[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,271)

Search Parameters:
Keywords = flotation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 5169 KiB  
Article
Research on the Flotation Mechanism of Microemulsion Collector Enhanced Removal of Dyeing Impurities from Phosphogypsum
by Xiaosheng Yu, Lijun Deng, Changpan Shen, Huiyong Li, Jingchao Li, Yijun Cao, Guoli Zhou and Guosheng Li
Separations 2025, 12(1), 7; https://doi.org/10.3390/separations12010007 (registering DOI) - 31 Dec 2024
Viewed by 111
Abstract
Phosphogypsum is an industrial byproduct that is limited in its high-value application due to the presence of dyeing impurities (such as organic matter and carbon black). The flotation method has been verified to be effective in separating these dyeing impurities from gypsum. In [...] Read more.
Phosphogypsum is an industrial byproduct that is limited in its high-value application due to the presence of dyeing impurities (such as organic matter and carbon black). The flotation method has been verified to be effective in separating these dyeing impurities from gypsum. In this study, microemulsion was used as the collector method of dyeing impurities for their separation from gypsum. The results of flotation tests showed that the microemulsion collector exhibited excellent collection capability and selectivity under natural pH conditions (pH = 1.5). With a microemulsion collector consumption of 400 g/t, purified gypsum of 65.1% whiteness, 95.74% yield, and 97.01% recovery was obtained. The purified gypsum of 65.1% whiteness, 95.74% yield, 97.01 recovery obtained by a used microemulsion collector amount of 400 g/t was better than using the same dosage of kerosene collector. The dispersion behavior of the microemulsion collector was studied by low-temperature transmission electron microscopy. The microemulsion collector demonstrated superior dispersibility, as it forms nano-oil droplets with an average size of 176.83 nm in the pulp, resolving issues associated with poor dispersibility observed in traditional kerosene collectors. Additionally, the nano-oil droplets effectively adsorbed onto the surface of dyeing impurities through hydrogen bonding, enhancing their hydrophobicity. Therefore, the microemulsion collector holds great potential for application in flotation whitening processes involving phosphogypsum. Full article
(This article belongs to the Special Issue Separation and Extraction Technology in Mineral Processing)
Show Figures

Figure 1

Figure 1
<p>SEM images of PG sample ((<b>a</b>) the microstructure of and (<b>b</b>) the elements mapping scanning of the PG raw ore.)</p>
Full article ">Figure 2
<p>XRD pattern of the PG sample.</p>
Full article ">Figure 3
<p>Flow chart of microemulsion preparation (<b>a</b>) and the reverse flotation tests (<b>b</b>).</p>
Full article ">Figure 3 Cont.
<p>Flow chart of microemulsion preparation (<b>a</b>) and the reverse flotation tests (<b>b</b>).</p>
Full article ">Figure 4
<p>TG-DTG pattern of the dyeing impurities.</p>
Full article ">Figure 5
<p>Size distribution of oil droplets generated by the dispersion of the microemulsion (<b>a</b>) and kerosene (<b>b</b>).</p>
Full article ">Figure 6
<p>Morphology of the microemulsion (<b>a</b>) and nano-oil droplets (<b>b</b>).</p>
Full article ">Figure 7
<p>Flotation results at different dosages of kerosene (<b>a</b>) and microemulsion (<b>b</b>) collectors.</p>
Full article ">Figure 8
<p>SEM images of raw PG (<b>a</b>) and gypsum concentrate (<b>b</b>).</p>
Full article ">Figure 9
<p>Contact angles of gypsum (<b>a</b>) and dyeing impurities (<b>b</b>) before and after being treated with kerosene and microemulsion.</p>
Full article ">Figure 10
<p>FT-IR results of gypsum and dyeing impurities treated by different collectors.</p>
Full article ">Figure 11
<p>XPS results of dyeing impurities treated by different collectors.</p>
Full article ">Figure 12
<p>Schematic diagram of the microemulsion collector enhancing flotation performance of PG.</p>
Full article ">
16 pages, 5499 KiB  
Article
Nanobubbles Adsorption and Its Role in Enhancing Fine Argentite Flotation
by Shunde Yan, Xihui Fang, Guanfei Zhao, Tingsheng Qiu and Kaiwei Ding
Molecules 2025, 30(1), 79; https://doi.org/10.3390/molecules30010079 (registering DOI) - 28 Dec 2024
Viewed by 326
Abstract
The efficient recovery of fine argentite from polymetallic lead–zinc (Pb–Zn) sulfide ore is challenging. This study investigated nanobubble (NB) adsorption on the argentite surface and its role in enhancing fine argentite flotation using various analytical techniques, including contact angle measurements, adsorption capacity analysis, [...] Read more.
The efficient recovery of fine argentite from polymetallic lead–zinc (Pb–Zn) sulfide ore is challenging. This study investigated nanobubble (NB) adsorption on the argentite surface and its role in enhancing fine argentite flotation using various analytical techniques, including contact angle measurements, adsorption capacity analysis, infrared spectroscopy, zeta potential measurements, turbidity tests, microscopic imaging, scanning electron microscopy, and flotation experiments. Results indicated that the NBs exhibited long-term stability and were adsorbed onto the argentite surface, thereby enhancing surface hydrophobicity, reducing electrostatic repulsion between fine argentite particles, and promoting particle agglomeration. Furthermore, the NBs formed a thin film on the argentite surface, which decreased the adsorption of sodium diethyldithiocarbamate. Microflotation tests confirmed that the introduction of NBs considerably enhanced the recovery of argentite using flotation technology. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The distribution state of argentite particle sizes.</p>
Full article ">Figure 2
<p>X-ray diffraction analysis results of argentite.</p>
Full article ">Figure 3
<p>DDTC molecular structure.</p>
Full article ">Figure 4
<p>ZJC-NM-200L Micro-Nano Bubble Generator.</p>
Full article ">Figure 5
<p>Adsorption standard curve of DDTC.</p>
Full article ">Figure 6
<p>The flowsheet of the microflotation test.</p>
Full article ">Figure 7
<p>The effect of cavitation time on the concentration and size of the nanobubbles. Cavitation time: (<b>A</b>) 3mim, 5min; (<b>B</b>) 7mim, 10min; (<b>C</b>) 12mim, 15min.</p>
Full article ">Figure 8
<p>The zeta potential of argentite under different conditions.</p>
Full article ">Figure 9
<p>Turbidity of the argentite particles under different pH.</p>
Full article ">Figure 10
<p>Turbidity of the argentite particles over different concentrations of DDTC.</p>
Full article ">Figure 11
<p>Optical microscopic observations of the fine argentite particles. (<b>A</b>) argentite + DI; (<b>B</b>) argentite + NBs; (<b>C</b>) argentite + DI + DDTC; (<b>D</b>) argentite + NBs + DDTC.</p>
Full article ">Figure 12
<p>SEM images of the argentite particles (<b>A</b>) before and (<b>B</b>) after NB adsorption.</p>
Full article ">Figure 13
<p>Contact angle of (<b>A</b>) argentite + DI; (<b>B</b>) argentite + NBs; (<b>C</b>) argentite + DI + DDTC; (<b>D</b>) argentite + NBs + DDTC.</p>
Full article ">Figure 14
<p>Infrared spectra of (<b>A</b>) DDTC and (<b>B</b>) argentite + DDTC.</p>
Full article ">Figure 15
<p>Effect of the concentration of DDTC on its adsorption on the surface of argentite.</p>
Full article ">Figure 16
<p>The recovery of argentite as a function of pH in the absence and presence of NBs (concentration of DDTC: 5 mg/L).</p>
Full article ">Figure 17
<p>The flotation recovery of argentite with particle sizes 38 μm to 74 μm varies with DDTC concentration (pH = 8).</p>
Full article ">Figure 18
<p>The flotation recovery of argentite as a function of DDTC concentration in the absence and presence of NBs (pH = 8).</p>
Full article ">
23 pages, 6361 KiB  
Article
Ozone/Thiosulfate-Assisted Leaching of Cu and Au from Old Flotation Tailings
by Stefan Trujić, Miroslav P. Popović, Vesna Conić, Miloš Janošević, Filip Alimpić, Dragoljub Bajić, Ana Milenković-Anđelković and Filip Abramović
Molecules 2025, 30(1), 69; https://doi.org/10.3390/molecules30010069 (registering DOI) - 27 Dec 2024
Viewed by 368
Abstract
The growing demand for metal production promotes the search for alternative sources and novel modalities in metallurgy. Flotation tailings are an important secondary mineral resource; however, they might pose a potential environmental threat due to containing toxic metals. Therefore, proper leaching reagent selection [...] Read more.
The growing demand for metal production promotes the search for alternative sources and novel modalities in metallurgy. Flotation tailings are an important secondary mineral resource; however, they might pose a potential environmental threat due to containing toxic metals. Therefore, proper leaching reagent selection is required. Ozone is an alternative oxidizing agent for metal leaching, as its use prevents contaminating product generation while increasing the noble metal extraction efficiency in the presence of complexing agents. In this study, the feasibility and efficiency of combining the use of thiosulfate and ozone for gold and silver extraction have been investigated as an eco-friendly alternative for recovery from flotation tailings. Two sets of samples from old flotation tailings of Copper Mine Bor (Serbia) were prepared and physico-chemically characterized, then treated in two experimental leaching procedures, followed by thorough XRD and SEM/EDS analyses of the products. It showed that after 1 h of leaching in a water medium at room temperature and a solid-to-liquid phase ratio of 1:4, 88.8% of Cu was obtained, while a high efficiency of Au extraction from solid residue (after Cu leaching) was attained (83.4%). The results suggest that ozone-assisted leaching mediated by Ca-thiosulfate can be an effective eco-friendly treatment for noble metals recovery from sulfide-oxide ores. Full article
Show Figures

Figure 1

Figure 1
<p>Separatory funnels for ozone leaching of Composite I and Composite II samples.</p>
Full article ">Figure 2
<p>XRD spectra of representative as-received samples of (<b>a</b>) Composite I and (<b>b</b>) Composite II from the old flotation tailings dump of the Copper Mine Bor.</p>
Full article ">Figure 2 Cont.
<p>XRD spectra of representative as-received samples of (<b>a</b>) Composite I and (<b>b</b>) Composite II from the old flotation tailings dump of the Copper Mine Bor.</p>
Full article ">Figure 3
<p>Separated phases after leaching of Composite I and Composite II samples.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD spectra of the dark precipitate of Composite I sample from experiment 2; (<b>b</b>) XRD spectra of the white precipitate of Composite I sample from experiment 2.</p>
Full article ">Figure 5
<p>(<b>a</b>) XRD spectra of the dark precipitate of Composite II sample from experiment 2; (<b>b</b>) XRD spectra of the white precipitate of Composite II sample from experiment 2.</p>
Full article ">Figure 6
<p>(<b>a</b>) EDS map-scan of the dark precipitate from Composite I sample and (<b>b</b>) SEM image of the sample (500× magnification) with labeled positions of EDS pointscan; (<b>c</b>) elemental distribution of mapscan.</p>
Full article ">Figure 7
<p>(<b>a</b>) EDS map-scan of the white precipitate from Composite I sample, and (<b>b</b>) SEM image of the sample (350× magnification) with labeled positions of EDS pointscan; (<b>c</b>) elemental distribution.</p>
Full article ">Figure 8
<p>(<b>a</b>) EDS map-scan of the dark precipitate from Composite II sample, and (<b>b</b>) SEM image of the sample (500× magnification) with labeled positions of EDS pointscan; (<b>c</b>) elemental distribution.</p>
Full article ">Figure 9
<p>(<b>a</b>) EDS map-scan of the white precipitate from Composite II sample, and (<b>b</b>) SEM image of the sample (500× magnification) with labeled positions of EDS pointscan; (<b>c</b>) elemental distribution.</p>
Full article ">Figure 10
<p>Pourbaix diagrams (E<sub>h</sub> vs. pH) at 25 °C of: (<b>a</b>) Au-O system, (<b>b</b>) Ag-O system, and (<b>c</b>) Fe-O system (assumed dissolved concentration of Au, Ag and Fe, respectively: 10<sup>−4</sup> molL<sup>−1</sup>) [<a href="#B46-molecules-30-00069" class="html-bibr">46</a>].</p>
Full article ">Figure 11
<p>Correlations between iron gain and redox potential increase during the leaching of (<b>a</b>) Composite I, (<b>b</b>) Composite II.</p>
Full article ">
19 pages, 3282 KiB  
Article
The Effect of Plasma Pretreatment on the Flotation of Lithium Aluminate and Gehlenite Using Light-Switchable Collectors
by Ali Zgheib, Maximilian Hans Fischer, Stéphanie Mireille Tsanang, Iliass El Hraoui, Shukang Zhang, Annett Wollmann, Alfred P. Weber, Ursula E. A. Fittschen, Thomas Schirmer and Andreas Schmidt
Separations 2024, 11(12), 362; https://doi.org/10.3390/separations11120362 - 23 Dec 2024
Viewed by 324
Abstract
The pyridinium phenolate punicine is a switchable molecule from Punica granatum. Depending on the pH, punicine exists as a cation, neutral molecule, anion, or dianion. In addition, punicine reacts to light, under the influence of which it forms radical species. We report [...] Read more.
The pyridinium phenolate punicine is a switchable molecule from Punica granatum. Depending on the pH, punicine exists as a cation, neutral molecule, anion, or dianion. In addition, punicine reacts to light, under the influence of which it forms radical species. We report on three punicine derivatives that possess an adamantyl, 2-methylnonyl, or heptadecyl substituent and on their performance in the flotation of lithium aluminate, an engineered artificial mineral (EnAM) for the recycling of lithium, e.g., from lithium-ion batteries. By optimizing the parameters: pH and light conditions (daylight, darkness), recovery rates of 92% of LiAlO2 are achieved. In all cases, the flotation of the gangue material gehlenite (Ca2Al[AlSiO7]) is suppressed. IR, the contact angle, zeta potential measurements, TG-MS, and PXRD confirm that the punicines interact with the surface of LiAlO2, which is covered by LiAl2(OH)7 after contact to water, resulting in a hydrophobization of the particle. The plasma pretreatment of the lithium aluminate has a significant influence on the flotation results and increases the recovery rates of lithium aluminate in blank tests by 58%. The oxidative plasma leads to a partial dehydratisation of the LiAl2(OH)7 and thus to a hydrophobization of the particles, while a reductive plasma causes a more hydrophilic particle surface. Full article
(This article belongs to the Special Issue Green Separation and Purification Technology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>TGA measurements of pure lithium aluminate and 5 min oxidative plasma-pretreated lithium aluminate samples.</p>
Full article ">Figure 2
<p>TG-MS of pure LiAlO<sub>2</sub> (<b>left</b>) and 5 min plasma pretreated LiAlO<sub>2</sub> (<b>right</b>).</p>
Full article ">Figure 3
<p>Comparison of the recovery rates of lithium aluminate at different light scenarios in the presence of the collectors <b>1</b>–<b>3</b> under identical conditions, i.e., 2.00 g of LiAlO<sub>2</sub>, 1 min of mixing (500 rpm) with 25 mL of distilled water, 60 µL (1 × 10<sup>−5</sup> M) of the corresponding collector, 1 min of mixing, 30 µL of frother, and another 1 min of mixing. The flotation was done for 3 min at an air flow rate of 32 mL/min using 250 mL distilled water in daylight at a pH of 10.9 ± 0.3.</p>
Full article ">Figure 4
<p>Comparison of the recovery rates of lithium aluminate at different pH values in the presence of the collectors <b>1</b>–<b>3</b> under identical conditions, i.e., 2.00 g of LiAlO<sub>2</sub>, 1 min of mixing (500 rpm) with 25 mL of distilled water, 60 µL (1 × 10<sup>−5</sup> M) of the corresponding collector, 1 min of mixing, 30 µL of frother, and another 1 min of mixing. The flotation was done for 3 min at an air flow rate of 32 mL/min using 250 mL distilled water in daylight.</p>
Full article ">Figure 5
<p>Single mineral flotation of lithium aluminate and gehlenite using punicine <b>3</b> as collector at pH 10.9 ± 0.3 for lithium aluminate and pH 8.3 ± 0.3 for gehlenite at different light scenarios. The conditioning was as follows: 120 µL collector, 1 min mixing, 30 µL frother, 1 min mixing, stirring speed 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and flotation time of 3 min.</p>
Full article ">Figure 6
<p>Single mineral flotation of plasma-pretreated lithium aluminate and the gehlenite-rich material without collector at pH 10.9 ± 0.3 for lithium aluminate and pH 8.3 ± 0.3 for gehlenite as a function of plasma treatment time. Conditioning was as follows: 30 µL frother, 1 min mixing, stirring speed 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and flotation time of 3 min.</p>
Full article ">Figure 7
<p>Single mineral flotation of the oxidative plasma pretreated lithium aluminate and gehlenite using punicine <b>3</b> as a collector at pH 10.9 ± 0.3 for lithium aluminate, and pH 8.3 for gehlenite as a function of plasma treatment time in min. The conditioning was as follows: 60 µL collector, 1 min mixing 30 µL frother, 1 min mixing, stirring speed of 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and applying a flotation time of 3 min.</p>
Full article ">Figure 8
<p>Single mineral flotation of the reductive plasma pretreated lithium aluminate in blank and using punicine <b>3</b> as a collector at pH 10.9 ± 0.3 as a function of plasma treatment time in min. The conditioning was as follows: 60 µL collector, 1 min mixing 30 µL frother, 1 min mixing, stirring speed of 500 rpm, air flow rate of 32 cm<sup>3</sup>/min using 250 mL distilled water, and flotation time of 3 min.</p>
Full article ">Figure 9
<p>Contact angle on LiAlO<sub>2</sub> before and after flotation with punicine <b>3</b> under identical conditions, i.e., 2.00 g of the mineral, 1 min mixing (500 rpm) with 25 mL distilled water, 120 µL (2 × 10<sup>−5</sup> M) of the corresponding collector, 1 min mixing, 30 µL foaming agent, and 1 min mixing again. Flotation was carried out for 3 min at an air flow rate of 32 mL/min with 250 mL of distilled water in daylight and a pH of 10.9 ± 0.3.</p>
Full article ">Figure 10
<p>Zeta potential of pure lithium aluminate, lithium aluminate with 7.2 µL of 42.5 mmol/L punicine <b>1</b>, lithium aluminate with 7.2 µL of 42.5 mmol/L punicine <b>2,</b> and lithium aluminate with 7.2 µL of 42.5 mmol/L punicine <b>3</b> in distilled water and under different pH values.</p>
Full article ">Figure 11
<p>FTIR spectra of LiAlO<sub>2</sub> in the presence of 60 µL 42.5 µmol/L punicine <b>3</b> with and without pretreatment with plasma.</p>
Full article ">Figure 12
<p>Zeta potential of pure and oxidative plasma-pretreated lithium aluminate, with and without 10 µL of 42.5 mmol/L punicine <b>3</b> at pH 10.9 ± 0.3.</p>
Full article ">Scheme 1
<p>Schematic representation of the switchability of punicine with respect to charges and radical status.</p>
Full article ">Scheme 2
<p>Intermolecular interactions and complexes of punicine.</p>
Full article ">Scheme 3
<p>Synthesis of the punicine derivatives.</p>
Full article ">Scheme 4
<p>Reactions of lithium aluminate surfaces with water.</p>
Full article ">Scheme 5
<p>Interactions of punicines with lithium aluminum surfaces.</p>
Full article ">
18 pages, 3016 KiB  
Article
Concepts for the Sustainable Hydrometallurgical Processing of End-of-Life Lithium Iron Phosphate (LFP) Batteries
by Marius Müller, Hüseyin Eren Obuz, Sebastian Keber, Firat Tekmanli, Luka Nils Mettke and Bengi Yagmurlu
Sustainability 2024, 16(24), 11267; https://doi.org/10.3390/su162411267 - 23 Dec 2024
Viewed by 555
Abstract
Lithium-ion batteries with an LFP cell chemistry are experiencing strong growth in the global battery market. Consequently, a process concept has been developed to recycle and recover critical raw materials, particularly graphite and lithium. The developed process concept consists of a thermal pretreatment [...] Read more.
Lithium-ion batteries with an LFP cell chemistry are experiencing strong growth in the global battery market. Consequently, a process concept has been developed to recycle and recover critical raw materials, particularly graphite and lithium. The developed process concept consists of a thermal pretreatment to remove organic solvents and binders, flotation for anode–cathode separation, and hydrometallurgical processes for product recovery. It has been shown that a pretreatment step is necessary for efficient flotation. By increasing the thermal treatment temperatures up to 450 °C, recovery rates of up to 73% are achieved. Similar positive effects are observed with leaching, where leaching efficiencies increase with higher treatment temperatures up to 400 °C. The results indicate that the thermal treatment of the black mass significantly influences both flotation and hydrometallurgical processes. Full article
Show Figures

Figure 1

Figure 1
<p>Generalized LFP processing methods.</p>
Full article ">Figure 2
<p>XRD diffractogram of the initial LFP sample.</p>
Full article ">Figure 3
<p>TGA data of the LFP black mass and possible reactions occurring during the thermal treatment [<a href="#B21-sustainability-16-11267" class="html-bibr">21</a>,<a href="#B22-sustainability-16-11267" class="html-bibr">22</a>,<a href="#B23-sustainability-16-11267" class="html-bibr">23</a>].</p>
Full article ">Figure 4
<p>Flotation behavior in the dependency of the thermal treatment. (<b>a</b>) Improper foam formation (thermal treatment at 350 °C; flotation parameters: cell volume, 0.125 L; flotation time, 10 min; conditioning time—collector, 3 min; conditioning time—foamer, 2 min; collector concentration, 350 g/t; foamer concentration, 150 g/t). (<b>b</b>) Foam after incomplete flotation. (<b>c</b>) Proper foam formation (thermal treatment at 450 °C; flotation parameters: cell volume, 0.125 L; flotation time, 10 min; conditioning time—collector, 3 min; conditioning time—foamer, 2 min; collector concentration, 350 g/t; foamer concentration, 150 g/t). (<b>d</b>) Cathode concentration after the flotation.</p>
Full article ">Figure 5
<p>Influence of the roasting temperature during the thermal pretreatment for 60 min on the graphite flotation efficiency (flotation parameters: cell volume, 0.125 L; flotation time, 10 min; conditioning time—collector, 3 min; conditioning time—foamer, 2 min; collector concentration, 350 g/t; foamer concentration, 150 g/t).</p>
Full article ">Figure 6
<p>Leaching yields with different stoichiometric ratios (STC) of acid concentrations (leaching conditions: solid/liquid ratio, 1/10; leaching time, 60 min; leaching temperature, 30 °C).</p>
Full article ">Figure 7
<p>Effect of oxidative media (H<sub>2</sub>O<sub>2</sub>) on the leaching yields (leaching conditions: stoichiometric amount of H<sub>2</sub>SO<sub>4</sub>; solid/liquid ratio, 1/10; leaching time, 60 min; leaching temperature, 30 °C).</p>
Full article ">Figure 8
<p>Effect of different roasting temperatures (roasting in a muffle furnace under atmospheric conditions with a heating rate of 10 K/min) on the leaching (leaching conditions: stoichiometric amount of H<sub>2</sub>SO<sub>4</sub>; solid/liquid ratio, 1/10; leaching time, 60 min; leaching temperature, 30 °C).</p>
Full article ">Figure 9
<p>The pH vs. precipitation yields of the constituent cathode metals.</p>
Full article ">Figure 10
<p>Flow sheet for the proposed process for leaching LFP black mass using phosphoric acid.</p>
Full article ">Figure 11
<p>Diffractogram of the precipitation product.</p>
Full article ">Figure 12
<p>Possible process routes for the recycling of LFP batteries.</p>
Full article ">
12 pages, 4751 KiB  
Article
Investigation of Appropriate Collector Selection for Hematite Removal from Pyrolusite and the Adsorption Mechanism on the Crystal Surface
by Yuhang Shi, Nan Nan, Baoxu Song, Fangyuan Ma, Jiquan Han, Enming Huang, Shuai Wang, Guang Yang and Lan Zhou
Minerals 2024, 14(12), 1300; https://doi.org/10.3390/min14121300 - 23 Dec 2024
Viewed by 374
Abstract
This study examined the appropriate hematite (Fe2O3) collector for the concentration of pyrolusite (MnO2) in a reverse flotation. Actual ore flotation studies were performed to determine how sodium oleate, sodium dodecyl sulfonate, and oxidized paraffin soap affect [...] Read more.
This study examined the appropriate hematite (Fe2O3) collector for the concentration of pyrolusite (MnO2) in a reverse flotation. Actual ore flotation studies were performed to determine how sodium oleate, sodium dodecyl sulfonate, and oxidized paraffin soap affect hematite removal during reverse flotation of pyrolusite ore. In order to explore the flotation mechanism, simulation experiments were carried out. Firstly, the crystal models of pyrolusite and hematite were established. Then, in order to verify the reliability of the simulation results, the simulated XRD spectra of the crystal model were compared with the measured spectra. Finally, density functional theory and molecular dynamics modeling were used to study the interaction between collector molecules and mineral surfaces. The flotation test results show that oxidized paraffin soap is the best hematite collector and promotes its flotation, removing iron from pyrolusite. Molecular dynamics simulations and density functional theory show that the three collectors (oxidized paraffin soap, sodium oleate, and sodium dodecyl sulfonate) have a much stronger interaction with hematite than with pyrolusite. Therefore, it is possible to separate pyrolusite and hematite through flotation. The simulation results also show that oxidized paraffin soap has the highest adsorption strength and selectivity for hematite. This characteristic makes oxidized paraffin soap an excellent collector for effectively removing hematite from pyrolusite in the reverse flotation process. Full article
(This article belongs to the Special Issue Desorption and/or Reuse of Collectors in Mineral Flotation)
Show Figures

Figure 1

Figure 1
<p>Flow chart of the flotation test.</p>
Full article ">Figure 2
<p>The effect of collector dosage on the hematite flotation efficiency.</p>
Full article ">Figure 3
<p>Original unit cell model of pyrolusite (<b>left</b>) and hematite (<b>right</b>).</p>
Full article ">Figure 4
<p>Convergence test of pyrolusite unit cell.</p>
Full article ">Figure 5
<p>Convergence test of hematite unit cell.</p>
Full article ">Figure 6
<p>Comparison between simulation results and experimental values (left—pyrolusite; right—hematite).</p>
Full article ">Figure 7
<p>The density of pyrolusite states.</p>
Full article ">Figure 8
<p>The density of hematite states.</p>
Full article ">Figure 9
<p>Optimized structures of collector anions (SDS, OPS, and NaOl, respectively).</p>
Full article ">Figure 10
<p>Interaction configurations of SDS–hematite (<b>a</b>) and SDS–pyrolusite (<b>b</b>).</p>
Full article ">Figure 11
<p>Interaction configurations of NaOl–hematite (<b>a</b>) and NaOl–pyrolusite (<b>b</b>).</p>
Full article ">Figure 12
<p>Interaction configurations of OPS–hematite (<b>a</b>) and OPS–pyrolusite (<b>b</b>).</p>
Full article ">Figure 13
<p>Adsorption energy of collector on mineral surface.</p>
Full article ">
21 pages, 5031 KiB  
Article
Interaction Between Nonionic Surfactants and Alkyl Amidoamine Cationic Collector in the Reverse Flotation of Iron Ore
by José Tadeu Gouvêa Junior, Flávia Paulucci Cianga Silvas, Christian Lariguet Taques Bittencourt, Vantuir Jorge de Morais, Ali Asimi Neisiani and Laurindo de Salles Leal Filho
Minerals 2024, 14(12), 1298; https://doi.org/10.3390/min14121298 - 22 Dec 2024
Viewed by 710
Abstract
This paper evaluates the performance of four ethoxylated nonionic surfactants (nonyl phenol vs. C13 alcohols) to act as ancillary collectors with Alkyl Amidoamine (AAA) in the reverse flotation of quartz at pH8 to concentrate iron ores. Compared to 100% AAA, the blend [...] Read more.
This paper evaluates the performance of four ethoxylated nonionic surfactants (nonyl phenol vs. C13 alcohols) to act as ancillary collectors with Alkyl Amidoamine (AAA) in the reverse flotation of quartz at pH8 to concentrate iron ores. Compared to 100% AAA, the blend composed of 80% AAA (Flotinor®5530) plus 20% of isotridecyl alcohol ethoxylated with five groups of ethylene oxide (DP-210 RO) improved quartz recovery (from 54% to 63%, p < 0.05) by increasing contact angle (from 55° to 56°, p < 0.05) and decreasing induction time (26 ms to 23 ms, p < 0.05). Compared to 100% AAA (200 g/t), the blend (160 g/t of AAA + 40 g/t of DP-210 RO) improved the flotation performance of iron ore, yielding richer hematite concentrate (65.3% Fe × 61.4% Fe) and less contaminated with quartz (4% SiO2 × 10.2% SiO2), coupled with an increase in Fe recovery from 79.8% × 81.6% in the sunken product as well as SiO2 recovery from 91.7% to 96.9% in the froth. Results from zeta potential, the hydrodynamic diameter of reagent droplets, and the surface tension of the solution provide insights into the synergism between AAA and DP-210 RO. Full article
Show Figures

Figure 1

Figure 1
<p>Adsorption of surfactants at air/solution (<b>I</b>) and mineral/solution (<b>II</b>) interface.</p>
Full article ">Figure 2
<p>Illustration of the rationale used to assess (<b>I</b>) HLB of surfactants (<b>II</b>) and Griffin’s HLB scale.</p>
Full article ">Figure 3
<p>Particle size distribution of the quartz and iron ore.</p>
Full article ">Figure 4
<p>Molecular structures of (<b>a</b>) Nonyl phenol with two EO groups, (<b>b</b>) Nonyl phenol with four EO groups, (<b>c</b>) Isotridecyl alcohol with three EO groups, (<b>d</b>) Isotridecyl alcohol with five EO groups, (<b>e</b>) N-[3-(Dimethylamino)propyl]dodecanamide. Atom color code: White (hydrogen), Red (oxygen), Blue (nitrogen), Gray (carbon) [<a href="#B52-minerals-14-01298" class="html-bibr">52</a>].</p>
Full article ">Figure 5
<p>Visual aspect of aqueous mixtures of AAA plus NS (1% <span class="html-italic">w</span>/<span class="html-italic">w</span>). Where A—100% AAA; B—80% AAA + 20% Ultranex<sup>®</sup>NP18; C—60% AAA + 40% Ultranex<sup>®</sup>NP18; D—80% AAA + 20% Arkopal<sup>®</sup>N040; E—60% AAA + 40% Arkopal<sup>®</sup>N040; F—80% AAA + 20% DP-209RO; G—60% AAA + 40% DP-209RO; H—80% AAA + 20% DP-210RO; I—60% AAA + 40% DP-210RO.</p>
Full article ">Figure 6
<p>Quartz recovery (R) at pH 8 with 20 g/t of Flotinor<sup>®</sup>5530 (AAA) vs. blends composed of 80% AAA (16 g/t) plus NIS (4 g/t) vs. blends of 60% AAA (12 g/t) plus 40% of NIS (8 g/t) at 20° ± 1 °C. Where the red dash compares the quartz recovery of the main collector and the mixtures.</p>
Full article ">Figure 7
<p>Distribution of hydrodynamic diameter (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>H</mi> </mrow> </msub> </mrow> </semantics></math>) of aggregates formed by AAA species versus AAA species plus NS suspended in electrolyte aqueous solution (1 × 10<sup>−3</sup> M NaCl) at pH 8 and 20 °C.</p>
Full article ">Figure 8
<p>Zeta potential of quartz vs. pH and droplets/aggregates formed by AAA species and 80%AAA + 20%NS (75 mg/L) at pH 8, 20 °C.</p>
Full article ">Figure 9
<p>Zeta potential of quartz at pH 8 in the absence (A) and presence of AAA (B) and 80%AAA + 20%NS (C).</p>
Full article ">Figure 10
<p>Contact angle (<b>I</b>) and Induction time (<b>II</b>) of quartz in the presence of: A—water; B—100% AAA; C—80% AAA + 20% DP 210 RO.</p>
Full article ">Figure 11
<p>Work of adhesion (<b>I</b>) and Free energy of attachment (<b>II</b>) of quartz in the presence of A—water; B—100% AAA; C—80% AAA + 20% DP 210 RO.</p>
Full article ">
22 pages, 6794 KiB  
Article
Metal Sulfide Nanoparticles as Sphalerite Surface Activators to Improve Zinc Recovery Through Flotation Process
by Delia Monserrat Ávila-Márquez, Alien Blanco-Flores, Helen Paola Toledo-Jaldin, Maribel González Torres, Alfredo Rafael Vilchis-Nestor, Iván Alejandro Reyes Domínguez and Ramiro de Aquino García
Separations 2024, 11(12), 358; https://doi.org/10.3390/separations11120358 - 22 Dec 2024
Viewed by 501
Abstract
CuS nanoparticles (Np) were synthesized and deposited on synthetic sphalerite (SP) using two different methods. Two nanoparticle products were obtained on the surface of SP, Np1 and Np2, resulting in two active materials (Np1-S [...] Read more.
CuS nanoparticles (Np) were synthesized and deposited on synthetic sphalerite (SP) using two different methods. Two nanoparticle products were obtained on the surface of SP, Np1 and Np2, resulting in two active materials (Np1-SP and Np2-SP) with specific characteristics. Nanoparticles and active materials were characterized by TEM, XRD, SEM, and XPS. The collectors PAX and SIPX were adsorbed on Np1-SP and Np2-SP to determine the adsorption capacity. Method 1 provides a higher quantity of nanoparticles on SP, which allows for the adsorption of a higher amount of SIPX. Method 1 was used to deposit nanoparticles on two natural sphalerites (SN) with different iron contents. SN, unlike SP, can be used to test nanoparticle activation results in microflotation experiments. SN was activated with nanoparticles (Np1-SN) and using the traditional method (Cu-SN). The recovery of 75% of zinc using the microflotation process suggests that the hydrophobicity of Np1-SN is higher than that of Cu-SN. Nanoparticles improve the hydrophobicity of SN compared to the traditional activation used in the mining industry. These results suggest that using nanoparticles is an excellent option to activate minerals in flotation processes, decreasing the consumption of reagents and helping to mitigate negative impacts on the environment. Full article
(This article belongs to the Section Materials in Separation Science)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Activation process with the deposition of CuS nanoparticles synthesized by (<b>a</b>) M1 and (<b>b</b>) by M2.</p>
Full article ">Figure 2
<p>XRD patterns of <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> and <b>Np2-S<sub>P</sub></b>.</p>
Full article ">Figure 3
<p>SEM micrographs of <b>Np1-S<sub>P</sub></b> (<b>c</b>,<b>d</b>) and <b>Np2-S<sub>P</sub></b> (<b>a</b>,<b>b</b>,<b>e</b>,<b>f</b>).</p>
Full article ">Figure 4
<p>EDS spectra of <b>Np1-S<sub>P</sub></b> using Na<sub>2</sub>S (<b>a</b>) and <b>Np2-S<sub>P</sub></b> using thiourea (<b>b</b>).</p>
Full article ">Figure 5
<p>TEM micrographs of <b>Np1-S<sub>P</sub></b> (<b>a</b>) and particle size histogram (<b>b</b>) and <b>Np2-S<sub>P</sub></b> (<b>c</b>) and particle size histogram (<b>d</b>).</p>
Full article ">Figure 6
<p>Adsorption kinetics of (<b>a</b>) SIPX and (<b>b</b>) PAX collectors onto <b>Np1-S<sub>P</sub></b>, <b>Np2-S<sub>P</sub></b>, <b>Cu-S<sub>P</sub></b> and (<b>c</b>) adsorption of the SIPX collector onto <b>Np1</b>, <b>Np2</b>, <b>Np1-S<sub>P</sub></b>, <b>Np2-S<sub>P</sub></b>, <b>Cu-S<sub>P</sub></b>.</p>
Full article ">Figure 7
<p>XPS survey spectra of <b>Np1-S<sub>P</sub></b> (<b>a</b>), spectra of Zn 2p (<b>b</b>), Zn2p comparation for <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> (<b>c</b>), S 2p (<b>d</b>), S2p comparation for <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> (<b>e</b>), Cu 2p (<b>f</b>) and O1s (<b>g</b>).</p>
Full article ">Figure 7 Cont.
<p>XPS survey spectra of <b>Np1-S<sub>P</sub></b> (<b>a</b>), spectra of Zn 2p (<b>b</b>), Zn2p comparation for <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> (<b>c</b>), S 2p (<b>d</b>), S2p comparation for <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> (<b>e</b>), Cu 2p (<b>f</b>) and O1s (<b>g</b>).</p>
Full article ">Figure 7 Cont.
<p>XPS survey spectra of <b>Np1-S<sub>P</sub></b> (<b>a</b>), spectra of Zn 2p (<b>b</b>), Zn2p comparation for <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> (<b>c</b>), S 2p (<b>d</b>), S2p comparation for <b>S<sub>P</sub></b> and <b>Np1-S<sub>P</sub></b> (<b>e</b>), Cu 2p (<b>f</b>) and O1s (<b>g</b>).</p>
Full article ">Figure 8
<p>Mechanism to deposition of <b>Np1</b> and collector adsorption onto <b>Np1-S<sub>P</sub></b>.</p>
Full article ">Figure 9
<p>Microflotation results of traditionally activated materials and CuS nanoparticles.</p>
Full article ">
15 pages, 3976 KiB  
Article
Mechanism of Efficient Smithsonite Flotation with a Ternary Composite Collector Under Sulfur-Free Conditions
by Rui Li, Yanhai Shao, Jinhui Li, Chenjie Liu, Hongqin Chen, Xiao Meng and Xinru Jia
Molecules 2024, 29(24), 6014; https://doi.org/10.3390/molecules29246014 - 20 Dec 2024
Viewed by 389
Abstract
The increasing demand for zinc resources and the declining availability of sulfide zinc ore reserves have made the efficient utilization of zinc oxide a topic of considerable interest. In this study, a ternary composite collector ABN (Al-BHA-NaOL system) was applied to the direct [...] Read more.
The increasing demand for zinc resources and the declining availability of sulfide zinc ore reserves have made the efficient utilization of zinc oxide a topic of considerable interest. In this study, a ternary composite collector ABN (Al-BHA-NaOL system) was applied to the direct flotation of smithsonite. Micro-flotation studies showed that at pH 9, ABN exhibited better adsorption on smithsonite, achieving a recovery rate of 80.62%. Visual MINTEQ 3.1 and zeta potential analysis confirmed that ABN predominantly reacted with Zn(OH)2(aq) on the surface of smithsonite. Furthermore, X-ray photoelectron spectroscopy (XPS) analysis results elucidated the formation of Al-O bonds through chemical adsorption on the smithsonite surface. Additionally, powder contact angle measurements indicated that ABN enhances the surface contact angle of smithsonite. These results illuminate that ABN is adsorbed by reacting with O sites on hydroxylated metal ions on the smithsonite surface, with Al serving as the adsorption center, thereby achieving separation and purification. Due to ABN’s adsorption characteristics, smithsonite can achieve efficient and clean direct flotation recovery without sulfidization. Full article
Show Figures

Figure 1

Figure 1
<p>Recovery rate of smithsonite under different experimental condition (<b>a</b>) different reagent regimes recovery rate of smithsonite; (<b>b</b>) different reagent regimes recovery rate of quartz; (<b>c</b>) dosage experiment (<b>d</b>) composition diagrams of smithsonite solution under different pH by Visual MITEQ3.1.</p>
Full article ">Figure 2
<p>Effect of pH on the zeta potential and composition diagram of smithsonite solution by Visual MINTEQ analysis.</p>
Full article ">Figure 3
<p>Full XPS spectrum of the smithsonite samples.</p>
Full article ">Figure 4
<p>Fine spectrum C 1s spectra of the smithsonite surface under different conditions. (<b>a</b>) Smithsonite and (<b>b</b>) smithsonite + ABN.</p>
Full article ">Figure 5
<p>Fine spectrum N 1s spectra of the smithsonite surface after ABN treatment.</p>
Full article ">Figure 6
<p>Fine spectrum of the smithsonite surface under different conditions; (<b>a</b>) O 1s smithsonite; (<b>b</b>) O 1s smithsonite + ABN; (<b>c</b>) Zn 2p smithsonite; (<b>d</b>) Zn 2p smithsonite + ABN.</p>
Full article ">Figure 7
<p>Powder contact angle test curve.</p>
Full article ">Figure 8
<p>Mechanism diagram of ABN on smithsonite surface.</p>
Full article ">Figure 9
<p>XRD spectrum of pure smithsonite.</p>
Full article ">Figure 10
<p>Chemical structure diagram of ABN.</p>
Full article ">Figure 11
<p>Flotation flowsheet.</p>
Full article ">
17 pages, 3508 KiB  
Article
Parasite Screening in Wild Passerines: Enhancing Diagnostic Approaches in Wildlife Rehabilitation Centers
by Catarina Ferreira Rebelo, Alicia Carrero Ruiz, Alberto Alvarado-Piqueras, Fernando González González and Luís Madeira de Carvalho
Animals 2024, 14(24), 3664; https://doi.org/10.3390/ani14243664 - 19 Dec 2024
Viewed by 363
Abstract
The order Passeriformes is the richest and most abundant group of birds, but despite numerous parasites being identified in wild birds, this order has received limited focus. This study analyzed 17 passerines admitted to the Grupo de Rehabilitación de la Fauna Autóctona y [...] Read more.
The order Passeriformes is the richest and most abundant group of birds, but despite numerous parasites being identified in wild birds, this order has received limited focus. This study analyzed 17 passerines admitted to the Grupo de Rehabilitación de la Fauna Autóctona y su Hábitat (GREFA), a wildlife rehabilitation center in Spain, during October to December 2022. Necropsies were conducted to determine the presence of parasites, and intestinal contents were analyzed using fecal smear, flotation and sedimentation techniques and the McMaster method. Sixteen passerines (94.1%) were positive for parasites. Identified species included Monojoubertia microhylla and the genera Ornithonyssus sp., Diplotriaena spp., Serratospiculum sp., Porrocaecum sp., Capillaria spp., Syngamus sp., Strongyloides sp. and Brachylecithum sp., besides cestodes and coccidia. The comparative analysis of parasitological diagnostic techniques showed that the Willis flotation technique was effective for detecting coccidia. However, to obtain more accurate results for other parasites, it is important to complement this technique with direct examination or sedimentation techniques. Among the 12 passerines positive for coccidia, oocyst counts per gram of intestinal contents ranged from 100 to 30,450, with a median of 7350. This study provides valuable insights into the parasitic fauna of Passeriformes, serving as a cornerstone for future research and enhancing our understanding of these avian species. Full article
(This article belongs to the Special Issue Advances in the Diagnosis of Parasitic Infections in Animals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of passerine sample by family.</p>
Full article ">Figure 2
<p>Distribution of the detected parasitic groups by passerine family.</p>
Full article ">Figure 3
<p>Female (<b>on the left</b>), male (<b>on the right</b>, below) and tritonymph (<b>on the right</b>, above) of <span class="html-italic">Monojoubertia microhylla</span> on a common chaffinch (<span class="html-italic">Fringilla coelebs</span>).</p>
Full article ">Figure 4
<p>Dorsal view (<b>on the left</b>) and ventral view (<b>on the right</b>) of <span class="html-italic">Ornithonyssus</span> sp. on a Eurasian blackcap (<span class="html-italic">Sylvia atricapilla</span>).</p>
Full article ">Figure 5
<p>Parasites of the genus <span class="html-italic">Diplotriaena</span> sp. in the air sacs of a Eurasian blackcap (<span class="html-italic">Sylvia atricapilla</span>).</p>
Full article ">Figure 6
<p>(<b>A</b>) Anterior end of <span class="html-italic">Diplotriaena</span> sp., where the trident can be observed (arrow). (<b>B</b>) Spicules at the posterior end of a <span class="html-italic">Diplotriaena</span> sp. male.</p>
Full article ">Figure 7
<p>(<b>A</b>) Anterior and posterior ends of <span class="html-italic">Serratospiculum</span> sp. from a song thrush (<span class="html-italic">Turdus philomelos</span>). (<b>B</b>) Posterior end of Porrocaecum sp. from a Eurasian blackbird (<span class="html-italic">Turdus merula</span>).</p>
Full article ">Figure 8
<p>Immature parasitic forms detected by coprological techniques: (<b>A</b>) Unsporulated oocyst of coccidia. (<b>B</b>–<b>D</b>) Eggs of <span class="html-italic">Diplotriaena</span> spp. (<b>E</b>,<b>F</b>) Eggs of <span class="html-italic">Capillaria</span> spp. (<b>G</b>) Egg of <span class="html-italic">Porrocaecum</span> sp. (<b>H</b>) Egg of <span class="html-italic">Syngamus</span> sp. (<b>I</b>) Egg of <span class="html-italic">Strongyloides</span> sp. (<b>J</b>) Egg of <span class="html-italic">Brachylecithum</span> sp. (<b>K</b>,<b>L</b>) Eggs of cestodes.</p>
Full article ">Figure 9
<p>Comparison of qualitative coprological techniques in the detection of positive parasitological samples.</p>
Full article ">Figure 10
<p>Results of coccidial oocyst counting using the McMaster method.</p>
Full article ">
20 pages, 1289 KiB  
Article
The Use of Diatomite-Based Composites for the Immobilization of Toxic Heavy Metals in Industrial Wastes Using Post-Flotation Sediment as an Example
by Krzysztof Gondek, Agnieszka Baran, Patrycja Boguta and Małgorzata Bołdak
Materials 2024, 17(24), 6174; https://doi.org/10.3390/ma17246174 - 17 Dec 2024
Viewed by 650
Abstract
Composite materials based on diatomite (DT) with the addition of biochar (BC), dolomite (DL), and bentonite (BN) were developed. The effect of chemical modification on the chemical structure of the resulting composites was investigated, and their influence on heavy metal immobilization and the [...] Read more.
Composite materials based on diatomite (DT) with the addition of biochar (BC), dolomite (DL), and bentonite (BN) were developed. The effect of chemical modification on the chemical structure of the resulting composites was investigated, and their influence on heavy metal immobilization and the ecotoxicity of post-flotation sediments was evaluated. It was demonstrated that the chemical modifications resulted in notable alterations to the chemical properties of the composites compared to pure DT and mixtures of DT with BC, DL, and BN. An increase in negative charge was observed in all variants. The addition of BC introduced valuable chemically and thermally resistant organic components into the composite. Among the chemical modifications, composites with the addition of perlite exhibited the lowest values of negative surface charge, which was attributed to the dissolution and transformation of silicon compounds and traces of kaolinite during their initial etching with sodium hydroxide. The materials exhibited varying efficiencies in metal immobilization, which is determined by both the type of DT additive and the type of chemical modification applied. The greatest efficacy in reducing the mobility of heavy metals was observed in the PFS with the addition of DT and BC without modification and with the addition of DT and BC after the modification of H2SO4 and H2O2: Cd 8% and 6%; Cr 71% and 69%; Cu 12% and 14%; Ni 10% and Zn 15%; and 4% and 5%. In addition, for Zn and Pb, good efficacy in reducing the content of mobile forms of these elements was observed for DT and DL without appropriate modification: 4% and 20%. The highest reduction in ecotoxicity was observed in the PFS with the addition of DT and BC, followed by BN and DL, which demonstrated comparable efficacy to materials with DT and BN. Full article
(This article belongs to the Special Issue Advances in Polymers and Functionalized Materials in the Environment)
Show Figures

Figure 1

Figure 1
<p>Distribution of surface charge (Q) derived from functional groups at increasing pH values.</p>
Full article ">Figure 2
<p>The distribution function of the apparent dissociation constants in a wide range of pK.</p>
Full article ">Figure 3
<p>The FI–IR spectra of the diatomite (DT) and DT composites with (<b>a</b>) biochar (BC), (<b>b</b>) dolomite (DL), and (<b>c</b>) bentonite (BN).</p>
Full article ">
16 pages, 10023 KiB  
Article
Silicon Extraction from a Diamond Wire Saw Silicon Slurry with Flotation and the Flotation Interface Behavior
by Lin Zhu, Dandan Wu, Shicong Yang, Keqiang Xie, Kuixian Wei and Wenhui Ma
Molecules 2024, 29(24), 5916; https://doi.org/10.3390/molecules29245916 - 15 Dec 2024
Viewed by 625
Abstract
Diamond wire saw silicon slurry (DWSSS) is a waste resource produced during the process of solar-grade silicon wafer preparation with diamond wire sawing. The DWSSS contains 6N grade high-purity silicon and offers a promising resource for high-purity silicon recycling. The current process for [...] Read more.
Diamond wire saw silicon slurry (DWSSS) is a waste resource produced during the process of solar-grade silicon wafer preparation with diamond wire sawing. The DWSSS contains 6N grade high-purity silicon and offers a promising resource for high-purity silicon recycling. The current process for silicon extraction recovery from DWSSS presents the disadvantages of lower recovery and secondary pollution. This study focuses on the original DWSSS as the target and proposes flotation for efficiently extracting silicon. The experimental results indicate that the maximal recovery of silicon reached 98.2% under the condition of a dodecylamine (DDA) dosage of 0.6 g·L−1 and natural pH conditions within 24 min, and the flotation conforms to the first-order rate model. Moreover, the mechanism of the interface behavior between DWSSS and DDA revealed that DDA is adsorbed on the surface of silicon though adsorption, and the floatability of silicon is improved. The DFT calculation indicates that DDA can be spontaneously adsorbed with the silicon. The present study demonstrates that flotation is an efficient method for extracting silicon from DWSSS and provides an available option for silicon recovery. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Experimental phenomena of different types of collectors: (<b>a</b>) kerosene; (<b>b</b>) sodium sulfide; (<b>c</b>) DDA; and (<b>d</b>) DDA and kerosene.</p>
Full article ">Figure 2
<p>Experimental results of DWSSS flotation: (<b>a</b>) recovery of silicon under different DDA dosages; and (<b>b</b>) recovery of silicon under different pH levels.</p>
Full article ">Figure 3
<p>The fitting results of the first-order classical flotation rate model for silicon recovery with different conditions: (<b>a1</b>) the recovery of silicon under different DDA dosages; (<b>a2</b>–<b>a6</b>) the fitting results under different DDA dosages; (<b>b1</b>) the recovery of silicon under different pH levels; and (<b>b2</b>–<b>b6</b>) the fitting results under different pH levels.</p>
Full article ">Figure 4
<p>(<b>a</b>) Effect of DDA dosage on maximum recovery and flotation rate constant at natural pH; (<b>b</b>) effect of pH on maximum recovery and flotation rate constant at DDA dosage of 0.6g·L<sup>−1</sup>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Effect of DDA on surface zeta potential of silicon; (<b>b</b>) contact angle of silicon before and after DDA addition.</p>
Full article ">Figure 6
<p>Infrared spectrum results: (<b>a</b>) DWSSS; (<b>b</b>) DDA; (<b>c</b>) DWSSS-DDA.</p>
Full article ">Figure 7
<p>(<b>a</b>) Three-dimensional height morphology and roughness changes in DWSSS before and after DDA addition; (<b>b1,b2</b>) two-dimensional geometric morphology and cross-sectional height of DWSSS; (<b>c1,c2</b>) two-dimensional geometric morphology and cross-sectional height of DWSSS-DDA.</p>
Full article ">Figure 8
<p>(<b>a1</b>) SEM image of DWSSS before DDA addition; (<b>a2</b>) EDS result of DWSSS before DDA addition; (<b>b1</b>,<b>b2</b>) SEM image of DWSSS after DDA addition; (<b>b2</b>) EDS result of DWSSS after DDA addition.</p>
Full article ">Figure 9
<p>(<b>a</b>) Best adsorption position of DDA on Si (111) surface in aqueous solution system; (<b>b</b>) density map of adsorption region of DDA on Si (111) surface in aqueous solution.</p>
Full article ">Figure 10
<p>(<b>a</b>) Analysis of Si (111) and DDA adsorption sites. Si (111) and DDA possible adsorption position stereogram and top view: (<b>b</b>) Site 1, (<b>c</b>) Site 2, (<b>d</b>) Site 2, (<b>e</b>) Site 2, (<b>f</b>) Site 2.</p>
Full article ">Figure 11
<p>DFT calculation model of DDA on Si (111): (<b>a</b>) before geometric optimization; (<b>b</b>) after geometric optimization.</p>
Full article ">Figure 12
<p>(<b>a</b>) Section position of differential charge density diagram of Si-DDA system; (<b>b</b>,<b>c</b>) differential charge density diagram of Si-DDA system.</p>
Full article ">Figure 13
<p>Mulliken population of band in Si (111) surface after DDA adsorption.</p>
Full article ">Figure 14
<p>Density of states for DDA, Si (111) surface, and Si-DDA.</p>
Full article ">Figure 15
<p>(<b>a</b>) Image of original DWSSS; (<b>b</b>) particle size distribution of DWSSS; (<b>c</b>) XRD phase analysis result of DWSSS.</p>
Full article ">Figure 16
<p>Experimental procedure of flotation: (<b>a</b>) flow chart of flotation experiment; (<b>b</b>) diagram of flotation process.</p>
Full article ">
20 pages, 4697 KiB  
Article
Utilization of Copper Flotation Tailings in Geopolymer Materials Based on Zeolite and Fly Ash
by Marija Štulović, Dragana Radovanović, Jelena Dikić, Nataša Gajić, Jovana Djokić, Željko Kamberović and Sanja Jevtić
Materials 2024, 17(24), 6115; https://doi.org/10.3390/ma17246115 - 14 Dec 2024
Viewed by 390
Abstract
Copper flotation tailings (FTs), resulting from the separation and beneficiation processes of ores, are a significant source of environmental pollution (acid mine drainage, toxic elements leaching, and dust generation). The most common disposal method for this industrial waste is dumping. However, due to [...] Read more.
Copper flotation tailings (FTs), resulting from the separation and beneficiation processes of ores, are a significant source of environmental pollution (acid mine drainage, toxic elements leaching, and dust generation). The most common disposal method for this industrial waste is dumping. However, due to their favorable physical and chemical properties—the high content of aluminosilicate minerals (60–90%)—flotation tailings can be effectively treated and reused through geopolymerization technology, thereby adding value to this waste. The objective of this study was to evaluate the potential of utilizing the geopolymerization of FTs to produce sustainable materials. Geopolymers based on natural zeolite (NZ), sodium-modified natural zeolite (NaZ), and fly ash (FA) were prepared using 20%, 35%, and 50% of FTs, activated with a 10 M NaOH solution. The study investigated the influence of Ca/Si, Si/Al, and Na/Al molar ratios on the structural, thermal, and mechanical properties (XRD, TG/DTG and unconfined compressive strength, UCS), and contaminant immobilization (TCLP method) of geopolymers. Geochemical modeling via the PHREEQC program was employed to interpret the results. The findings indicated that the UCS value decreased in zeolite-based geopolymers as the content of FT increased due to the inertness of the tailings and the low calcium content in the system (Ca/Si ≤ 0.3), in contrast to the FA-based geopolymer. The highest UCS of 44.3 MPa was recorded in an FA-based geopolymer containing 50% flotation tailings, with optimal molar ratios of 0.4 for Ca/Si, 3.0 for Si/Al, and 1.1 for Na/Al. In conclusion, the geopolymerization process has been determined to be a suitable technological approach for the sustainable treatment and reuse of FTs. Full article
(This article belongs to the Section Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p>X-ray patterns of the flotation tailings (FT, NZ, and FA) and geopolymers FT50NZ50, FT50NaZ50, and FT50FA50.</p>
Full article ">Figure 2
<p>SEM-EDS analysis of the raw samples: (<b>a</b>) FT, (<b>b</b>) NZ, and (<b>c</b>) FA.</p>
Full article ">Figure 3
<p>SEM-EDS analysis of the geopolymers: (<b>a</b>) FT50NZ50, (<b>b</b>) FT50NaZ50, and (<b>c</b>) FT50FA50.</p>
Full article ">Figure 4
<p>TG and DTG curvesof samples flotation tailings (FTs) and geopolymer products FT50NaZ50 and FT50FA50.</p>
Full article ">Figure 5
<p>DSC curves for samples of flotation tailings (FTs) and geopolymer products FT50NaZ50 and FT50FA50.</p>
Full article ">Figure 6
<p>Unconfined compressive strength of geopolymer based on (<b>a</b>) NZ, (<b>b</b>) NaZ, and (<b>c</b>) FA with FT content 20 wt.%, 35wt.%, and 50 wt.%, aged 7, 14, and 28 days.</p>
Full article ">Figure 7
<p>SI values of hydrated and mineral phases potentially formed in the geopolymer systems.</p>
Full article ">Figure 8
<p>Concentrations of heavy metals in leachate after TCLP test on (<b>a</b>) natural zeolite, (<b>b</b>) Na-modified zeolite, and (<b>c</b>) fly-ash-based geopolymer.</p>
Full article ">
16 pages, 14767 KiB  
Article
Molecular Design and Mechanism Study of Non-Activated Collectors for Sphalerite (ZnS) Based on Coordination Chemistry Theory and Quantum Chemical Simulation
by Xiaoqin Tang, Yilang Pan, Jianhua Chen and Ye Chen
Molecules 2024, 29(24), 5882; https://doi.org/10.3390/molecules29245882 - 13 Dec 2024
Viewed by 386
Abstract
Sphalerite flotation is generally achieved by copper activation followed by xanthate collection. This study aims to propose a design idea to find novel collectors from the perspective of molecular design and prove the theoretical feasibility that the collector can effectively recover sphalerite without [...] Read more.
Sphalerite flotation is generally achieved by copper activation followed by xanthate collection. This study aims to propose a design idea to find novel collectors from the perspective of molecular design and prove the theoretical feasibility that the collector can effectively recover sphalerite without copper activation. To address this, 30 compounds containing different structures of sulfur atoms and different neighboring atoms were designed based on coordination chemistry. Twelve potential collectors were screened, and their properties and interactions with a hydrated sphalerite (110) surface were evaluated. Compound 27 (C2H4S22−) showed the greatest reactivity, suggesting that the double-coordination structure of two sulfhydryl groups is an effective molecular structure for direct sphalerite flotation. The DFTB+ and MD results demonstrate that 1,2-butanedithiol (C4H10S2), having a similar coordination structure to compound 27, has the potential to replace the traditional reagent scheme of sphalerite flotation. The strong reagent–surface interaction is attributed to the overlap of Zn 3d with S 3p orbitals, the most negative electrostatic potential, the relatively high EHOMO and low average local ionization energy, and the eliminated steric hindrance effect. It is expected that this study can provide a design idea for the targeted design and development of novel reagents for complex sulfide ore flotation. Full article
Show Figures

Figure 1

Figure 1
<p>Optimized geometry configurations of 12 collectors adsorbed on the sphalerite (110) surface in the presence of 3 layers of water molecules (unit: Å): (<b>a</b>) 1-BT, (<b>b</b>) DS, (<b>c</b>) TP, (<b>d</b>) DTP, (<b>e</b>) DDTC, (<b>f</b>) DET, (<b>g</b>) ETC, (<b>h</b>) DPT, (<b>i</b>) BX, (<b>j</b>) EDX, (<b>k</b>) 1,2-BDT and (<b>l</b>) DTPP.</p>
Full article ">Figure 2
<p>DOSs of the surface Zn atoms and (<b>a</b>) water O (<b>b</b>) collector S atoms and their interactions on the sphalerite (110) surface.</p>
Full article ">Figure 3
<p>RDFs of Zn atoms on the sphalerite (110) surface with respect to O atoms in water molecules in the presence of 12 collectors: RDFs of (<b>a</b>) 1-BT, (<b>b</b>) DS, (<b>c</b>) TP, (<b>d</b>) DTP, (<b>e</b>) DDTC, (<b>f</b>) DET, (<b>g</b>) ETC, (<b>h</b>) DPT, (<b>i</b>) BX, (<b>j</b>) EDX, (<b>k</b>) 1,2-BDT and (<b>l</b>) DTPP.</p>
Full article ">Figure 4
<p>Flotation recoveries of natural sphalerite (pH = 7) and sphalerite collected by BX (pH = 7; [BX] = 1 × 10<sup>−4</sup> mol/L), BX + CuSO<sub>4</sub> (pH = 7; [BX] = 1 × 10<sup>−4</sup> mol/L and [CuSO<sub>4</sub>] = 1 × 10<sup>−4</sup> mol/L), and 1,2-BDT (pH = 7; [1,2-BDT] = 1 × 10<sup>−4</sup> mol/L).</p>
Full article ">
18 pages, 8535 KiB  
Article
New Insight into the Effect of Particle Surface Roughness on Flotation Efficiency: An Experimental and Theoretical Analysis
by Rui Zhang, Shiqi Dai, Jincheng Liu, Zhili Yang, Baisheng Nie, Yaowen Xing and Xiahui Gui
Minerals 2024, 14(12), 1267; https://doi.org/10.3390/min14121267 - 13 Dec 2024
Viewed by 477
Abstract
The significance of the roughness of particles in flotation has been recognized for several decades. In this study, to investigate the effect of particle roughness, a series of rod milling with different grinding durations is performed on the quartz particles. Subsequently, the particle [...] Read more.
The significance of the roughness of particles in flotation has been recognized for several decades. In this study, to investigate the effect of particle roughness, a series of rod milling with different grinding durations is performed on the quartz particles. Subsequently, the particle roundness, wetting rate, and floatability are measured. Finally, the flotation kinetic experiments are conducted based on a fixed particle size, and the interaction energy between rough particles and bubbles is calculated based on extended DLVO (E-DLVO) theory. Results show that by increasing the grinding time, the hydrophobic particle surface roughness increases, and the improved flotation rate constant increases. While the results of the E-DLVO calculation show that the interaction energy barrier between particles and bubbles decreases significantly as the roughness increases, the theoretical calculation and the experimental results confirm that an increase in the particle surface roughness decreases the interaction energy barrier between particle bubbles and increases the flotation rate. Full article
(This article belongs to the Special Issue Interfacial Chemistry of Critical Mineral Flotation)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of rough particles in water and their interaction with a bubble.</p>
Full article ">Figure 2
<p>Size distribution of the different samples after sieving.</p>
Full article ">Figure 3
<p>Image processing steps in Image J software 6.0: (<b>a</b>) Original; (<b>b</b>) threshold; (<b>c</b>) analyzed; (<b>d</b>) results.</p>
Full article ">Figure 4
<p>Grayscale surface of SEM image of particle subjected to different samples: (<b>a</b>) sample 1; (<b>b</b>) sample 2; (<b>c</b>) sample 3; (<b>d</b>) sample 4.</p>
Full article ">Figure 5
<p>Fractal dimension fitting results of particle surface for different samples: (<b>a</b>) sample 1; (<b>b</b>) sample 2; (<b>c</b>) sample 3; (<b>d</b>) sample 4.</p>
Full article ">Figure 6
<p>(<b>a</b>) First-order classical kinetic model fitted to quartz bubble flotation test data; (<b>b</b>) improved flotation rate constants of quartz particles with grinding time.</p>
Full article ">Figure 7
<p>(<b>a</b>) Wetting curve of quartz particles on deionized water after subjection to different grinding durations: (<b>b</b>) Wetting rate fitting of quartz particle at wetting stag.</p>
Full article ">Figure 8
<p>Quartz particle—bubble attachment angle of different samples varying with time.</p>
Full article ">Figure 9
<p>Potential energy curves between a bubble and particles with different asperity radii: (<b>a</b>) 4 nm; (<b>b</b>) 6 nm; (<b>c</b>) 8 nm; (<b>d</b>) 10 nm.</p>
Full article ">Figure 10
<p>Potential energy between rough particles and bubbles in water: variations in (<b>a</b>) van der Waals potential energy, (<b>b</b>) electrical double-layer interaction energy, and (<b>c</b>) hydrophobic interaction potential energy. (<b>d</b>) Variation law of potential energy barrier with asperity radius.</p>
Full article ">Figure 11
<p>Summary of findings.</p>
Full article ">
Back to TopTop