Ma, Y.; Mu, B.; Zhang, X.; Xu, H.; Qu, Z.; Gao, L.; Li, B.; Tian, J. Ag-Fe3O4@rGO ternary magnetic adsorbent for gaseous elemental mercury removal from coal-fired flue gas. Fuel 2019, 239, 579–586. [Google Scholar] [CrossRef]
  3. Ma, Y.; Mu, B.; Zhang, X.; Zhang, H.; Xu, H.; Qu, Z.; Gao, L. Hierarchical Ag-SiO2@Fe3O4 magnetic composites for elemental mercury removal from non-ferrous metal smelting flue gas. J. Environ. Sci. 2019, 79, 111–120. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, H.; Yang, G.; Gao, X.; Pang, C.H.; Kingman, S.W.; Wu, T. Hg0 Capture over CoMoS/γ-Al2O3 with MoS2 Nanosheets at Low Temperatures. Environ. Sci. Technol. 2016, 50, 1056–1064. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, H.-w.; Chen, J.-y.; Zhao, K.; Niu, Q.-X.; Wang, L. Removal of vapor-phase elemental mercury from simulated syngas using semi-coke modified by Mn/Ce doping. J. Fuel. Chem. Technol. 2016, 44, 394–400. [Google Scholar] [CrossRef]
  6. Zhang, C.; Song, W.; Zhang, X.; Li, R.; Zhao, S.; Fan, C. Synthesis of CeO2-modified activated carbon spheres by grafting and coordinating reactions for elemental mercury removal. J. Mater. Sci. 2019, 54, 2836–2852. [Google Scholar] [CrossRef]
  7. Li, J.; Yan, N.; Qu, Z.; Qiao, S.; Yang, S.; Guo, Y.; Liu, P.; Jia, J. Catalytic oxidation of elemental mercury over the modified catalyst Mn/alpha-Al2O3 at lower temperatures. Environ. Sci. Technol. 2010, 44, 426–431. [Google Scholar] [CrossRef]
  8. Wang, T.; Liu, J.; Zhang, Y.; Zhang, H.; Chen, W.-Y.; Norris, P.; Pan, W.-P. Use of a non-thermal plasma technique to increase the number of chlorine active sites on biochar for improved mercury removal. Chem. Eng. J. 2018, 331, 536–544. [Google Scholar] [CrossRef]
  9. Liu, Y.; Wang, Y. Gaseous elemental mercury removal using VUV and heat coactivation of Oxone/H2O/O2 in a VUV-spraying reactor. Fuel 2019, 243, 352–361. [Google Scholar] [CrossRef]
  10. Guan, Y.; Hu, T.; Wu, J.; Zhao, L.; Tian, F.; Pan, W.; He, P.; Qi, X.; Li, F.; Xu, K. Enhanced photocatalytic activity of TiO2/graphene by tailoring oxidation degrees of graphene oxide for gaseous mercury removal. Korean J. Chem. Eng. 2019, 36, 115–125. [Google Scholar] [CrossRef]
  11. Chen, C.; Duan, Y.; Zhao, S.; Hu, B.; Li, N.; Yao, T.; Zhao, Y.; Wei, H.; Ren, S. Experimental Study on Mercury Removal and Regeneration of SO2 Modified Activated Carbon. Ind. Eng. Chem. 2019, 58, 13190–13197. [Google Scholar] [CrossRef]
  12. Ma, Y.; Mu, B.; Zhang, X.; Yuan, D.; Ma, C.; Xu, H.; Qu, Z.; Fang, S. Graphene enhanced Mn-Ce binary metal oxides for catalytic oxidation and adsorption of elemental mercury from coal-fired flue gas. Chem. Eng. J. 2019, 358, 1499–1506. [Google Scholar] [CrossRef]
  13. Xu, W.; Adewuyi, Y.G.; Liu, Y.; Wang, Y. Removal of elemental mercury from flue gas using CuOx and CeO2 modified rice straw chars enhanced by ultrasound. Fuel Process. Technol. 2018, 170, 21–31. [Google Scholar] [CrossRef]
  14. Wang, P.; Su, S.; Xiang, J.; You, H.; Cao, F.; Sun, L.; Hu, S.; Zhang, Y. Catalytic oxidation of Hg0 by MnOx–CeO2/γ-Al2O3 catalyst at low temperatures. Chemosphere 2014, 101, 49–54. [Google Scholar] [CrossRef]
  15. Fan, X.; Li, C.; Zeng, G.; Zhang, X.; Tao, S.; Lu, P.; Tan, Y.; Luo, D. Hg0 Removal from Simulated Flue Gas over CeO2/HZSM-5. Energy Fuel 2012, 26, 2082–2089. [Google Scholar] [CrossRef]
  16. Xiao, H.; Dou, C.; Li, J.; Yuan, Z.; Lv, H. Experimental Study on SO2-to-SO3 Conversion Over Fe-Modified Mn/ZSM-5 Catalysts During the Catalytic Reduction of NOx. Catal. Surv. Asia 2019, 23, 332–343. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Wu, J.; Li, B.; Xu, H.; Liu, D. Removal of elemental mercury from simulated flue gas by ZSM-5 modified with Mn-Fe mixed oxides. Chem. Eng. J. 2019, 375, 121946. [Google Scholar] [CrossRef]
  18. Huo, Q.; Wang, Y.; Chen, H.; Han, L.; Wang, J.; Bao, W.; Chang, L.; Xie, K. ZnS/AC sorbent derived from the high sulfur petroleum coke for mercury removal. Fuel Process. Technol. 2019, 191, 36–43. [Google Scholar] [CrossRef]
  19. Quan, Z.; Huang, W.; Liao, Y.; Liu, W.; Xu, H.; Yan, N.; Qu, Z. Study on the regenerable sulfur-resistant sorbent for mercury removal from nonferrous metal smelting flue gas. Fuel 2019, 241, 451–458. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Han, X.; Huang, Z.; Hou, Y.; Guo, Y.; Wu, M. Superior activity of CeO2 modified V2O5/AC catalyst for mercury removal at low temperature. Chem. Eng. J. 2018, 337, 741–749. [Google Scholar] [CrossRef]
  21. Xu, H.; Qu, Z.; Zong, C.; Huang, W.; Quan, F.; Yan, N. MnOx/Graphene for the Catalytic Oxidation and Adsorption of Elemental Mercury. Environ. Sci. Technol. 2015, 49, 6823–6830. [Google Scholar] [CrossRef]
  22. Li, X.; Zhou, J.; Zhou, Q.; Mao, J. Removal of elemental mercury using titania sorbents loaded with cobalt ceria oxides from syngas. New J. Chem. 2018, 42, 12503–12510. [Google Scholar] [CrossRef]
  23. Yang, W.; Chen, H.; Han, X.; Ding, S.; Shan, Y.; Liu, Y. Preparation of magnetic Co-Fe modified porous carbon from agricultural wastes by microwave and steam activation for mercury removal. J. Hazard. Mater. 2020, 381, 120981. [Google Scholar] [CrossRef]
  24. Zhang, X.; Tan, B.; Wang, J.; Zhang, H.; Li, C.; He, G. Removal of elemental mercury by Ce and Co modified MCM-41 catalyst from simulated flue gas. Can. J. Chem. Eng. 2019, 97, 734–742. [Google Scholar] [CrossRef]
  25. Deng, S.; Chen, N.; Deng, D.; Li, Y.; Xing, X.; Wang, Y. Meso- and macroporous coral-like Co3O4 for VOCs gas sensor. Ceram. Int. 2015, 41 Pt A, 11004–11012. [Google Scholar] [CrossRef]
  26. Oton, L.F.; Oliveira, A.C.; de Araujo, J.C.S.; Araujo, R.S.; de Sousa, F.F.; Saraiva, G.D.; Lang, R.; Otubo, L.; Carlos da Silva Duarte, G.; Campos, A. Selective catalytic reduction of NOx by CO (CO-SCR) over metal-supported nanoparticles dispersed on porous alumina. Adv. Powder Technol. 2020, 31, 464–476. [Google Scholar] [CrossRef]
  27. Silas, K.; Karim, A.; Choong, T.; Rashid, T.D.U. Optimization of Activated Carbon Monolith Co3O4-Based Catalyst for Simultaneous SO2/NOx Removal from Flue Gas Using Response Surface Methodology. Combust. Sci. Technol. 2019, 192, 1–18. [Google Scholar] [CrossRef]
  28. Zhang, X.; Zhang, H.; Song, X.; Han, X.; Bao, J.; Zhang, N.; He, G. Co3O4-based catalysts derived from natural wood with hierarchical structure for elemental mercury oxidation. J. Energy Inst. 2021, 94, 285–293. [Google Scholar] [CrossRef]
  29. Zhang, X.P.; Song, X.N.; Wang, X.X.; Wei, Y.Y.; Han, X.K.; Bao, J.J.; Zhang, N.; He, G.H. In-situ grown Co3O4 nanoparticles on wood-derived carbon with natural ordered pore structure for efficient removal of Hg0 from flue gas. J. Energy Inst. 2021, 98, 206–215. [Google Scholar] [CrossRef]
  30. Xu, Y.L.; Zhong, Q.; Xing, L.L. Gas-phase elemental mercury removal from flue gas by cobalt-modified fly ash at low temperatures. Environ. Technol. 2014, 35, 2870–2877. [Google Scholar] [CrossRef]
  31. Yang, W.; Li, C.; Wang, H.; Li, X.; Zhang, W.; Li, H. Cobalt doped ceria for abundant storage of surface active oxygen and efficient elemental mercury oxidation in coal combustion flue gas. Appl. Catal. B-Environ. 2018, 239, 233–244. [Google Scholar] [CrossRef]
  32. Xu, W.; Tong, L.; Qi, H.; Zhou, X.; Wang, J.; Zhu, T. Effect of Flue Gas Components on Hg0 Oxidation over Fe/HZSM-5 Catalyst. Ind. Eng. Chem. 2015, 54, 146–152. [Google Scholar] [CrossRef]
  33. Yang, J.; Zhao, Y.; Ma, S.; Zhu, B.; Zhang, J.; Zheng, C. Mercury Removal by Magnetic Biochar Derived from Simultaneous Activation and Magnetization of Sawdust. Environ. Sci. Technol. 2016, 50, 12040–12047. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, W.; Hussain, A.; Zhang, J.; Liu, Y.X. Removal of elemental mercury from flue gas using red mud impregnated by KBr and KI reagent. Chem. Eng. J. 2018, 341, 483–494. [Google Scholar] [CrossRef]
  35. Liu, M.; Li, C.; Zeng, Q.; Du, X.; Gao, L.; Li, S.; Zhai, Y. Study on removal of elemental mercury over MoO3-CeO2/cylindrical activated coke in the presence of SO2 by Hg-temperature-programmed desorption. Chem. Eng. J. 2019, 371, 666–678. [Google Scholar] [CrossRef]
  36. Liu, D.; Zhou, W.; Wu, J. CeO2–MnOx/ZSM-5 sorbents for H2S removal at high temperature. Chem. Eng. J. 2016, 284, 862–871. [Google Scholar] [CrossRef]
  37. Shen, B.; Liu, Z.; Xu, H.; Wang, F. Enhancing the absorption of elemental mercury using hydrogen peroxide modified bamboo carbons. Fuel 2019, 235, 878–885. [Google Scholar] [CrossRef]
  38. Peng, C.; Yan, R.; Peng, H.; Mi, Y.; Liang, J.; Liu, W.; Wang, X.; Song, G.; Wu, P.; Liu, F. One-pot synthesis of layered mesoporous ZSM-5 plus Cu ion-exchange: Enhanced NH3-SCR performance on Cu-ZSM-5 with hierarchical pore structures. J. Hazard. Mater. 2020, 385, 121593. [Google Scholar] [CrossRef]
  39. An, D.; Wang, X.; Cheng, X.; Cui, L.; Zhang, X.; Zhou, P.; Dong, Y. Regeneration performance of activated coke for elemental mercury removal by microwave and thermal methods. Fuel Process. Technol. 2020, 199, 106303. [Google Scholar] [CrossRef]
  40. Wang, Y.; Li, C.; Zhao, L.; Xie, Y.E.; Zhang, X.; Zeng, G.; Wu, H.; Zhang, J. Study on the removal of elemental mercury from simulated flue gas by Fe2O3-CeO2/AC at low temperature. Environ. Sci. Pollut. Res. 2016, 23, 5099–5110. [Google Scholar] [CrossRef]
  41. Zhang, X.; Wang, J.; Tan, B.; Zhang, N.; Bao, J.; He, G. Ce-Co interaction effects on the catalytic performance of uniform mesoporous Cex-Coy catalysts in Hg0 oxidation process. Fuel 2018, 226, 18–26. [Google Scholar] [CrossRef]
  42. Gao, L.; Li, C.; Zhang, J.; Du, X.; Li, S.; Zeng, J.; Yi, Y.; Zeng, G. Simultaneous removal of NO and Hg0 from simulated flue gas over CoOx-CeO2 loaded biomass activated carbon derived from maize straw at low temperatures. Chem. Eng. J. 2018, 342, 339–349. [Google Scholar] [CrossRef]
  43. Cai, X.; Nie, J.; Yang, G.; Wang, F.; Ma, C.; Lu, C.; Chen, Z. Phosphorus-rich network polymer supported ruthenium nanoparticles for nitroarene reduction. Mater. Lett. 2019, 240, 80–83. [Google Scholar] [CrossRef]
  44. Xia, Y.; Wei, G.; Liang, X.; Zhu, J.; Xian, H.; Su, X.; He, H.; Zhu, R. Sequestration of Gaseous Hg0 by Sphalerite with Fe Substitution: Performance, Mechanism, and Structure–Activity Relationship. J. Phys. Chem. C 2019, 123, 2828–2836. [Google Scholar] [CrossRef]
  45. Gao, L.; Li, C.; Li, S.; Zhang, W.; Du, X.; Huang, L.; Zhu, Y.; Zhai, Y.; Zeng, G. Superior performance and resistance to SO2 and H2O over CoOx-modified MnOx/biomass activated carbons for simultaneous Hg0 and NO removal. Chem. Eng. J. 2019, 371, 781–795. [Google Scholar] [CrossRef]
  46. Tan, P. Active phase, catalytic activity, and induction period of Fe/zeolite material in nonoxidative aromatization of methane. J. Catal. 2016, 338, 21–29. [Google Scholar] [CrossRef]
  47. Wu, H.; Li, C.; Zhao, L.; Zhang, J.; Zeng, G.; Xie, Y.E.; Zhang, X.; Wang, Y. Removal of Gaseous Elemental Mercury by Cylindrical Activated Coke Loaded with CoOx-CeO2 from Simulated Coal Combustion Flue Gas. Energy Fuel 2015, 29, 6747–6757. [Google Scholar] [CrossRef]
  48. Cao, T.; Zhou, Z.; Chen, Q.; Li, Z.; Xu, S.; Wang, J.; Xu, M.; Bisson, T.; Xu, Z. Magnetically responsive catalytic sorbent for removal of Hg0 and NO. Fuel Process. Technol. 2017, 160, 158–169. [Google Scholar] [CrossRef]
  49. Yang, Z.; Li, H.; Liu, X.; Li, P.; Yang, J.; Lee, P.-H.; Shih, K. Promotional effect of CuO loading on the catalytic activity and SO2 resistance of MnOx/TiO2 catalyst for simultaneous NO reduction and Hg0 oxidation. Fuel 2018, 227, 79–88. [Google Scholar] [CrossRef]
  50. Yang, W.; Li, Y.; Shi, S.; Chen, H.; Shan, Y.; Liu, Y. Mercury removal from flue gas by magnetic iron-copper oxide modified porous char derived from biomass materials. Fuel 2019, 256, 115977. [Google Scholar] [CrossRef]
  51. Xu, Y.; Luo, G.; Zhang, Q.; Li, Z.; Zhang, S.; Cui, W. Cost-effective sulfurized sorbents derived from one-step pyrolysis of wood and scrap tire for elemental mercury removal from flue gas. Fuel 2021, 285, 119221. [Google Scholar] [CrossRef]
  52. Zhou, Z.; Liu, X.; Li, C.; Cao, X.E.; Xu, M. Seawater-assisted synthesis of MnCe/zeolite-13X for removing elemental mercury from coal-fired flue gas. Fuel 2020, 262, 116605. [Google Scholar] [CrossRef]
  53. Yang, L.; Bing, L.; He, Y.; YANG, D.-w.; HU, H.-q. Removal of elemental mercury (Hg0) from simulated flue gas over MnOx-TiO2 sorbents. J. Fuel. Chem. Technol. 2020, 48, 513–524. [Google Scholar]
  54. Presto, A.A.; Granite, E.J. Noble Metal Catalysts for Mercury Oxidation in Utility Flue Gas GOLD, PALLADIUM AND PLATINUM FORMULATIONS. Platin. Met. Rev. 2008, 52, 144–154. [Google Scholar] [CrossRef]
  55. Wang, Z.; Liu, J.; Yang, Y.; Yu, Y.; Yan, X.; Zhang, Z. AMn2O4 (A=Cu, Ni and Zn) sorbents coupling high adsorption and regeneration performance for elemental mercury removal from syngas. J. Hazard. Mater. 2020, 388, 121738. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the experimental setup.
Figure 1. Schematic diagram of the experimental setup.
Applsci 12 03769 g001
Figure 2. (a) Removal of mercury over various samples; (b) calculation of adsorption capacity for the pseudo-first-order model and pseudo-second-order model.
Figure 2. (a) Removal of mercury over various samples; (b) calculation of adsorption capacity for the pseudo-first-order model and pseudo-second-order model.
Applsci 12 03769 g002
Figure 3. FTIR spectra of a series of samples.
Figure 3. FTIR spectra of a series of samples.
Applsci 12 03769 g003
Figure 4. XRD patterns of a series of samples.
Figure 4. XRD patterns of a series of samples.
Applsci 12 03769 g004
Figure 5. (a) N2 adsorption–desorption isotherms and (b) pore distributions of a series samples.
Figure 5. (a) N2 adsorption–desorption isotherms and (b) pore distributions of a series samples.
Applsci 12 03769 g005
Figure 6. SEM images of (a,b) ZSM-5 and (c,d) Co4Fe1-ZSM-5.
Figure 6. SEM images of (a,b) ZSM-5 and (c,d) Co4Fe1-ZSM-5.
Applsci 12 03769 g006
Figure 7. (a) SEM image of fresh Co4Fe1-ZSM-5. (b,c) The corresponding EDX elemental mapping images of Co (green) and Fe (red). (d) EDX analysis.
Figure 7. (a) SEM image of fresh Co4Fe1-ZSM-5. (b,c) The corresponding EDX elemental mapping images of Co (green) and Fe (red). (d) EDX analysis.
Applsci 12 03769 g007
Figure 8. H2-TPR patterns of a series of samples.
Figure 8. H2-TPR patterns of a series of samples.
Applsci 12 03769 g008
Figure 9. XPS spectra of fresh Co-ZSM-5, Fe-ZSM-5, and fresh Co4Fe1-ZSM-5: (a) Co 2p, (b) Fe 2p.
Figure 9. XPS spectra of fresh Co-ZSM-5, Fe-ZSM-5, and fresh Co4Fe1-ZSM-5: (a) Co 2p, (b) Fe 2p.
Applsci 12 03769 g009
Figure 10. Effect of gas components on Hg0 removal efficiency. T = 120 °C.
Figure 10. Effect of gas components on Hg0 removal efficiency. T = 120 °C.
Applsci 12 03769 g010
Figure 11. (a) Hg-TPD curves on sorbent at 3 °C/min. (b) Effect of regeneration temperature on Hg0 removal performance. (c) Effect of regeneration atmosphere on Hg0 removal performance. (d) Hg0 removal efficiency of the sample over five repeated adsorption-regeneration; condition: N2 + 5% O2 + 200 μg/m3 Hg0.
Figure 11. (a) Hg-TPD curves on sorbent at 3 °C/min. (b) Effect of regeneration temperature on Hg0 removal performance. (c) Effect of regeneration atmosphere on Hg0 removal performance. (d) Hg0 removal efficiency of the sample over five repeated adsorption-regeneration; condition: N2 + 5% O2 + 200 μg/m3 Hg0.
Applsci 12 03769 g011
Figure 12. XPS spectra of fresh and spent Co4Fe1-ZSM-5: (a) Co 2p, (b) Fe 2p, (c) O1s, (d) Hg 4f.
Figure 12. XPS spectra of fresh and spent Co4Fe1-ZSM-5: (a) Co 2p, (b) Fe 2p, (c) O1s, (d) Hg 4f.
Applsci 12 03769 g012
Table 1. N2 adsorption results of a series samples.
Table 1. N2 adsorption results of a series samples.
SamplesSurface Area/(m2/g)Micropore Surface Area/(m2/g)External Surface Area/(m2/g)Total Pore Volume/(cm3/g)
ZSM-5356289670.18
Co-ZSM-5336260760.16
Fe-ZSM-5343269740.19
Co4Fe1-ZSM-5351264870.18
Table 2. XPS data of a series of samples.
Table 2. XPS data of a series of samples.
SamplesConcentration of the OxygenCo3+/Co2+Fe3+/Fe2+
OαOβ
Fresh Co-ZSM-5--0.77-
Fresh Fe-ZSM-5---1.62
Fresh Co4Fe1-ZSM-57.8%92.2%1.371.29
Spent Co4Fe1-ZSM-55.9%94.1%0.840.93
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, W.; Ye, D.; Wang, H. Experimental Study on the Elemental Mercury Removal Performance and Regeneration Ability of CoOx–FeOx-Modified ZSM-5 Adsorbents. Appl. Sci. 2022, 12, 3769. https://doi.org/10.3390/app12083769

AMA Style

Ma W, Ye D, Wang H. Experimental Study on the Elemental Mercury Removal Performance and Regeneration Ability of CoOx–FeOx-Modified ZSM-5 Adsorbents. Applied Sciences. 2022; 12(8):3769. https://doi.org/10.3390/app12083769

Chicago/Turabian Style

Ma, Wei, Dong Ye, and Haining Wang. 2022. "Experimental Study on the Elemental Mercury Removal Performance and Regeneration Ability of CoOx–FeOx-Modified ZSM-5 Adsorbents" Applied Sciences 12, no. 8: 3769. https://doi.org/10.3390/app12083769

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