Next Article in Journal
Hydrogen Absorption Performance and O2 Poisoning Resistance of Pd/ZrCo Composite Film
Next Article in Special Issue
Mineral Phase Reconstruction and Separation Behavior of Zinc and Iron from Zinc-Containing Dust
Previous Article in Journal
Effect of Strontium Substitution on the Tribocatalytic Performance of Barium Titanate
Previous Article in Special Issue
Removal of Sodium from Vanadium Tailings by Calcification Roasting in Reducing Atmosphere
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Achieving Large-Capability Adsorption of Hg0 in Wet Scrubbing by Defect-Rich Colloidal Copper Sulfides under High-SO2 Atmosphere

1
School of Metallurgy and Environment, Central South University, Changsha 410083, China
2
State Key Laboratory of Advanced Metallurgy for Non-Ferrous Metals, Changsha 410083, China
3
Chinese National Engineering Research Center for Control & Treatment of Heavy Metal Pollution, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(8), 3157; https://doi.org/10.3390/ma16083157
Submission received: 28 March 2023 / Revised: 13 April 2023 / Accepted: 13 April 2023 / Published: 17 April 2023
(This article belongs to the Special Issue Frontier of Environmental Friendly Recycling Technology for Metals)

Abstract

:
This paper reports on a novel method to remove Hg0 in the wet scrubbing process using defect-rich colloidal copper sulfides for reducing mercury emissions from non-ferrous smelting flue gas. Unexpectedly, it migrated the negative effect of SO2 on mercury removal performance, while also enhancing Hg0 adsorption. Colloidal copper sulfides demonstrated the superior Hg0 adsorption rate of 306.9 μg·g−1·min−1 under 6% SO2 + 6% O2 atmosphere with a removal efficiency of 99.1%, and the highest-ever Hg0 adsorption capacity of 736.5 mg·g−1, which was 277% higher than all other reported metal sulfides. The Cu and S sites transformation results reveal that SO2 could transform the tri-coordinate S sites into S22− on copper sulfides surfaces, while O2 regenerated Cu2+ via the oxidation of Cu+. The S22− and Cu2+ sites enhanced Hg0 oxidation, and the Hg2+ could strongly bind with tri-coordinate S sites. This study provides an effective strategy to achieve large-capability adsorption of Hg0 from non-ferrous smelting flue gas.

Graphical Abstract

1. Introduction

Mercury has been a contaminant of global concern due to its toxicity, persistence, and bioaccumulation since the implementation of the Minamata Convention on August, 2017 [1]. Among anthropogenic sources, the non-ferrous smelting sector is a major contributor, accounting for 14.9% of all worldwide emissions, as reported in the Global Mercury Assessment Report 2018 [2,3]. Therefore, it is urgent to efficiently remove Hg0 from non-ferrous smelting flue gas to abate the significant challenge of global mercury pollution control.
There are various forms of mercury in non-ferrous smelting flue gas, including elemental mercury (Hg0), oxidized mercury (Hg2+), and particulate mercury (Hgp) [4,5,6]. In the existing flue gas purification system, Hgp would be captured by the electrostatic precipitator (ESP) [7,8,9]. Hg2+ has strong water solubility and can be removed by wet scrubber, whereas the capture of Hg0 is difficult due to its strong stability, high volatility, and low solubility [10]. Wet methods for Hg0 removal promote the conversion of Hg0 to Hg2+, such as the Boliden–Norzink and advanced oxidation process, thereby improving mercury removal efficiency [11,12]. Hence, oxidation is considered as a vital way for Hg0 removal in the wet scrubbing of non-ferrous smelting flue gas [13].
Oxidative demercurization remains a major bottleneck in the application of wet scrubbing from non-ferrous smelting flue gas [14,15]. There is a large amount of reductive SO2 (>6% vol) in the flue gas, which is thousands of times higher than Hg0 concentration [16,17]. Selective oxidation of Hg0 poses a great challenge since the standard oxidation potential of Hg0 (0.85 V) is higher than that of SO2 (0.17 V) [18,19]. Consequently, SO2 is preferentially oxidized over Hg0, bringing about the depletion of oxidants and making it difficult to oxidize Hg0. Therefore, the efficient oxidation of Hg0 from high-SO2 flue gas is key to overcoming the bottleneck of Hg0 removal in wet scrubbing.
Transition metal sulfides (TMSs) have become a hot material for Hg0 capture due to the strong affinity of sulfur sites with Hg0 [20,21,22]. Previous studies had shown that SO2 could react with defective S2− on TMSs to generate highly active Sn2−, which might promote Hg0 oxidation [23]. O2 dissolves into aqueous phase and forms dissolved oxygen, which has a high oxidation activity due to hydrogen bonding and Van Der Waals force with water molecules, providing an oxidative environment for TMSs to capture Hg0 [24]. Therefore, if TMSs are added to the wet scrubber, it is expected to eliminate the negative impact of SO2 on the removal of Hg0, which could enhance the ability of TMSs to oxidize mercury, thereby greatly improving the adsorption capacity of mercury. However, the mechanism of SO2 and O2 in wet scrubbing for mercury capture by TMSs has not been reported.
Based on the above considerations, this paper involves a novel method of using metal sulfides in the wet scrubbing process for large-capability adsorption of Hg0 from high-SO2 flue gas. The optimal mercury adsorbent was prepared by sulfide precipitation and optimized systematically. The morphology and structure of the adsorbents were characterized. The excellent Hg0 removal performance of colloidal copper sulfides (c-CuS) was examined in the flue gas wet scrubber. The effects of SO2 and O2 on Hg0 removal by c-CuS were investigated, and the mechanism of the activation of oxidizing sites (S22− and Cu2+) by SO2 and O2 was proposed after identifying the structural changes of c-CuS under different atmospheres.

2. Experimental Section

2.1. Chemicals and Reagents

The chemicals in analytical grade were used in this study, including copper chloride (CuCl2), copper nitrate (CuNO3·6H2O), Copper sulfate (CuSO4·5H2O), sodium sulfide (NaS·9H2O), sodium hydroxide (NaOH), sodium chloride (NaCl), and nitric acid (HNO3). They were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) All the agents were directly used without any further purification. Ultrapure water was used in all experiments unless otherwise stated.

2.2. Synthesis of TMSs

TMSs suspensions were synthesized by the double-jet liquid-phase sulfidation precipitation method. Simply, 50 mL of a 10 mmol·L−1 metal chloride solution and an equal volume of 10 mmol·L−1 Na2S·9H2O solution were added simultaneously to the beaker at a speed of 1.2 mL min−1. These solutions were mixed by stirring at a speed of 380 rpm for 0.5 h, and then TMSs suspensions were obtained. c-CuS was synthesized using a single-jet liquid-phase sulfidation method. Then, 50 mL of Na2S·9H2O was rapidly added into a CuCl2 solution of 1.0 mmol·L−1, and then they were mixed by stirring. The experimental investigation on raw material concentration and ratio was undertaken to judiciously optimize the synthesis conditions.

2.3. Sample Characterization

These suspensions and colloids were filtered with nanopore filter membranes. The filtered residues were washed three times with ultrapure water and then dehydrated at −89 °C in a vacuum freeze-dryer for 8 h. The obtained filtered residues of TMSs suspensions and c-CuS were analyzed for crystalline structure by X-ray diffraction (XRD, Empyrean 2, PANalytic, Malvern, UK) with Cu-Kα radiation. High-resolution transmission electron microscopy (HRTEM, Titan G260-300, FEI, Lausanne, Switzerland) was used to observe the morphology of c-CuS particles. The scanning electron microscope (SEM, JSM-6360LV, Jeol, Tokyo, Japan) of c-CuS was characterized at the accelerating voltage of 200 kV after splashing a thin Au layer. The structure information of samples under different atmospheres was characterized by X-ray photoelectron spectroscopy (XPS, Escalab 250 Xi, Thermo Fisher Scientific, Waltham, MA, USA). The ultimate vacuum degree of the sample analysis room was 5 × 10−7 Pa, and the Al Kα X-ray was used as the excitation source with a power of 16 mA × 12.5 kV. Correction was made using C ls = 284.6 eV as the internal standard for electron binding energy. The fingerprint and structural properties of c-CuS were analyzed by three-dimensional excitation–emission matrix (3D-EEM, F7000, Hitachi, Ibaraki, Japan) fluorescence spectroscopy. 3D-EEM spectra were generated by scanning in the range of 200 nm to 600 nm, while excitation and emission sampling wavelengths were 10 nm, and the scanning speed was 12,000 nm·min−1. Finally, temperature-programmed desorption (TPD) was employed to identify the form of mercury species adsorbed on c-CuS. The TPD tests were carried out under pure nitrogen at the flow rate of 500 mL·min–1 from 50 °C to 600 °C at a rate of 5 °C·min−1.

2.4. Hg0 Removal Test

An experiment was conducted in a bubbling reactor to remove Hg0. Pure N2 was used as the carrier gas for the Hg0 generator (VICI Metronics) with a flow rate of 200 mL·min−1. The simulated flue gas (SFG) containing N2, SO2, and O2 at a flow rate of 300 mL·min−1 was mixed with the Hg0 carrier gas. The mixture gas was then blown into a bubbling reactor containing the scrubbing solution. Then, the SFG was treated with 5 mol·L−1 NaOH and quartz wool before the mercury detection to reduce the influence of SO2 and water vapor. The mercury analyzer (RA-915M, Lumex Zeeman) was used to monitor the outlet Hg0 concentration. All data were obtained by averaging three measurement results. Finally, the remaining Hg0 in the tail gas was absorbed by KMnO4 solution and activated carbon. The operating conditions in the experiment are listed in Table S1. The Hg0 removal efficiency and adsorption capacity were calculated according to Equations (1) and (2), respectively.
η = C i n C o u t C i n × 100 %
Q t = 1 / M × t 1 t 2 ( C i n l e t C o u t l e t ) t × u
where η (%) represents Hg0 removal efficiency, and Cin and Cout (μg·m−3) are the Hg0 concentration at the inlet and outlet of the experimental system, respectively. Q t (mg·g−1) is the adsorption capacity of Hg0, M (mg) is the mass of adsorbent, t1 and t2 (min) are the start and end times, Δt (min) is the scrubbing time, and u is the gas flow rate (m3·min−1).

2.5. DFT Calculation

Perdew–Burke-Ernzerhof (PBE0) functional was used to characterize the exchange-correlation effects in the density functional theory (DFT) by CP2K [25,26]. The CuS (110) system was simulated using an orthorhombic box with dimensions of 16.44 × 13.15 × 34.49 Å3, and the dipole correction technique was applied in the XY orientation. BROYDEN-MIXING was used to optimize the geometries, while the DZVP-MOLOPT-SR-GTH basis set was used to describe the basis. A plane wave cutoff of 400 Ry was set, and Goedecker–Teter–Hutter (GTH) pseudopotentials were used to represent all electrons [27,28]. In addition, the D3 dispersion correction by Grimme et al. was employed, and the adsorption energy (Ead) was calculated using Equation (3), based on the optimized structure.
Ead = E(Hg+CuS) − EHg − ECuS
where EHg is the state energy of a free Hg atom, ECuS represents the total energy of the CuS of different configurations, and E(Hg+CuS) is the total energy of the Hg atoms being adsorbed on CuS.

3. Results and Discussion

3.1. Selection and Optimization of Transition Metal Sulfides for Hg0 Removal

Hg0 removal performance was evaluated by adding TMSs suspensions into a bubble reactor as the scrubbing solution. Figure 1a shows that TMSs suspensions had a good ability to capture Hg0 in wet scrubbing. Especially, copper sulfides suspension was the preferred Hg0 removal scrubbing agent, with the highest efficiency of 87.6%. Furthermore, the effect of regulating feeding methods on the dispersibility of detergent was investigated to improve the Hg0 removal performance of copper sulfides. Figure 1b displays that c-CuS prepared by single-jet was homodispersed in solution, and the Tyndall effect was evident, while the precipitation of suspensions copper sulfides (s-CuS) was visible. The removal efficiency gradually decreased from 65.8% to 46.4% due to the poor contact between gas and s-CuS, while the mercury removal efficiency of c-CuS could be maintained at 97.5 ~ 98.9%. Therefore, the c-CuS synthesized by single-jet has better Hg0 capture performance.
Optimization of the raw material concentration and ratio to synthesize copper sulfides is crucial for enhancing Hg0 removal. As shown in Figure 1c and Figure S1, when the concentration of c-CuS was 1/32 mmol·L−1, the efficiency of Hg0 removal was only 65% due to the insufficient content of active components in the solution. Increasing the concentration of CuCl2 and Na2S fourfold to synthesize c-CuS could immediately improve the removal efficiency of Hg0 to 93%. When the concentrations were optimized to 1/2 mmol·L−1 and 2 mmol·L−1, the efficiencies were higher than 96.3%. Unfortunately, a large amount of precipitation appeared in the solution of 2 mmol·L−1 concentration, which was not conducive to sufficient contact with Hg0. Finally, the Cu/S ratio of copper sulfides was optimized, as shown in Figure 1d. Copper sulfides had the best removal efficiency for Hg0, which was achieved by reaching 98.1% when the Cu:S ratio was 1:1. However, as shown in Figure S2, further increasing the S2− ratio caused possible instability of colloid, which would have a negative effect on the Hg0 removal. In summary, the c-CuS was prepared with the concentration of 1/2 mmol·L−1 and the Cu:S raw material ratio of 1:1, which was the optimal mercury capture agent.

3.2. Structural Characterizations of c-CuS

Figure 2 shows the XRD patterns of the obtained c-CuS and s-CuS filter residues. The phase was confirmed to be Cu4(S2)2(CuS)2 (JCPDS card NO. 74-1234), with overlapping diffraction peaks that could be divided into lattice planes such as (101), (102), (103), (110), and (116) [29]. The diffraction peaks of s-CuS were stronger and sharper, indicating a better crystallinity. According to Equation (4), the crystallinity of s-CuS was calculated to be 70.9%. Surprisingly, the crystallinity of the c-CuS sample was only about 3.4% with severely broadened diffraction peaks. It indicated that c-CuS might contain abundant defects, which was beneficial for exposing more active adsorption sites to capture Hg0 [15,30].
X c = I c I c + I a × 100 %
where Ic is the crystal diffraction intensity and Ia is the amorphous scattering intensity.
Figure 3 shows the XPS full spectrum of c-CuS. The peak analysis result indicated that the Cu:S atomic ratio was 0.98:1 on the surface of c-CuS filter residue (the O, C peak belongs to the peak of carrier sheet). The mismatched stoichiometric ratio suggested that rich sulfur sites were exposed. Figure 4a shows that c-CuS was composed of near-spherical nanoparticles with the diameter of about 8.0 nm, similar to those of SEM in Figure S3. HRTEM images in Figure 4b display many orange-marked vacancies and amorphous regions in the lattice, indicating that rich defect that could provide more active sites for mercury capture. The measured lattice spacing of 1.89 nm was consistent with the (110) plane of c-CuS. The broad diffraction rings in the selected-area electron diffraction (SAED) shown in Figure 4c further confirmed the existence of amorphous regions in c-CuS [31]. Figure 4d–f shows the high-angle annular dark-field image (HADDF), Cu element, and S element scanning images of c-CuS, respectively. It confirms that Cu and S were evenly distributed. In summary, c-CuS with abundant defect sites has the potential for high-activity mercury capture.

3.3. Hg0 Removal Performance of c-CuS under SO2 and O2 Atmospheres

SO2 is typically considered to be a negative effects gas that inhibits Hg0 removal in industrial flue gas [23,32]. SO2 resistance has a crucial impact on the mercury capture performance of adsorbents. SO2-contained simulated flue gas was inputted into a bubbling reactor with 10 mL of c-CuS to investigate the effect of SO2 on the Hg0 removal performance of c-CuS. As shown in Figure 5a, the removal efficiency was 80.1% in the absence of SO2. The efficiency significantly increased to 95.1 ~ 97.4% under 3 ~ 9% vol SO2. Furthermore, the performance of c-CuS for Hg0 removal was compared with and without SO2 pretreatment, as shown in Figure 5b. The outlet concentration of Hg0 was about 10 μg·m−3 after passing through c-CuS solution pretreated without SO2, but it decreased immediately to ~0 μg·m−3 after turning on SO2. When c-CuS solution was pretreated with SO2, the outlet Hg0 concentration was about 3 μg·m−3 after opening gas pipeline of Hg0. Meanwhile, the outlet concentration exhibited almost no fluctuation after supplying SO2 and Hg0 simultaneously. This might mean that SO2 did not occupy the adsorption sites, and SO2 could interact with c-CuS to enhance adsorption with Hg0. The 3D-EEM was used to compare the fluorescence fingerprint spectra of c-CuS under N2 + 6% vol SO2 atmospheres. As shown in Figure 5c,d, the region II area of the sample treated with SO2 was smaller than the region I area of c-CuS under N2 atmosphere, confirming that SO2 leads to the formation of more vacancies. The formation of vacancies meant that low-active S2− had been converted to more active S22− [33,34]. The specific conversion process of activity sites will be further studied in the following.
In Figure S4, it could be observed that the capture of mercury was almost unaffected under an atmosphere of 3 ~ 9% vol O2. This was attributed to the fact that c-CuS solution could dissolve only 0.076 mmol·L−1 oxygen, which was much higher than the amount of mercury and much lower than the O2 proportion in the flue gas. Therefore, we conducted further investigations to explore the influence of dissolved oxygen on Hg0 oxidation. As displayed in Figure 6a, the dissolved oxygen concentration in c-CuS solution increased from 0.6 mg∙L−1 to 7.6 mg∙L−1. The redox potential of the c-CuS solution gradually increased from 166 mV to 222 mV, which meant enhancing the oxidation ability of the scrubbing solution. It led to a significant increase in the removal efficiency of mercury from 45.0% to 95.3%. As shown in Figure 6b, the results displayed that the Cu LMM auger spectral peaks included Cu2+ of binding energy at 568.1 eV and Cu+ at 565.8 eV [35]. The Cu+ spectral peak of c-CuS in oxygen-free water was visible, while it was weakened for c-CuS in oxic water, conversely indicating the oxidation of Cu+ to Cu2+ by O2. It could be inferred that dissolved oxygen in an aqueous solution could increase the Cu2+ content of c-CuS and strengthen the oxidation of Hg0.
The Hg0 adsorption capacity and rate of c-CuS were investigated and calculated by penetration experiments. The prepared 10 mL c-CuS (containing 0.48 mg CuS particles) was input with 6% vol SO2 + 6% vol O2 flue gas containing an inlet Hg0 concentration of 890.0 µg·m−3. As shown in Figure 7, the outlet concentration of Hg0 gradually increased from 303.2 µg·m−3 to 835.5 µg·m−3 after 2400 min of scrubbing experiment. The Hg0 adsorption capacity of c-CuS was calculated to be 736.5 mg·g−1, and the average adsorption rate was 306.9 μg·g−1·min−1. Figure 8 and Table S2 demonstrate the comparison of the adsorption capacity and rate of c-CuS with previously reported metal sulfides. c-CuS has the highest adsorption capacity and rate among the reported sulfide adsorbents. Compared to all other adsorbents, the Hg0 adsorption capacity of c-CuS is exceptionally high, exceeding them by an impressive 277%, such as nano-CuS, Co3S4, S/FeS2, ZnS, and so on. Additionally, the Hg0 adsorption rate of c-CuS is much higher than that of other mineral sulfides, which is due to the abundant adsorption sites under the positive effect of SO2 and O2 [7,14,36,37,38,39,40].

3.4. Mechanism for Hg0 Adsorption

As mentioned above, SO2 and O2 have positive effects on c-CuS for Hg0 capture. To determine the mechanism of Hg0 adsorption under SO2 and O2 atmospheres, as shown in Figure 9, the XPS spectra of spent c-CuS samples were analyzed under different atmospheres. Shown in Figure 9a–d are the S 2p peaks of samples under N2, 6% vol SO2, 6% vol SO2 + 6% vol O2, and 6% vol SO2 + 6% vol O2 + Hg0, respectively. The forms of S on the c-CuS surface are multiple, such as tri-coordinated S2−(CN=3), tetra-coordinated S2−(CN=4), and sulfur–sulfur-coordinated S22− [23]. After peak fitting, the peaks at 161.1 eV and 162.2 eV were assigned to S2−(CN=3), while the peaks at 161.9 eV and 163.0 eV were attributed to S2−(CN=4). The peaks at 163.2 eV and 164.4 eV were attributed to S22−, while the peaks at 167.9 eV and 168.9 eV belonged to SO42− [37,38,41].
Under N2 atmosphere, the ratio of S and Cu total amount (St, Cut) on c-CuS was found to be 1.01:1, with S 2p consisting of 26.8% S2−(CN=3), 39.0% S2−(CN=4), 33.5% S22−, and 0.7% SO42−. Upon treatment with 6% vol SO2, the St:Cut ratio increased to 1.79:1, with a decrease in S2−(CN=3) content to 10.4%, an increase in S22− to 49.5%, and a slight change in other forms of S, indicating the formation of new S22− by the combination of S2−(CN=3) with SO2. Additionally, the Cu 2p peak in Figure 10 shifted towards higher binding energy, indicating the reduction of a part of Cu2+ to Cu+ by SO2. Similarly, under 6% vol SO2 + 6% vol O2 conditions, St:Cut decreased with the oxidation of Cu+ into Cu2+ by O2 and inhibition of S2−(CN=4) combination with SO2 adsorbed on the c-CuS surface. The S2−(CN=3) content decreased with an increase in S2−(CN=4), which led to less S22−. The further supply of Hg0 resulted in a decrease in the ratio of S22− and S2−(CN=4), but an increase in S2−(CN=3), indicating the transformation of S22− and S2−(CN=4) into S2−(CN=3) after capture of Hg0. The results of Cu 2p shifting towards higher binding energy confirmed that Cu2+ was also an active site for Hg0 oxidation. It suggests that S22− and Cu2+ serve as oxidation sites, while S2−(CN=3) is the binding site with Hg after adsorption. The TPD results in Figure 11 indicated that the decomposition temperature of Hg on the c-CuS was about 195 ~ 220 °C, which suggested the presence of black HgS [38].
The experimental results were inconclusive about the main binding site of the adsorbed mercury. Therefore, the Ead at each site on the c-CuS (110) crystal was analyzed at the molecular level using DFT calculations. The result in Figure 12 indicated that the tri-coordinate sulfur sites had the highest adsorption energy of −125.23 kJ·mol−1. It means that the tri-coordinate sulfur site might be the main binding site for mercury, which is consistent with XPS results.
In summary, the mechanism for Hg0 capture under SO2 and O2 atmospheres could be described by Equations (5) ~ (10). After the adsorbing of Hg0 on the c-CuS surface, Cu2+ and S22− sites oxidized Hg0ad to Hg2+ and binded with S2−(CN=3) to form HgS. As described in Equations (9) and (10), SO2 and O2 respectively activate S2−(CN=3) and Cu+ transformed into S22− and Cu2+ as Hg0 oxidation sites.
Hg0 → Hg0ad
Hg0ad + 2Cu2+ − S2−(CN=4) → Hg2+ + 2Cu+ − S2−(CN=3)
Hg2+ + S2−(CN=3) → HgS
Hg0ad + S22− → HgS + S2−(CN=3)
4Cu+ + O2 + 4H+ → 4Cu2+ + 2H2O
3S2−(CN=3) + HSO3+ 3H+ → 2S22− + 2H2O

4. Conclusions

In this paper, the optimal conditions for preparing c-CuS with a Cu–S ratio of 1:1 and a concentration of 0.5 mmol·L−1 were synthesized by a double-jet liquid-phase sulfidation precipitation method. Defect-rich c-CuS with low crystallinity was used as a scrubbing agent to remove mercury from flue gas in wet scrubbing. It migrated the negative effect of SO2 on mercury removal performance, while also enhancing mercury capture. The Hg0 removal efficiency of c-CuS was 99.1% under 6% vol SO2 + 6% vol O2 atmosphere. The Hg0 adsorption capacity of c-CuS reached 736.5 mg·g−1, and the average adsorption rate was 306.9 μg·g−1·min−1, which was far better than other reported metal sulfides adsorbents. Based on structural characterization and DFT calculation, it was found that SO2 and O2 can enhance the formation of Cu2+ and S22− sites from Cu+ and S2−(CN=3) to promote the oxidation of Hg0 to Hg2+, and then Hg2+ could strongly bind with S2−(CN=3). However, reusability of the material should be developed in future studies.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma16083157/s1: Figure S1. Pictures of c-CuS prepared at different concentrations. Figure S2. (a) The pictures of c-CuS prepared with different Cu:S ratios. (b) Their phenomenon of Tyndall Effect. Figure S3. The SEM pictures of c-CuS. Figure S4. The effect of O2. Figure S5. The structure of configurations A, B, C, and D after mercury adsorption. Table S1. List of experimental conditions for Hg0 capture. Table S2. Hg0 adsorption capacities and rates of different sulfide sorbents. References [36,38,39,42,43,44,45,46,47,48,49] are cited in the supplementary materials.

Author Contributions

X.X.: investigation, formal analysis, writing—original draft, writing—review and editing; H.C. and X.L.: formal analysis; K.X.: formal analysis, writing—review and editing, funding acquisition; H.L.: formal analysis, writing—review and editing, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 52234011), and the Innovative Research Group Project of the National Natural Science Foundation of China (No. 52121004).

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. Selin, H. Global environmental law and treaty-making on hazardous substances: The Minamata Convention and mercury abatement. Glob. Environ. Politics 2014, 14, 1–19. [Google Scholar] [CrossRef]
  2. Outridge, P.M.; Mason, R.; Wang, F.; Guerrero, S.; Heimburger-Boavida, L. Updated global and oceanic mercury budgets for the United Nations Global Mercury Assessment 2018. Environ. Sci. Technol. 2018, 52, 11466–11477. [Google Scholar] [CrossRef]
  3. Liu, W.; Xu, H.; Guo, Y.; Yuan, Y.; Liao, Y.; Qu, Z.; Yan, N. Immobilization of elemental mercury in non-ferrous metal smelting gas using ZnSe1−xSx nanoparticles. Fuel 2019, 254, 115641. [Google Scholar] [CrossRef]
  4. Sun, X.; Huang, W.; Ji, L.; Xu, H.; Qu, Z.; Yan, N. Establishing a Self-supporting System of H2S Production from SO2 with Induced Catalytic Reduction Process for Mercury Capture with Super-large Enrichment. Chem. Eng. J. 2023, 459, 141493. [Google Scholar] [CrossRef]
  5. Liu, H.; Shen, F.; Li, Q.; Wen, M.; Zhang, H.; Jiang, L.; Zheng, C.; Liu, Y.; Liu, T.; Chai, L. Systematic control technologies for gaseous pollutants from non-ferrous metallurgy. J. Environ. Sci. 2023, 123, 65–82. [Google Scholar] [CrossRef] [PubMed]
  6. Luo, H.; Zhao, S.; Ma, A.; Sun, K.; Zhu, Y.; Sun, Z. Influence of flue gas components on the mercury adsorption/oxidation by mechanochemical S2Cl2-modified sawdust coke. Fuel 2023, 336, 127154. [Google Scholar] [CrossRef]
  7. Xie, X.; Zhang, Z.; Chen, Z.; Wu, J.; Li, Z.; Zhong, S.; Liu, H.; Xu, Z.; Zhilou, L. In-situ preparation of zinc sulfide adsorbent using local materials for elemental mercury immobilization and recovery from zinc smelting flue gas. Chem. Eng. J. 2022, 429, 132115. [Google Scholar] [CrossRef]
  8. Yang, J.; Zhu, P.; Meng, F.; Guo, Q.; He, T.; Yang, Z.; Qu, W.; Li, H. Charge distribution modulation and morphology controlling of copper selenide for an enhanced elemental mercury adsorption activity in flue gas. Chem. Eng. J. 2022, 442, 136145. [Google Scholar] [CrossRef]
  9. Wu, S.; Ma, Y.; Yu, W.; Wang, H.; Yang, W.; Che, L.; Han, Z. SO2 resistance of CeO2-and Co3O4-supported activated carbon during removal of mercury from flue gas: A comparative study. Fuel 2023, 334, 126636. [Google Scholar] [CrossRef]
  10. Ishag, A.; Yue, Y.; Xiao, J.; Huang, X.; Sun, Y. Recent advances on the adsorption and oxidation of mercury from coal-fired flue gas: A review. J. Clean. Prod. 2022, 367, 133111. [Google Scholar] [CrossRef]
  11. Peng, J.; Yin, R.; Yang, X.; Shang, C. A novel UVA/ClO2 advanced oxidation process for the degradation of micropollutants in water. Environ. Sci. Technol. 2022, 56, 1257–1266. [Google Scholar] [CrossRef]
  12. Huang, T.; Geng, X.; Liu, X.; Liu, J.; Duan, Y.; Zhao, S.; Gupta, R. Reduction of HgCl2 to Hg0 in flue gas at high temperature. Part I: Influences of oxidative species. Fuel 2022, 324, 124417. [Google Scholar] [CrossRef]
  13. Ye, D.; Wang, X.-X.; Wang, R.-X.; Liu, X.; Liu, H.; Wang, H.-N. Review of elemental mercury (Hg0) removal by CuO-based materials. J. Zhejiang Univ.-Sci. A 2022, 23, 505–526. [Google Scholar] [CrossRef]
  14. Kumar, A.S.; Kalidhasan, S.; Rajesh, V.; Rajesh, N. A meticulous study on the adsorption of mercury as tetrachloromercurate (II) anion with trioctylamine modified sodium montmorillonite and its application to a coal fly ash sample. Ind. Eng. Chem. Res. 2012, 51, 11312–11327. [Google Scholar] [CrossRef]
  15. Isinkaralar, K. High-efficiency removal of benzene vapor using activated carbon from Althaea officinalis L. biomass as a lignocellulosic precursor. Environ. Sci. Pollut. Res. 2022, 29, 66728–66740. [Google Scholar] [CrossRef]
  16. Liu, H.; Liu, X.; Wang, Z.; Yi, H.; Fu, Y.; Wu, L.; Shen, F.; Wang, P.; Liu, M.; Lin, Z.; et al. A gas permeable membrane electrode for selective and durable H2S production from SO2. Chem. Eng. J. 2023, 454, 140052. [Google Scholar] [CrossRef]
  17. Liu, T.; Xiong, Z.; Ni, P.; Ma, Z.; Tan, Y.; Li, Z.; Deng, S.; Li, Y.; Yang, Q.; Zhang, H. Review on adsorbents in elemental mercury removal in coal combustion flue gas, smelting flue gas and natural gas. Chem. Eng. J. 2023, 454, 140095. [Google Scholar] [CrossRef]
  18. He, Z.; Zhou, Z.; Wei, P.; Xu, T.; Han, J.; Huang, K.; Guo, K.; Huang, W.; Akasaka, T.; Lu, X. Fullerene-Derived Porous and Defective N-Doped Carbon Nanosheets as Advanced Trifunctional Metal-Free Electrocatalysts. Chem. Asian J. 2023, 18, e202200994. [Google Scholar] [CrossRef]
  19. Yan, W.; Wang, D.; Botte, G.G. Electrochemical decomposition of urea with Ni-based catalysts. Appl. Catal. B Environ. 2012, 127, 221–226. [Google Scholar] [CrossRef]
  20. Martellaro, P.; Moore, G.; Peterson, E.; Abbott, E.; Gorenbain, A. Environmental application of mineral sulfides for removal of gas-phase Hg (0) and aqueous Hg2+. Sep. Sci. Technol. 2001, 36, 1183–1196. [Google Scholar] [CrossRef]
  21. Liu, H.; Ruan, W.; Zhang, Z.; Zhou, Y.; Shen, F.; Liu, J.; Yang, H. Performance and mechanism of CuS-modified MWCNTs on mercury removal: Experimental and density functional theory study. Fuel 2022, 309, 122238. [Google Scholar] [CrossRef]
  22. Zheng, W.; Li, H.; Yang, Z.; Yang, J.; Qu, W.; Meng, F.; Feng, Y.; Xu, Z.; Guo, X. Advances in flue gas mercury abatement by mineral chalcogenides. Chem. Eng. J. 2021, 411, 128608. [Google Scholar] [CrossRef]
  23. Li, H.; Zu, H.; Deng, Y.; He, W.; Yang, Z.; Yang, J.; Zhao, S.; Qu, W. Mechanisms of gas-phase mercury immobilized by metal sulfides from combustion flue gas: A mini review. Energy Fuels 2022, 36, 6027–6037. [Google Scholar] [CrossRef]
  24. Li, Y.; Dong, H.; Li, L.; Tang, L.; Tian, R.; Li, R.; Chen, J.; Xie, Q.; Jin, Z.; Xiao, J.; et al. Recent advances in waste water treatment through transition metal sulfides-based advanced oxidation processes. Water Res. 2021, 192, 116850. [Google Scholar] [CrossRef] [PubMed]
  25. Kühne, T.D.; Iannuzzi, M.; Del Ben, M.; Rybkin, V.V.; Seewald, P.; Stein, F.; Laino, T.; Khaliullin, R.Z.; Schütt, O.; Schiffmann, F.; et al. CP2K: An electronic structure and molecular dynamics software package-Quickstep: Efficient and accurate electronic structure calculations. J. Chem. Phys. 2020, 152, 194103. [Google Scholar] [CrossRef]
  26. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  27. Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703. [Google Scholar] [CrossRef] [PubMed]
  28. Hartwigsen, C.; Gœdecker, S.; Hutter, J. Relativistic separable dual-space Gaussian pseudopotentials from H to Rn. Phys. Rev. B 1998, 58, 3641. [Google Scholar] [CrossRef]
  29. Sathiyaraj, E.; Thirumaran, S. Structural, morphological and optical properties of iron sulfide, cobalt sulfide, copper sulfide, zinc sulfide and copper-iron sulfide nanoparticles synthesized from single source precursors. Chem. Phys. Lett. 2020, 739, 136972. [Google Scholar] [CrossRef]
  30. Liu, K.; Liu, Z.; Yao, S.; Sun, S.; Ma, J.; Sun, R. CuInS2 quantum dots anchored onto the three-dimensional flexible self-supporting graphene oxide array with regulatable crystallinity and defect density for efficient photocatalytic synthesis of xylonic acid. Appl. Catal. B Environ. 2022, 316, 121573. [Google Scholar] [CrossRef]
  31. Xiao, J.; Xie, Y.; Tang, R.; Chen, M.; Tian, X. Novel ultrasonically assisted templated synthesis of palladium and silver dendritic nanostructures. Adv. Mater. 2001, 13, 1887–1891. [Google Scholar] [CrossRef]
  32. Liu, X.; Wu, L.; Wang, Z.; Fu, Y.; Xie, X.; Chen, H.; Long, J.; Xiang, K.; Liu, H. Influence of the Interface Wettability of Membranes on Electrolysis Reduction of SO2 to H2S. Ind. Eng. Chem. Res. 2023, 62, 1550–1557. [Google Scholar] [CrossRef]
  33. Li, Z.; Ma, H.; Zang, L.; Li, D.; Guo, S.; Shi, L. Construction of nano-flower MIL-125 (Mo)-In2Se3 Z-scheme heterojunctions by one-step solvothermal method for removal of tetracycline from wastewater in the synergy of adsorption and photocatalysis way. Sep. Purif. Technol. 2021, 276, 119355. [Google Scholar] [CrossRef]
  34. Lesnyak, V.; Brescia, R.; Messina, G.C.; Manna, L. Cu vacancies boost cation exchange reactions in copper selenide nanocrystals. J. Am. Chem. Soc. 2015, 137, 9315–9323. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Y.; Chao, D.; Wang, Z.; Ni, J.; Li, L. An energetic CuS–Cu battery system based on CuS nanosheet arrays. ACS Nano 2021, 15, 5420–5427. [Google Scholar] [CrossRef]
  36. Li, H.; Yang, Q.; Qu, W.; Feng, Y.; Leng, L.; Zhao, J.; Meng, F.; Yang, Z.; Yang, J. Facile preparation of nanosized copper sulfide functionalized macroporous skeleton for efficient vapor-phase mercury sequestration. Chem. Eng. J. 2021, 419, 129561. [Google Scholar] [CrossRef]
  37. Ma, S.; Shim, Y.; Islam, S.M.; Subrahmanyam, K.S.; Wang, P.; Li, H.; Wang, S.; Yang, X.; Kanatzidis, M.G. Efficient Hg vapor capture with polysulfide intercalated layered double hydroxides. Chem. Mater. 2014, 26, 5004–5011. [Google Scholar] [CrossRef]
  38. Liu, W.; Xu, H.; Liao, Y.; Quan, Z.; Li, S.; Zhao, S.; Qu, Z.; Yan, N. Recyclable CuS sorbent with large mercury adsorption capacity in the presence of SO2 from non-ferrous metal smelting flue gas. Fuel 2019, 235, 847–854. [Google Scholar] [CrossRef]
  39. Li, H.; Zhu, W.; Yang, J.; Zhang, M.; Zhao, J.; Qu, W. Sulfur abundant S/FeS2 for efficient removal of mercury from coal-fired power plants. Fuel 2018, 232, 476–484. [Google Scholar] [CrossRef]
  40. Isinkaralar, K.; Gullu, G.; Turkyilmaz, A.; Dogan, M.; Turhan, O. Activated carbon production from horse chestnut shells for hydrogen storage. Int. J. Glob. Warm. 2022, 26, 361–373. [Google Scholar] [CrossRef]
  41. Waltrovitz, R.R.; Qian, G.; de Groot, F.; Quinton, J.S.; Harmer, S.L. Influence of Co: Fe: Ni ratio on cobalt Pentlandite’s electronic structure and surface speciation. Miner. Eng. 2022, 190, 107935. [Google Scholar] [CrossRef]
  42. 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]
  43. Hong, Q.; Xu, H.; Li, J.; Huang, W.; Liu, P.; Qu, Z.; Yan, N. Shell-thickness-induced spontaneous inward migration of mercury in porous ZnO@ CuS for gaseous mercury immobilization. Chem. Eng. J. 2021, 420, 127592. [Google Scholar] [CrossRef]
  44. Yang, Z.; Li, H.; Feng, S.; Li, P.; Liao, C.; Liu, X.; Zhao, J.; Yang, J.; Lee, P.-H.; Shih, K. Multiform sulfur adsorption centers and copper-terminated active sites of nano-CuS for efficient elemental mercury capture from coal combustion flue gas. Langmuir 2018, 34, 8739–8749. [Google Scholar] [CrossRef]
  45. 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]
  46. Liu, Z.; Li, Z.; Xie, X.; Yang, S.; Fei, J.; Li, Y.; Xu, Z.; Liu, H. Development of recyclable iron sulfide/selenide microparticles with high performance for elemental mercury capture from smelting flue gas over a wide temperature range. Environ. Sci. Technol. 2019, 54, 604–612. [Google Scholar] [CrossRef]
  47. Yang, S.; Wang, D.; Liu, H.; Liu, C.; Xie, X.; Xu, Z.; Liu, Z. Highly stable activated carbon composite material to selectively capture gas-phase elemental mercury from smelting flue gas: Copper polysulfide modification. Chem. Eng. J. 2019, 358, 1235–1242. [Google Scholar] [CrossRef]
  48. 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]
  49. 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]
Figure 1. (a) Hg0 removal performances of different TMSs suspensions. (b) Long-time Hg0 removal performances of c-CuS and s-CuS. (c) The effect of raw material concentration on Hg0 removal. (d) The effect of Cu and S ratio on Hg0 removal (Experimental condition: solution volume = 80 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was pure N2, the inlet concentration of Hg0 was 200 µg·m−3).
Figure 1. (a) Hg0 removal performances of different TMSs suspensions. (b) Long-time Hg0 removal performances of c-CuS and s-CuS. (c) The effect of raw material concentration on Hg0 removal. (d) The effect of Cu and S ratio on Hg0 removal (Experimental condition: solution volume = 80 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was pure N2, the inlet concentration of Hg0 was 200 µg·m−3).
Materials 16 03157 g001
Figure 2. The XRD patterns of c-CuS and s-CuS.
Figure 2. The XRD patterns of c-CuS and s-CuS.
Materials 16 03157 g002
Figure 3. The XPS full spectra survey of c-CuS.
Figure 3. The XPS full spectra survey of c-CuS.
Materials 16 03157 g003
Figure 4. (a) TEM image of c-CuS. (b) HRTEM image of c-CuS. (c) SAED image of c-CuS. (d) HADDF image of c-CuS. (e) Cu and (f) S element mapping images of c-CuS (Prepared condition of c-CuS was the concentration of 1/2 mmol·L−1 and the Cu:S raw material ratio of 1:1 by using a single-jet liquid-phase sulfidation method).
Figure 4. (a) TEM image of c-CuS. (b) HRTEM image of c-CuS. (c) SAED image of c-CuS. (d) HADDF image of c-CuS. (e) Cu and (f) S element mapping images of c-CuS (Prepared condition of c-CuS was the concentration of 1/2 mmol·L−1 and the Cu:S raw material ratio of 1:1 by using a single-jet liquid-phase sulfidation method).
Materials 16 03157 g004
Figure 5. (a) The effect of SO2 on Hg0 removal efficiency. (b) Hg0 removal performances under SO2 on–off experiment. (c) The 3D-EEM fluorescence spectrum of c-CuS under pure N2. (d) The 3D-EEM fluorescence spectrum of c-CuS under SO2 (Experimental condition: solution volume = 10 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was 0 ~ 9% vol SO2 + N2, the inlet concentration of Hg0 was 200 µg·m−3).
Figure 5. (a) The effect of SO2 on Hg0 removal efficiency. (b) Hg0 removal performances under SO2 on–off experiment. (c) The 3D-EEM fluorescence spectrum of c-CuS under pure N2. (d) The 3D-EEM fluorescence spectrum of c-CuS under SO2 (Experimental condition: solution volume = 10 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was 0 ~ 9% vol SO2 + N2, the inlet concentration of Hg0 was 200 µg·m−3).
Materials 16 03157 g005
Figure 6. (a) The relationship of redox potential and Hg0 removal efficiency at varied dissolved O2 concentrations (Experimental condition: solution volume = 10 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was 0 ~ 9% vol O2 + N2, the inlet concentration of Hg0 was 200 µg·m−3). (b) XPS spectra of Cu LMM for c-CuS in oxygen-free water and oxic water.
Figure 6. (a) The relationship of redox potential and Hg0 removal efficiency at varied dissolved O2 concentrations (Experimental condition: solution volume = 10 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was 0 ~ 9% vol O2 + N2, the inlet concentration of Hg0 was 200 µg·m−3). (b) XPS spectra of Cu LMM for c-CuS in oxygen-free water and oxic water.
Materials 16 03157 g006
Figure 7. The adsorption breakthrough curve of c-CuS under 6% vol SO2+ 6% vol O2 (Experimental condition: solution volume = 10 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was 6% vol SO2 + 6% vol O2 + N2, the inlet concentration of Hg0 was 890 µg·m−3).
Figure 7. The adsorption breakthrough curve of c-CuS under 6% vol SO2+ 6% vol O2 (Experimental condition: solution volume = 10 mL, solution pH = 3.0, solution temperature = 20 °C, flue gas was 6% vol SO2 + 6% vol O2 + N2, the inlet concentration of Hg0 was 890 µg·m−3).
Materials 16 03157 g007
Figure 8. Comparison of adsorption capacity and rate with reported metal sulfide materials.
Figure 8. Comparison of adsorption capacity and rate with reported metal sulfide materials.
Materials 16 03157 g008
Figure 9. XPS spectra of S 2p for c-CuS under (a) N2, (b) 6% vol SO2, (c) 6% vol SO2 + 6% vol O2, (d) 6% vol SO2 + 6% vol O2 +Hg0 (Experimental condition: solution volume = 80 mL, solution pH = 3.0, solution temperature = 20 °C, the inlet concentration of Hg0 was 200 µg·m−3).
Figure 9. XPS spectra of S 2p for c-CuS under (a) N2, (b) 6% vol SO2, (c) 6% vol SO2 + 6% vol O2, (d) 6% vol SO2 + 6% vol O2 +Hg0 (Experimental condition: solution volume = 80 mL, solution pH = 3.0, solution temperature = 20 °C, the inlet concentration of Hg0 was 200 µg·m−3).
Materials 16 03157 g009
Figure 10. Cu 2p of spent c-CuS samples under pure N2, 6% vol SO2, 6% vol SO2 + 6% vol O2, and 6% vol SO2 + 6% vol O2 +Hg0.
Figure 10. Cu 2p of spent c-CuS samples under pure N2, 6% vol SO2, 6% vol SO2 + 6% vol O2, and 6% vol SO2 + 6% vol O2 +Hg0.
Materials 16 03157 g010
Figure 11. The TPD curves of spent samples after Hg0 adsorption under different atmospheres.
Figure 11. The TPD curves of spent samples after Hg0 adsorption under different atmospheres.
Materials 16 03157 g011
Figure 12. The Ead on CuS (110) of Hg0 with different sites (A: di-coordinated Cu, B: tri-coordinated Cu, C: tri-coordinated S, and D: tetra-coordinated S. The structural configurations shown in Figure S4).
Figure 12. The Ead on CuS (110) of Hg0 with different sites (A: di-coordinated Cu, B: tri-coordinated Cu, C: tri-coordinated S, and D: tetra-coordinated S. The structural configurations shown in Figure S4).
Materials 16 03157 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xie, X.; Chen, H.; Liu, X.; Xiang, K.; Liu, H. Achieving Large-Capability Adsorption of Hg0 in Wet Scrubbing by Defect-Rich Colloidal Copper Sulfides under High-SO2 Atmosphere. Materials 2023, 16, 3157. https://doi.org/10.3390/ma16083157

AMA Style

Xie X, Chen H, Liu X, Xiang K, Liu H. Achieving Large-Capability Adsorption of Hg0 in Wet Scrubbing by Defect-Rich Colloidal Copper Sulfides under High-SO2 Atmosphere. Materials. 2023; 16(8):3157. https://doi.org/10.3390/ma16083157

Chicago/Turabian Style

Xie, Xiaofeng, Hao Chen, Xudong Liu, Kaisong Xiang, and Hui Liu. 2023. "Achieving Large-Capability Adsorption of Hg0 in Wet Scrubbing by Defect-Rich Colloidal Copper Sulfides under High-SO2 Atmosphere" Materials 16, no. 8: 3157. https://doi.org/10.3390/ma16083157

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