Next Article in Journal
Quantitative Comparison of Infrared Thermography, Visual Inspection, and Electrical Analysis Techniques on Photovoltaic Modules: A Case Study
Previous Article in Journal
Can Clean Heating in Winter in Northern China Reduce Air Pollution?—Empirical Analysis Based on the PSM-DID Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Molten-Salt Pyrolysis Synthesis Strategy toward Sulfur-Functionalized Carbon for Elemental Mercury Removal from Coal-Combustion Flue Gas

1
School of Energy Science and Engineering, Central South University, Changsha 410083, China
2
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University, Yinchuan 750021, China
3
Shandong Shiheng Thermal Power Co., Ltd., Taian 271600, China
*
Author to whom correspondence should be addressed.
Energies 2022, 15(5), 1840; https://doi.org/10.3390/en15051840
Submission received: 29 January 2022 / Revised: 25 February 2022 / Accepted: 27 February 2022 / Published: 2 March 2022
(This article belongs to the Topic Energy Efficiency, Environment and Health)

Abstract

:
The emission of mercury from coal combustion has caused consequential hazards to the ecosystem. The key challenge to abating the mercury emission is to explore highly efficient adsorbents. Herein, sulfur-functionalized carbon (S-C) was synthesized by using a molten-salt pyrolysis strategy and employed for the removal of elemental mercury from coal-combustion flue gas. An ideal pore structure, which was favorable for the internal diffusion of the Hg0 molecule in carbon, was obtained by using a SiO2 hard template and adjusting the HF etching time. The as-prepared S-C with an HF etching time of 10 h possessed a saturation Hg0 adsorption capacity of 89.90 mg·g−1, far exceeding that of the commercial sulfur-loaded activated carbons (S/C). The S-C can be applied at a wide temperature range of 25–125 °C, far exceeding that of commercial S/C. The influence of flue gas components, such as SO2, NO, and H2O, on the Hg0 adsorption performance of S-C was insignificant, indicating a good applicability in real-world applications. The mechanism of the Hg0 removal by S-C was proposed, i.e., the reduced components, including sulfur thiophene, sulfoxide, and C-S, displayed a high affinity toward Hg0, which could guarantee the stable immobilization of Hg0 as HgS in the adsorbent. Thus, the molten-salt pyrolysis strategy has a broad prospect in the application of one-pot carbonization and functionalization sulfur-containing organic precursors as efficient adsorbents for Hg0.

1. Introduction

The excessive emission of mercury from industrial activities has caused consequential hazards to the ecosystem and human health [1,2,3]. Coal combustion is one of the largest industrial sources of mercury emission. The mercury emitted from typical coal-combustion flue gas generally existed in three forms, i.e., elemental mercury (Hg0), oxidized mercury (Hg2+), and particulate bound mercury (Hgp) [4,5,6,7,8]. The Hgp can be captured by using particulate matter control devices, while the Hg2+ can be removed by using wet flue-gas scrubbers due to Hg2+’s water solubility [9,10,11,12]. However, Hg0 is difficult to remove owing to its water insolubility and high volatility [13,14,15]. As a consequence, Hg0 is the primary species of mercury discharged into the atmosphere from coal-combustion flue gases. The highly efficient removal of Hg0 is a key challenge to reducing mercury pollution from coal-fired power plants.
Activated carbons (ACs) are the most widely researched and skilled mercury adsorbents for coal-fired power plants [16,17,18,19,20,21]. However, ACs were generally limited by adsorption kinetics and equilibrium capacities, hence causing large consumptions of ACs during Hg0 removal [22,23]. Moreover, the weak binding affinity of ACs toward mercury induced leaching risks of mercury when the ACs were dumped in landfills together with fly ashes [24,25]. Thus, active components that can accommodate Hg0 were generally introduced to the ACs to improve the Hg0 adsorption capacity. In nature, the mercury was tied to sulfides owing to its sulphophile affinity [26,27]. Inspired by this natural law, sulfur was widely employed to modify the Hg0 adsorption capacity of ACs [28,29]. This strategy was relatively simple, cheap, and effective, but was still found to suffer from drawbacks [30,31]. The most significant one is that the sulfur was not actually anchored to the carbon surface during the impregnation process, hence affecting the activity, abundance, and accessibility of the sulfur groups during Hg0 removals [32]. The impregnation method for introducing sulfur would plug the pores of ACs. As a result, the efficient diffusion of Hg0 on the sulfur-modified ACs would be limited [33]. In addition, the anchoring of sulfur on ACs was not firm, hence causing leaching risks of sulfur as well as mercury adsorption products when ACs were disposed of in landfills [32,34]. Therefore, a new strategy is urgently undergoing exploration to overcome the disadvantages associated with sulfur-impregnated ACs [35,36].
Traditionally, the sulfur-functionalized carbon materials were prepared by using two steps, i.e., the carbonization of natural products (e.g., cellulose, chitin, starch, alginic acid, and chitosan) as well as some synthetic polymers (e.g., poly-acrylonitrile, polyaniline, and phenolic resins) first and then introducing sulfur onto carbons by using an impregnation method [37,38,39,40]. A one-pot carbonization and functionalization step, which could guarantee the uniform distribution and firm fixation of sulfur functionality on the carbon matrix, was required. Very recently, a simple approach via the carbonization of small organic molecules with the assistance of transition metals was reported to prepare a series of functional carbon materials [37]. This approach was realized via the pyrolysis of a mixture of small organic molecules and transition-metal salts in a conventional tubular furnace, hence avoiding the barriers to large-scale production, such as complicated equipment and harsh conditions. The salts would act as a heat-transfer medium and provide an oxygen-free environment for pyrolysis [41]. Moreover, the salts catalyze the formation of a thermally stable intermediate polymerization structure, avoiding the direct sublimation of small organic molecules during heating [37,38]. More attractive, template-like SiO2 could be mixed into the precursors and hence adopted to adjust the porosity of carbons. This strategy is commonly available and easy to use to control the surface sulfur functionalities, porosities, and morphologies of carbons.
In this work, porous sulfur-functionalized carbons were prepared and employed for Hg0 removal from coal-fired flue gas. The removal performance of sulfur-functionalized carbon (S-C) on Hg0 was studied and compared with that of commercial sulfur-loaded ACs (S/C). The Hg0 removal performances of S-C and S/C under various adsorption temperatures and flue gas conditions were studied. The excellent adsorption mechanism of Hg0 by S-C was further investigated. This work not only provides a promising trap for highly efficient Hg0 sequestration from coal-fired power plants but also illustrates a versatile platform for preparing functional carbon materials by using a one-pot carbonization and functionalization organic precursor.

2. Experimental Section

2.1. Sample Preparation

S-C was prepared by using a molten-salt pyrolysis strategy [38]. Two grams of 2,2-bithiophene and 2.0 g of SiO2 were added into 150 mL of tetrahydrofuran and stirred at room temperature for 6 h. The mixture was then dried and then ground to a powder. After that, the powder was placed into a rail boat and covered with 5.0 g of Co(NO3)2·6H2O, which was heated at 800 °C for 2 h. Then, the SiO2 was removed by using 10% HF solution. To obtain different porosity of S-C, the sample was etched by using HF for different times (0, 24, 48 and 56 h). After drying at 105 °C for 10 h, the S-C was finally obtained.

2.2. Sample Characterization

The morphology of the sample was studied by using a scanning electronic microscope (SEM, FEI F50, New York, NY, USA). Transmission electron microscope (TEM, EOL JEM 2100F, microscope Tokyo, Japan) and high-resolution TEM (HRTEM) were used to study the morphology and structure of sample. The valence states of samples were characterized by using X-ray photoelectron spectroscopy (XPS, Thermo ESCALAB 250Xi, New York, NY, USA).

2.3. Hg0 Adsorption Activity Tests

The Hg0 adsorption activity test was measured on a fixed-bed reactor [42,43]. The adsorbent was placed in a quartz reactor, reaction temperature of which was controlled by using a tube furnace. The flue gas was composed of N2, O2, SO2, NO, and H2O, with the total flow rate of 1 L·min−1. Hg0 was provided by a mercury permeation tube placed in a constant temperature water bath, delivering Hg0 by using N2 to ensure a stable concentration of 65 µg·m−3. The concentration of Hg0 was monitored by using an online mercury analyzer (RA-915M, Lumex, Tianjin, China). The Hg0 adsorption capability was calculated by using the following equations:
Q = 1 m t 1 t 2 ( C i n C o u t ) × f × d t
where Cin (μg·m−3) and Cout (μg·m−3) represent the inlet and outlet concentrations of Hg0, Q (μg·g−1) is the Hg0 adsorption capacity, m (g) is the sample amount, f (m3·h−1) is the gas flow rate, and t (h) is the reaction time.

3. Description of Sorption Kinetic Models

3.1. Pseudo-First-Order Model

This model is described as follows [44]:
d q t d t = k 1 ( q e q t )
Based on the initial conditions, i.e., t = 0 qt = 0, Equation (2) is revised as:
q t = q e ( 1 e k 1 t )
where qt and qe represent the adsorbed mercury amount at any time t and equilibrium time (µg·g−1). k1 represents the rate constant (min−1).

3.2. Pseudo-Second-Order Model

This model is described as follows [45]:
d q t d t = k 2 ( q e q t ) 2
On the basis of initial conditions, i.e., t = 0 and qt = 0, Equation (4) is modified as:
q t = t 1 k 2 q e 2 + 1 q e t
where k2 represents the rate constant (µg/(cm3·min)).

3.3. Elovich Model

This model is described as follows [46]:
d q t d t = α exp ( β q t )
where α represents the initial rate and β is related to surface coverage and activation energy. If t is much larger than t0, this equation is modified as follows:
q t = 1 β ln ( α β ) + ( 1 β ) ln t

3.4. Intra-Particle Diffusion Model

This model is described by the following formula [47]:
q t = k i d t 0.5 + C
where kid represents the diffusion rate constant within the particle and C is a constant, which is related to the boundary layer.

4. Results and Discussion

4.1. Preparation and Characterization of Samples

The S-C was prepared by using the molten Co(NO3)2·6H2O-assisted carbonization of sulfur-containing small organic molecule precursors. SiO2 nanoparticles were adopted as hard templates, followed by an HF etching step to remove the SiO2 template. The porosity of S-C was adjusted by using the HF etching time (shown in Figure 1). As listed in Table 1, the HF etching step will generate mesopores on the S-C. The S-C with HF etching for 10 h possessed the largest BET surface area of 318.5 m2·g−1, with an average pore size of 7.31 nm and total pore volume of 0.58 cm3·g−1. Pore size affects the internal mass transfer process of Hg0 on adsorbents, thus affecting the removal rate of Hg0. As shown in Figure 2a, the N2 absorption–desorption isotherms for S-C belonged to a typical II isotherm, indicating that the porosity of S-C was very limited. The S-C after the HF etching step presented a typical IV isotherm with a negligible absorption at lower pressures but significant absorption at higher pressures (p/p0 = 0.2-1.0). This indicates that S-C is composed of mesopores or macropores rather than micropores. The pore distribution curves shown in Figure 2b demonstrate that the pore size of the S-C was in the range of 2–10 nm. The mesoporous structure in the range of 2–50 nm is beneficial for Hg0’s diffusion to the inner surface of an adsorbent [48]. Thus, the pore structure of S-C is favorable for Hg0 adsorption.
Figure 3 shows the morphologies of S-Cs with different HF etching times. These images display the high-solution local microscopies of individual particles, in which the sizes of individual particles are in the range of 100–200 mesh. As shown, the morphology of S-C was significant dependent on the HF etching time. The S-C without HF etching displayed a dense surface, and there were negligible pores observed on the S-C surface. After etching with HF, abundant pores were generated on the S-C surface, and more pores were generated with the extension of the HF etching time from 5 to 10 h. Thus, the SiO2 was embedded in the carbon structure as a template, which could be removed by using the HF solution to induce the formation of a porous structure. However, upon reaching a higher HF etching time, the resultant carbon framework collapsed, which might have affected the internal diffusion of mercury. Therefore, a SiO2 template was used to achieve the adjustable preparation of porous carbons.
The surface-functional groups of S-C were studied by using FTIR. As shown in Figure 4, there were four main peaks on the FTIR spectra, which could be assigned to the C-OH (3392 and 615 cm−1) and aromatic C = C (1626 and 1100 cm−1) [49]. Figure 5 shows the XPS spectra for S-C. As shown in Figure 5a, the spectra for high-solution C 1s spectra could be divided into three peaks at 284.8, 286.0, and 288.8 eV, assigned to the characteristic peaks of C = C, C-O, and C = O [30]. Figure 5b shows the O 1s spectra for C-S, in which the peaks at 531.5, 532.3, and 533.3 eV are regarded as C-OH, C-O, and O-H [31]. Figure 5c shows the S spectra for S-C. The sulfur on the S-C existed in four forms. The two peaks at 164.2 and 165.3 eV corresponded to S 2p3/2 and S 2p1/2 for thiophenic sulfur (i.e., -C-Sx-C-, x = 1-2), the small peak at 169.1 eV could be assigned to sulfoxide [50], the peak at 166.2 eV could be assigned to C-S, while the peaks at 167.2 and 170.3 eV could be assigned to sulfate [51,52].

4.2. Hg0 Adsorption Capacity Tests

Figure 6 shows the Hg0 adsorption performances of S-Cs with various HF etching times. As shown, the normalized-outlet Hg0 concentration of S-C without etching climbed dramatically to above 0.90. Thus, the S-C without etching has a poor Hg0 adsorption performance, and the removal rate of Hg0 is only below 10%. However, after the HF etching, the S-C displayed far superior Hg0 removal performances. The HF etching times played significant roles in the Hg0 adsorption of S-C. After etching for 5 h, the concentration of Hg0 at the normalized outlet remained below 0.2. With the extension of the etching time to 10 h, the Hg0 adsorption performance of S-C can be further improved. However, a too-long etching time resulted in a decrease in Hg0 adsorption performance. This is attributed to the fact that HF etching can remove the SiO2 template to generate abundant pores on S-C, which allows for accessible mercury molecule transportation. As a result, the sulfur in S-C can be accessible sufficiently for binding mercury. However, the excessive etching will result in the collapse of the framework of carbons, hence the weakening of the Hg0 adsorption on S-C.
The saturated adsorption capacity of an adsorbent is an important index to use to evaluate the adsorption performances of materials, so it is necessary to study the change in S-C adsorption capacity with time. The curve of Hg0 adsorption capacity changing over time is shown in Figure 7. When the reaction duration was 1200 min, the adsorption capacity of Hg0 exceeded 50% and reached 45.39 mg·g−1. In addition, it can be seen from the figure that the adsorption rate increases first and then decreases with time, and the slope is zero when the adsorption is saturated. Adsorption kinetic models can be adopted to investigate the process of Hg0 adsorption and the dominant controlling factors. The pseudo-first-order-based kinetic model is the most common adsorption kinetic model. Ln(qe-qt) is used to plot t, and the adsorption mechanism conforms to the pseudo-first-order model if a straight line can be obtained. The pseudo-second-order model is based on the assumption that the adsorption rate is controlled by the chemisorption mechanism. Elovich describes a series of reaction mechanism processes, which are suitable for processes with large activation energy changes in reactions and for complex heterogeneous diffusion processes. The intra-particle diffusion model is commonly used to analyze the control steps in the reaction. Generally, the material adsorption process is divided into two processes: adsorbent surface adsorption and slow pore diffusion. If the fitting result fails to reach the origin, it indicates that the internal diffusion of the material is not the only step to controlling the adsorption process. As shown in Figure 8, the pseudo-first-order model was closest to the adsorption process of Hg0 on S-C, with an extremely high correlation coefficient (R2 = 0.9982). The saturation adsorption capacity of S-C was simulated as 89.90 mg·g−1. When the molar ratio of Hg:S is 1:1, it is equivalent to 55% of the sulfur accessibility in Hg0 adsorption. For comparison, the Hg0 saturation adsorption capacity of a sulfur-loaded commercial activated carbon (S/C) was also investigated. It should be noted that the S/C possessed a higher sulfur content and surface area compared with S-C. However, the saturation adsorption capacity of S/C was less than 1 mg·g−1, which was much lower compared with its theoretical adsorption capacity of 352.1 mg·g−1. Thus, most of the sulfur in S/C was not adopted sufficiently for binding mercury, although it possessed a higher surface area. It fully shows that the key to Hg0 adsorption improvements lies in the high dispersion of active sulfur species [53]. Thus, the molten-salt pyrolysis synthesis strategy toward sulfur-functionalized carbon would be more superior compared with other traditional methods, such as impregnation when significantly dispersing the active components (i.e., sulfur).

4.3. Impact of Operation Conditions on Hg0 Adsorption Capacity

An excellent adsorbent should have a good stability at various conditions. Figure 9 shows the influence of temperature on the Hg0 adsorption capacities of S-C and S/C. The normalized-outlet Hg0 concentration when passing through the S-C was maintained below 0.05 in a wide reaction temperature range of 25–125 °C. This suggests that an above 95% Hg0 adsorption efficiency was obtained; even the sorbent dosage was as low as 5 mg. This wide temperature range proves that the S-C can be applied flexibly at different scenes for Hg0 removal. In contrast, the S/C exhibited significantly different Hg0 adsorption capacities under various temperatures, in which relatively limited Hg0 adsorption capacities were obtained at low temperatures. Even at the optimum reaction temperature and the same amount of adsorbent, the concentration of Hg0 at the normalized outlet is still higher than 0.15, much higher than S-C. Depending on the preferred temperature range, setting the S/C upstream of the ESP is the ideal application in which the Hg0 adsorption capacity of S-C might be affected by various flue gas components, such as high concentrations of fly ash, SO2, and NO2.
According to previous studies, SO2 and H2O generally compete with Hg0 for adsorption sites, hence exhibiting inhibitive effects on the Hg0 removal over carbonaceous sorbents [54,55]. However, the S-C exhibited an excellent resistance to these detrimental flue gas components. Figure 10a–c show that, in the presence of SO2, H2O, as well as NO, the Hg0 removal performance was very similar to that under a pure N2 atmosphere. Even adding 1200 ppm of SO2, 12% H2O, or 400 ppm of NO, the normalized-outlet Hg0 concentration was kept at less than 0.05. As a direct comparison, the adsorption performance of S/C for Hg0 was investigated. As shown in Figure 10d, the SO2, H2O, and NO significantly weakened the adsorption capacity of S/C. These results could fully demonstrate the excellent resistance of S-C to the detrimental impacts of flue gas impurities compared with commercial activated carbon, which would facilitate the real-world applications.

4.4. Reaction Mechanism

To study the mechanism of a Hg0 removal on S-C, the valence states of elements on spent S-C after adsorbing mercury were investigated by using XPS. Figure 11a shows that there is no significant change in the spectral binding energy of C 1s for the spent adsorbent, indicating that C did not participate in the Hg0 adsorption [56]. Figure 11b shows the O 1s spectra for the spent adsorbent. The spectra of O 1s can be divided into four separate corresponding peaks: C = O, C-O-C, COOH, and H2O at 531.3, 532.5, 533.8, and 536.1 eV, respectively [57]. After the Hg0 adsorption, part of the C-O groups changed to C = O, and H2O was generated on the adsorbent surface, which may have been caused by a charge imbalance after binding Hg0. Figure 11c shows the S 2p for the spent adsorbent, including only two peaks: C-S at 163.94 eV and thiophene at 165.22 eV [58]. The transformations of sulfur species on the spent adsorbent indicate that sulfur played a crucial role in the Hg0 removal. The absolute content of thiophene and C-S decreased after adsorbing Hg0, especially the peaks for sulfoxide, which disappeared compared with the fresh adsorbent. This variation indicates that thiophene (i.e., -C-Sx-C-, x = 1-2), sulfoxide, and C-S were beneficial to improving the Hg0 removal capacity. This is in line with a previous study [59], i.e., findings that the sulfur in sulfide existing in a low state as well as the reducing sulfur (such as thiosulfone) had a high affinity with the Hg0 atom. Electrons around the sulfur atom were easy to combine with Hg0, which was conducive to the removal of mercury. Figure 11d shows that the two peaks of Hg 4f corresponded to the peaks of 103 and 107 eV of HgS [60], further demonstrating that the sulfur in the adsorbent was active for binding Hg0.

5. Conclusions

Sulfur-functionalized carbon (S-C) derived from a molten-salt pyrolysis strategy was employed for a Hg0 removal. Abundant pores could be generated by using a SiO2 hard template and a subsequent HF etching step. The HF etching time significantly affected the Hg0 removal performances of S-Cs. With an HF etching time of 10 h, the S-C presented the best Hg0 removal performance, the saturation Hg0-adoption capacity of which reached 89.90 mg·g−1. This value far exceeded that of the commercial sulfur-loaded activated carbon (S/C), which was specialized for Hg0 removals. The S-C displayed a good applicability at 25–125 °C, while the S/C could be adopted around only 125 °C since the Hg0 adsorption performances decreased when deviating from this temperature. The SO2, NO, and H2O had negligible adverse effects on Hg0 adsorption over S-C. The good Hg0 removal performance of S-C was ascribed to the existence of reduced sulfurs, such as thiophene, sulfoxide, and C-S, which have high affinities toward Hg0, and the fact that gaseous Hg0 was converted into HgS after the adsorption. These results indicate that molten-salt pyrolysis can simultaneously achieve the carbonization and functionalization of sulfur-containing organic precursors and is an ideal method for preparing carbonaceous adsorbents for Hg0 removal.

Author Contributions

Conceptualization, J.Y.; Writing—original draft, J.Y.; Formal analysis, J.Y.; Investigation, H.X.; Writing—review & editing, H.X.; Software, F.M. and W.Q.; Methodology, F.M.; Supervision, Q.G.; Validation, Q.G., T.H. and Z.Y.; Data curation, T.H. and W.Q.; Visualization, Z.Y.; Resources, H.L.; Project administration, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No data was reported.

Acknowledgments

This project was supported by the National Natural Science Foundation of China (51906260), the Natural Science Foundation of Hunan Province, China (2021JJ30851), the Postgraduate Scientific Research Innovation Project of Hunan Province (CX20210098), the Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (Grant No. 2022-K09 and 2022-K52), and the Key Research and Development Program of Sichuan Province (2021YFG0117).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Landrigan, P.J.; Wright, R.O.; Birnbaum, L.S. Mercury toxicity in children. Science 2013, 342, 1447. [Google Scholar] [CrossRef] [PubMed]
  2. Richardson, J.B.; Moore, L. A tale of three cities: Mercury in urban deciduous foliage and soils across land-uses in Poughkeepsie NY, Hartford CT, and Springfield MA USA. Sci. Total Environ. 2020, 715, 136869. [Google Scholar] [CrossRef] [PubMed]
  3. Li, M.; Yang, H.E.; Lu, H.; Wu, T.; Zhou, D.; Liu, Y. Investigation into the Classification of Tight Sandstone Reservoirs via Imbibition Characteristics. Energies 2018, 11, 2619. [Google Scholar] [CrossRef] [Green Version]
  4. Masoomi, I.; Kamata, H.; Yukimura, A.; Ohtsubo, K.; Schmid, M.O.; Scheffknecht, G. Investigation on the behavior of mercury across the flue gas treatment of coal combustion power plants using a lab-scale firing system. Fuel Process. Technol. 2020, 201, 106340. [Google Scholar] [CrossRef]
  5. Lee, S.S.; Wilcox, J. Behavior of mercury emitted from the combustion of coal and dried sewage sludge: The effect of unburned carbon, Cl, Cu and Fe. Fuel 2017, 203, 749–756. [Google Scholar] [CrossRef]
  6. Yang, S.; Liu, C.; Wang, P.; Yi, H.; Shen, F.; Liu, H. Co9S8 nanoparticles-embedded porous carbon: A highly efficient sorbent for mercury capture from nonferrous smelting flue gas. J. Hazard. Mater. 2021, 412, 124970. [Google Scholar] [CrossRef]
  7. Fernández-Miranda, N.; Rodríguez, E.; Lopez-Anton, M.; García, R.; Martínez-Tarazona, M. A New Approach for Retaining Mercury in Energy Generation Processes: Regenerable Carbonaceous Sorbents. Energies 2017, 10, 1311. [Google Scholar] [CrossRef] [Green Version]
  8. Mei, J.; Liao, Y.; Qin, R.; Sun, P.; Wang, C.; Ma, Y.; Qu, Z.; Yan, N.; Yang, S. Acceleration of Hg0 adsorption onto natural sphalerite by Cu2+ activation during flotation: Mechanism and applications in Hg0 recovery. Environ. Sci. Technol. 2020, 54, 7687–7696. [Google Scholar] [CrossRef]
  9. Yu, M.Y.; Luo, G.Q.; Sun, R.Z.; Zou, R.J.; Li, X.; Yao, H. A mechanism study on effects of bromide ion on mercury re-emission in WFGD slurry. Chem. Eng. J. 2021, 406, 127010. [Google Scholar] [CrossRef]
  10. Zhang, Y.; Yang, J.; Yu, X.; Sun, P.; Zhao, Y.; Zhang, J.; Chen, G.; Yao, H.; Zheng, C. Migration and emission characteristics of Hg in coal-fired power plant of China with ultra low emission air pollution control devices. Fuel Process. Technol. 2017, 158, 272–280. [Google Scholar] [CrossRef]
  11. Wen, M.; Wu, Q.; Li, G.; Wang, S.; Li, Z.; Tang, Y.; Xu, L.; Liu, T. Impact of ultra-low emission technology retrofit on the mercury emissions and cross-media transfer in coal-fired power plants. J. Hazard. Mater. 2020, 396, 122729. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, X.; Cui, L.; Li, Y.; Zhao, Y.; Dong, Y.; Cao, S. Adsorption and oxidation of mercury by montmorillonite powder modified with different copper compounds. Energy Fuels 2019, 33, 7852–7860. [Google Scholar] [CrossRef]
  13. Heidel, B.; Klein, B. Reemission of elemental mercury and mercury halides in wet flue gas desulfurization. Int. J. Coal. Geol. 2017, 170, 28–34. [Google Scholar] [CrossRef]
  14. Jia, T.; Ji, Z.; Wu, J.; Zhao, X.; Wang, F.; Xiao, Y.; Qi, X.; He, P.; Li, F. Nanosized ZnIn2S4 supported on facet-engineered CeO2 nanorods for efficient gaseous elemental mercury immobilization. J. Hazard. Mater. 2021, 419, 126436. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, H.; Qu, Z.; Zong, C.; Quan, F.; Mei, J.; Yan, N. Catalytic oxidation and adsorption of Hg0 over low-temperature NH3-SCR LaMnO3 perovskite oxide from flue gas. Appl. Catal. B 2016, 186, 30–40. [Google Scholar] [CrossRef]
  16. Choi, S.; Lee, S.S. Mercury adsorption characteristics of Cl-impregnated activated carbons in simulated flue gases. Fuel 2021, 299, 120822. [Google Scholar] [CrossRef]
  17. Dastgheib, S.A.; Mock, J.; Ilangovan, T. Evaluation of modified activated carbons for mercury reemission control during neutralization of a simulated wastewater from the direct contact cooler of a pressurized oxy-combustion process. Ind. Eng. Chem. Res. 2021, 60, 947–954. [Google Scholar] [CrossRef]
  18. Huang, T.F.; Duan, Y.F.; Luo, Z.K.; Zhao, S.L.; Geng, X.Z.; Xu, Y.F.; Huang, Y.J.; Wei, H.Q.; Ren, S.J.; Wang, H.; et al. Influence of flue gas conditions on mercury removal by activated carbon injection in a pilot-scale circulating fluidized bed combustion system. Ind. Eng. Chem. Res. 2019, 58, 15553–15561. [Google Scholar] [CrossRef]
  19. Zhao, W.M.; Geng, X.Z.; Lu, J.C.; Duan, Y.F.; Liu, S.; Hu, P.; Xu, Y.F.; Huang, Y.J.; Tao, J.; Gu, X.B. Mercury removal performance of brominated biomass activated carbon injection in simulated and coal-fired flue gas. Fuel 2021, 285, 119131. [Google Scholar] [CrossRef]
  20. Liu, D.J.; Li, B.; Liu, Y.X.; Wu, J. Copper sulfide-loaded boron nitride nanosheets for elemental mercury removal from simulated flue gas. Energy Fuel 2021, 35, 2234–2242. [Google Scholar] [CrossRef]
  21. Park, J.; Lee, S.S. Adsorption of mercury by activated carbon prepared from dried sewage sludge in simulated flue gas. J. Air Waste Manag. Assoc. 2018, 68, 1077–1084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yang, Z.; Li, H.; Liao, C.; Zhao, J.; Feng, S.; Li, P.; Liu, X.; Yang, J.; Shih, K. Magnetic rattle-type Fe3O4@CuS nanoparticles as recyclable sorbents for mercury capture from coal combustion flue gas. ACS Appl. Nano Mater. 2018, 1, 4726–4736. [Google Scholar] [CrossRef]
  23. Oliveira, L.C.A.; Rios, R.V.R.A.; Fabris, J.D.; Garg, V.; Sapag, K.; Lago, R.M. Activated carbon/iron oxide magnetic composites for the adsorption of contaminants in water. Carbon 2002, 40, 2177–2183. [Google Scholar] [CrossRef]
  24. Padak, B.; Wilcox, J. Understanding mercury binding on activated carbon. Carbon 2009, 47, 2855–2864. [Google Scholar] [CrossRef]
  25. Li, H.; Zhu, L.; Wang, J.; Li, L.; Shih, K. Development of nano-sulfide sorbent for efficient removal of elemental mercury from coal combustion fuel gas. Environ. Sci. Technol. 2016, 50, 9551–9557. [Google Scholar] [CrossRef] [PubMed]
  26. Qiao, G.; Liu, L.; Hao, X.; Zheng, J.; Liu, W.; Gao, J.; Zhang, C.C.; Wang, Q. Signal transduction from small particles: Sulfur nanodots featuring mercury sensing, cell entry mechanism and in vitro tracking performance. Chem. Eng. J. 2020, 382, 122907. [Google Scholar] [CrossRef]
  27. Zhao, L.; Wu, Y.-W.; Han, J.; Wang, H.-X.; Liu, D.-J.; Lu, Q.; Yang, Y.-P. Density Functional Theory Study on Mechanism of Mercury Removal by CeO2 Modified Activated Carbon. Energies 2018, 11, 2872. [Google Scholar] [CrossRef] [Green Version]
  28. Manceau, A.; Merkulova, M.; Murdzek, M.; Batanova, V.; Baran, R.; Glatzel, P.; Saikia, B.K.; Paktunc, D.; Lefticariu, L. Chemical Forms of Mercury in Pyrite: Implications for Predicting Mercury Releases in Acid Mine Drainage Settings. Environ. Sci. Technol. 2018, 52, 10286–10296. [Google Scholar] [CrossRef] [Green Version]
  29. Deng, J.; Li, M.; Wang, Y. Biomass-derived carbon: Synthesis and applications in energy storage and conversion. Green Chem. 2016, 18, 4824–4854. [Google Scholar] [CrossRef]
  30. Wang, W.; Lu, Y.C.; Huang, H.; Wang, A.J.; Chen, J.R.; Feng, J.J. Solvent-free synthesis of sulfur- and nitrogen-co-doped fluorescent carbon nanoparticles from glutathione for highly selective and sensitive detection of mercury(II) ions. Sens. Actuators B Chem. 2014, 202, 741–747. [Google Scholar] [CrossRef]
  31. Ma, J.; Zhou, B.; Zhang, H.; Zhang, W. Fe/S modified sludge-based biochar for tetracycline removal from water. Powd. Technol. 2020, 364, 889–900. [Google Scholar] [CrossRef]
  32. Shin, Y.; Fryxell, G.E.; Um, W.; Parker, K.; Mattigod, S.V.; Skaggs, R. Sulfur-Functionalized Mesoporous Carbon. Adv. Funct. Mater. 2007, 17, 2897–2901. [Google Scholar] [CrossRef]
  33. Marczak-Grzesik, M.; Budzyn, S.; Tora, B.; Szufa, S.; Kogut, K.; Burmistrz, P. Low-cost organic adsorbents for elemental mercury removal from lignite flue gas. Energies 2021, 14, 2174. [Google Scholar] [CrossRef]
  34. Gwenzi, W.; Chaukura, N.; Wenga, T.; Mtisi, M. Biochars as media for air pollution control systems: Contaminant removal, applications and future research directions. Sci. Total Environ. 2021, 753, 142249. [Google Scholar] [CrossRef]
  35. Shen, B.; Li, G.; Wang, F.; Wang, Y.; He, C.; Zhang, M.; Singh, S. Elemental mercury removal by the modified bio-char from medicinal residues. Chem. Eng. J. 2015, 272, 28–37. [Google Scholar] [CrossRef]
  36. Li, G.; Wang, S.; Wang, F.; Wu, Q.; Tang, Y.; Shen, B. Role of inherent active constituents on mercury adsorption capacity of chars from four solid wastes. Chem. Eng. J. 2017, 307, 544–552. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, L.; Chen, M.X.; Yan, Q.Q.; Xu, S.L.; Chu, S.Q.; Chen, P.; Lin, Y.; Liang, H.W. A sulfur-tethering synthesis strategy toward high-loading atomically dispersed noble metal catalysts. Sci. Adv. 2019, 5, 6322. [Google Scholar] [CrossRef] [Green Version]
  38. Wu, Z.Y.; Xu, S.L.; Yan, Q.Q.; Chen, Z.Q.; Ding, Y.W.; Li, C.; Liang, H.W.; Yu, S.H. Transition metal–assisted carbonization of small organic molecules toward functional carbon materials. Sci. Adv. 2018, 4, 0788. [Google Scholar] [CrossRef] [Green Version]
  39. 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]
  40. 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]
  41. Yang, J.; Xu, H.; Zhao, Y.; Li, H.; Zhang, J. Mercury Removal from Flue Gas by Noncarbon Sorbents. Energy Fuel 2021, 35, 3581–3610. [Google Scholar] [CrossRef]
  42. Yang, J.; Li, Q.; Zu, H.; Yang, Z.; Qu, W.; Li, M.; Li, H. Surface-engineered sponge decorated with copper selenide for highly efficient gas-phase mercury Immobilization. Environ. Sci. Technol. 2020, 54, 16195–16203. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, Z.; Yang, J.; Li, H.; Qu, W.; Leng, L.; Zhao, J.; Feng, Y.; Xu, Z.; Liu, H.; Shih, K. Advances in magnetically recyclable remediators for elemental mercury degradation in coal combustion flue gas. J. Mater. Chem. A 2020, 8, 18624–18650. [Google Scholar] [CrossRef]
  44. 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]
  45. Revellame, E.D.; Fortela, D.L.; Sharp, W.; Hernandez, R.; Zappi, M.E. Adsorption kinetic modeling using pseudo-first order and pseudo-second order rate laws: A review. Cleaner Eng. Technol. 2020, 1, 100032. [Google Scholar] [CrossRef]
  46. Pichler, C.; Lackner, R. Post-peak decelerating reaction of Portland cement: Monitoring by heat flow calorimetry, modelling by Elovich-Landsberg model and reaction-order model. Constr. Build. Mater. 2020, 231, 117107. [Google Scholar] [CrossRef]
  47. Cabooter, D.; Song, H.; Makey, D.; Sadriaj, D.; Dittmann, M.; Stoll, D.; Desmet, G. Measurement and modelling of the intra-particle diffusion and b-term in reversed-phase liquid chromatography. J. Chromatogr. A 2021, 1637, 461852. [Google Scholar] [CrossRef]
  48. Yang, J.; Li, Q.; Zhu, W.; Qu, W.; Li, M.; Xu, Z.; Yang, Z.; Liu, H.; Li, H. Recyclable chalcopyrite sorbent for mercury removal from coal combustion flue gas. Fuel 2021, 290, 120049. [Google Scholar] [CrossRef]
  49. Cheng, F.; Zhou, P.; Huo, X.; Liu, Y.; Liu, Y.; Zhang, Y. Enhancement of bisphenol A degradation by accelerating the Fe(III)/Fe(II) cycle in graphene oxide modified Fe(III)/peroxymonosulfate system under visible light irradiation. J. Colloid Interf. Sci. 2020, 580, 1021. [Google Scholar] [CrossRef]
  50. Feng, P.; Wang, W.; Wang, K.; Cheng, S.; Jiang, K. A high-performance carbon with sulfur doped between interlayers and its sodium storage mechanism as anode material for sodium ion batteries. J. Alloys Compd. 2019, 795, 223–232. [Google Scholar] [CrossRef]
  51. Yu, H.; Qian, C.; Ren, H.; Chen, M.; Tang, D.; Wu, H.; Lv, R. Enhanced catalytic properties of bimetallic sulfides with the assistance of graphene oxide for accelerating triiodide reduction in dye-sensitized solar cells. Sol. Energy 2020, 207, 1037–1044. [Google Scholar] [CrossRef]
  52. Wen, X.; Zhao, Z.; Zhai, S.; Wang, X.; Li, Y. Stable nitrogen and sulfur co-doped carbon dots for selective folate sensing, in vivo imaging and drug delivery. Diam. Relat. Mater. 2020, 105, 107791. [Google Scholar] [CrossRef]
  53. Asasian, N.; Kaghazchi, T.; Faramarzi, A.; Hakimi-Siboni, A.; Asadi-Kesheh, R.; Kavand, M.; Mohtashami, S.-A. Enhanced mercury adsorption capacity by sulfurization of activated carbon with SO2 in a bubbling fluidized bed reactor. J. Taiwan Inst. Chem. E. 2014, 45, 1588–1596. [Google Scholar] [CrossRef]
  54. Guo, Y.; Niu, J.; Zhang, H.; Cheng, F.; Wu, H.; Jin, D. Enhanced SO2 and Rhodamine B Removal by Blending Coke-Making Waste Benzene Residue (BR) for Pelletized Activated Coke (PAC) Production and Mechanisms. Energy Fuel 2019, 33, 5173–5181. [Google Scholar] [CrossRef]
  55. Liu, H.; Xie, X.; Chen, H.; Yang, S.; Liu, C.; Liu, Z.; Yang, Z.; Li, Q.; Yan, X. SO2 promoted ultrafine nano-sulfur dispersion for efficient and stable removal of gaseous elemental mercury. Fuel 2020, 261, 116367. [Google Scholar] [CrossRef]
  56. Zhou, M.; Xie, X.; Liu, Q.; Zhang, M.; Peng, C.; Li, F.; Liu, Q.; Song, Y.; Wu, J.; Qiao, Z. Spherical In2S3 anchored on g-C3N4 nanosheets for efficient elemental mercury removal in the wide temperature range. Chem. Eng. J. 2022, 430, 132857. [Google Scholar] [CrossRef]
  57. Alegre, C.; Sebastián, D.; Lázaro, M.J. Carbon xerogels electrochemical oxidation and correlation with their physico-chemical properties. Carbon 2019, 144, 382–394. [Google Scholar] [CrossRef] [Green Version]
  58. Hsu, C.J.; Chiou, H.J.; Chen, Y.H.; Lin, K.S.; Rood, M.J.; Hsi, H.C. Mercury adsorption and re-emission inhibition from actual WFGD wastewater using sulfur-containing activated carbon. Environ. Res. 2019, 168, 319–328. [Google Scholar] [CrossRef]
  59. Zhao, S.; Luo, H.; Ma, A.; Sun, Z.; Zheng, R. Experimental study on mercury removal from coal-fired flue gas by sulfur modified biomass coke with mechanochemical method. Fuel 2022, 309, 122201. [Google Scholar] [CrossRef]
  60. Yang, Q.; Yang, Z.; Li, H.; Zhao, J.; Yang, J.; Qu, W.; Shih, K. Selenide functionalized natural mineral sulfides as efficient sorbents for elemental mercury capture from coal combustion flue gas. Chem. Eng. J. 2020, 398, 125611. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthesis process for S-C.
Figure 1. Schematic illustration of the synthesis process for S-C.
Energies 15 01840 g001
Figure 2. (a) N2 absorption–desorption isotherms and (b) pore size distributions of S-C.
Figure 2. (a) N2 absorption–desorption isotherms and (b) pore size distributions of S-C.
Energies 15 01840 g002
Figure 3. SEM images of S-C with HF etchings for (a) 0 h, (b) 5 h, (c) 10 h, and (d) 20 h.
Figure 3. SEM images of S-C with HF etchings for (a) 0 h, (b) 5 h, (c) 10 h, and (d) 20 h.
Energies 15 01840 g003
Figure 4. FTIR spectra of the synthetic S-C sample.
Figure 4. FTIR spectra of the synthetic S-C sample.
Energies 15 01840 g004
Figure 5. XPS spectra of S-C for (a) C 1s, (b) O 1s, and (c) S 2p.
Figure 5. XPS spectra of S-C for (a) C 1s, (b) O 1s, and (c) S 2p.
Energies 15 01840 g005
Figure 6. Hg0 removal performances of S-C with different etching times.
Figure 6. Hg0 removal performances of S-C with different etching times.
Energies 15 01840 g006
Figure 7. Hg0 adsorption-capacity curve for S-AC as a function of time.
Figure 7. Hg0 adsorption-capacity curve for S-AC as a function of time.
Energies 15 01840 g007
Figure 8. Hg0 adsorption behaviors of S-C simulated by the (a) pseudo-first-order, (b) pseudo-second-order, (c) Elovich, and (d) intra-particle diffusion models.
Figure 8. Hg0 adsorption behaviors of S-C simulated by the (a) pseudo-first-order, (b) pseudo-second-order, (c) Elovich, and (d) intra-particle diffusion models.
Energies 15 01840 g008aEnergies 15 01840 g008b
Figure 9. Hg0 removal performances of (a) S-C and (b) S/C under different temperatures.
Figure 9. Hg0 removal performances of (a) S-C and (b) S/C under different temperatures.
Energies 15 01840 g009
Figure 10. Effects of (a) SO2, (b) NO, and (c) H2O on the Hg0 removal performances of S-Cs and (d) effects of flue gas components on the Hg0 removal performances of S/Cs.
Figure 10. Effects of (a) SO2, (b) NO, and (c) H2O on the Hg0 removal performances of S-Cs and (d) effects of flue gas components on the Hg0 removal performances of S/Cs.
Energies 15 01840 g010
Figure 11. (a) C 1s, (b) O 1s, (c) S 2p, and (d) Hg 4f spectra of spent S-C.
Figure 11. (a) C 1s, (b) O 1s, (c) S 2p, and (d) Hg 4f spectra of spent S-C.
Energies 15 01840 g011
Table 1. Pore structure parameters of various S-Cs with different HF etching times.
Table 1. Pore structure parameters of various S-Cs with different HF etching times.
HF Etching Time (h)Total Surface Area (m2·g−1)Surface Area of Micropores (m2·g−1)Pore Volume
(cm3·g−1)
Pore Diameter
(nm)
0180.92143.330.296.32
5267.66263.680.507.43
10318.50335.880.587.31
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, J.; Xu, H.; Meng, F.; Guo, Q.; He, T.; Yang, Z.; Qu, W.; Li, H. A Molten-Salt Pyrolysis Synthesis Strategy toward Sulfur-Functionalized Carbon for Elemental Mercury Removal from Coal-Combustion Flue Gas. Energies 2022, 15, 1840. https://doi.org/10.3390/en15051840

AMA Style

Yang J, Xu H, Meng F, Guo Q, He T, Yang Z, Qu W, Li H. A Molten-Salt Pyrolysis Synthesis Strategy toward Sulfur-Functionalized Carbon for Elemental Mercury Removal from Coal-Combustion Flue Gas. Energies. 2022; 15(5):1840. https://doi.org/10.3390/en15051840

Chicago/Turabian Style

Yang, Jianping, Hong Xu, Fanyue Meng, Qingjie Guo, Tao He, Zequn Yang, Wenqi Qu, and Hailong Li. 2022. "A Molten-Salt Pyrolysis Synthesis Strategy toward Sulfur-Functionalized Carbon for Elemental Mercury Removal from Coal-Combustion Flue Gas" Energies 15, no. 5: 1840. https://doi.org/10.3390/en15051840

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