Next Article in Journal
Sediment-Hosted Rare-Earth Elements Mineralization from the Dian-Qian District, Southwest China: Mineralogy and Mode of Occurrence
Previous Article in Journal
A New Mineral Calcioveatchite, SrCaB11O16(OH)5·H2O, and the Veatchite–Calcioveatchite Isomorphous Series
Previous Article in Special Issue
Bubble Size Characterization in the HydroFloat® Fluidized-Bed Flotation Cell Using Tap Water and Seawater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Open-Circuit Technology of Zinc Oxide Ore Flotation with Ternary Collector and Its Adsorption Characteristics on Smithsonite Surface

1
State Key Laboratory of Complex Nonferrous Metal Resources Clean Utilization, Faculty of Land Resource Engineering, Kunming University of Science and Technology, Kunming 650093, China
2
Yunnan Key Laboratory of Green Separation and Enrichment of Strategic Mineral Resources, Kunming University of Science and Technology, Kunming 650093, China
*
Authors to whom correspondence should be addressed.
Minerals 2024, 14(9), 902; https://doi.org/10.3390/min14090902
Submission received: 2 August 2024 / Revised: 29 August 2024 / Accepted: 30 August 2024 / Published: 2 September 2024
(This article belongs to the Special Issue Advances in Flotation of Copper, Lead and Zinc Minerals)

Abstract

:
The sulfidization-amine flotation method is commonly used for the beneficiation of zinc oxide ores. Lanping zinc oxide ores contains 8.40% zinc, with the main mineral being smithsonite; additionally, they have a high mud content. Conventional sulfidization–ammonium flotation presents a low flotation index and unsatisfactory flotation froth. A new open-circuit technology is employed to treat Lanping zinc oxide ores, where Na2S, KG-248, and dodecyl amine + sodium isoamyl xanthate + ammonium dibutyl dithiophosphate are used as the regulator, depressant, and ternary collector, respectively. Consequently, the flotation indices for the zinc grade and recovery are 28.71% and 86.24%, respectively, and the flotation froth becomes more stable. Subsequently, the flotation behavior and adsorption mechanism of smithsonite with a ternary collector are investigated. The flotation recovery of smithsonite increases to 94.40% after treatment with the ternary collector. Surface-analysis results indicate that the ternary collector can synergistically adsorb onto the sulfidized smithsonite surface to enhance its hydrophobicity, thus increasing the floatability of smithsonite. Meanwhile, the total consumption of the collector in the ternary-collector system is lower than that in the binary- or unitary-collector system.

1. Introduction

Zinc is a crucial component in the field of non-ferrous metals and is widely used in various industries, such as the electrical, mechanical, military, metallurgical, chemical, optical, and pharmaceutical industries [1,2]. The primary source for zinc extraction is zinc sulfide ore; however, the continuous increase in global demand for zinc resources has resulted in the gradual depletion of zinc sulfide ore resources. Therefore, the possibility of using zinc oxide ores must be investigated to satisfy the increasing demand [3,4,5].
Flotation is the most commonly used method for enriching zinc oxide ores and includes direct flotation (fatty acid, flocculation, and chelation) and sulfidization flotation (sulfidization amine and sulfidation xanthate) [6,7,8,9,10,11,12,13,14,15,16]. Among them, sulfidization-amine flotation treated with zinc oxide ore has been widely employed and investigated owing to its high flotation recovery [17,18]. Feng et al. [19] discovered that the quality of the sulfidation effect determines whether zinc oxide can yield excellent results and that sulfidation can yield zinc monosulfide and zinc polysulfide. Additionally, adding an appropriate amount of the sulfiding agent can yield more polysulfides, thus exerting activation and promotion effects on the flotation of zinc oxide ores. However, excessive sulfidation causes the generation of excessive sulfide ions in the pulp, thereby inhibiting the flotation effect. Mehdilo et al. [20] discovered that the dosage of sulfidizing agents and the pulp pH significantly affected the effectiveness of sulfidization-amine flotation methods. In the flotation process of smithsonite, when the optimal dosage of sodium sulfide was 500 g/t and Armac C and Armac T were used as collectors, 82.60% and 83.70% recoveries were achieved, respectively. However, a continued increase in the sulfide concentration decreased the smithsonite recovery. Additionally, sulfidization-amine methods pose several severe issues, such as the generation of significant amounts of viscous froth when amine collectors are added. These viscous froths are not easily defoamed and are sensitive to the slurry, thus resulting in a significant decline in the flotation index and the emergence of empty and large bubbles [21,22,23,24].
To address these challenges, researchers have investigated combinations of anionic and cationic collectors for the treatment of zinc oxide ores [25,26,27,28]. Anionic or cationic collectors with opposite charges promote adsorption through electrostatic interactions. Hosseini et al. [29] discovered that the recovery of smithsonite was low regardless of the collector concentration when using a single collector, whereas the recovery of smithsonite increased significantly by employing potassium isoamyl xanthate (KAX) and dodecyl amine (DDA) combined collectors. This is attributable to the use of anion and cation collectors, which decreased the electrostatic head–head repulsion between the surface and ammonium ions and increased the lateral tail–tail hydrophobic bonds. Majid et al. [30] employed Armac C and KAX as anionic and cationic collectors, respectively, to obtain zinc oxide minerals synergistically. Sulfidized regions on the zinc oxide surface with a negative charge were termed “strong spots”. Positively charged Armac C ions were firmly adsorbed onto these “strong spots”, whereas negatively charged KAX ions were easily adsorbed with Armac C, thereby increasing the surface coverage of the collectors on the mineral surface and enhancing the floatability. Wang et al. [31] mixed sodium oleate and butyl hydroxy anisole in a molar ratio of 2:1 to prepare a new combined collector named “BHOA”. It can adsorb onto the surface of smithsonite, thus reducing the surface potential to below −50 mV. Consequently, the sulfidization of smithsonite is improved, thus enhancing its flotation performance. Furthermore, Wang et al. [32] discovered that by combining KAX and DDA for smithsonite flotation, DDA ions occupied the space between KAX anions, thereby preventing adjacent KAX from experiencing electrostatic repulsion. This increased the attraction between the tail groups, thus allowing the further adsorption of DDA cations on the surface of smithsonite; consequently, a remarkable concentrate recovery of 96.80% was achieved. Combining anionic and cationic collectors to treat zinc oxide minerals enhances the hydrophobicity of the mineral surface, owing to the synergistic effects between the agents. In the alkaline pH range, where the mineral surface is negatively charged, cations are preferentially adsorbed onto the mineral surface, which are subsequently co-adsorbed through electrostatic neutralization, thus resulting in complexes that enhance the floatability of the mineral [33,34]. However, in industrial production for low-grade zinc oxide ores characterized by a fine particle size, a high mud content, and abundant gangue minerals, the floatability difference between gangue and purpose minerals is minimal. Furthermore, gangue reacts easily with the amine collector, thus deteriorating the concentrate grade and recovery [35,36,37,38]. Therefore, the appropriate combination of anionic and cationic collectors that exhibits synergistic effects must be identified.
In this work, owing to its low grade, fine particle size, and more kinds of gangue minerals, the Lanping zinc oxide ore renders sulfidization-amine flotation ineffective. After optimizing the reagent system and flotation process flow, a ternary-collector system (DDA + NaIX + ADD) was used to treat zinc oxide ore. Subsequently, a new open-circuit process and depressant (KG-248) were introduced to minimize the effects of gangue, slimes, and residual froth in the pulp circulation on the flotation concentrate. Additionally, microflotation, zeta-potential analysis, adsorption tests, X-ray photoelectron spectroscopy (XPS), and contact-angle measurements were performed to elucidate the surface mechanisms of the cooperative adsorption of the ternary collector on the smithsonite surface. This study offers theoretical support for the development and utilization of zinc oxide ores with a high mud content that cannot be beneficiated easily.

2. Experiment

2.1. Materials and Reagents

Zinc oxide ore was obtained from Jinding Zinc Industry Co., Ltd., Lanping County, Yunnan Province, China. Samples smaller than 2 mm were obtained by crushing and sieving. The samples were ground to −0.074 µm and then prepared for analysis after being dried. Results of chemical multi-element analysis revealed that the zinc and lead grades in the ore were 8.40% and 1.26%, respectively. The contents of calcite and quartz in the raw ore are 35.33% and 26.26%, respectively, which are the main components of gangue minerals. Meanwhile, results of phase analysis indicated that the zinc appeared primarily in the form of smithsonite, which constituted 77.99% of the total element. The occurrence form and distribution rate of the remaining zinc are, respectively, 11.53% for sphalerite, 7.74% for hemimorphite, 2.63% for limonite, 0.10% for siderite, and 0.01% for rhodochrosite. According to the distribution rate of zinc in the ore, smithsonite should be considered in the recovery.
Smithsonite samples were obtained from Lanping County, Yunnan Province, China. A high-purity smithsonite sample was obtained via meticulous selection, after which it was cleaned ultrasonically and dried naturally. Subsequently, it was ground using an agate grinder, following which smithsonite samples of different sizes were obtained via screening using a Taylor standard sieve. Finally, microflotation experiments and analytical tests were conducted. The X-ray diffraction (XRD) results of the smithsonite sample, as shown in Figure 1, indicated no clear impurity peaks. Meanwhile, the elemental-analysis results of the smithsonite sample indicated that the zinc content was 50.14%. These results confirmed the high purity of the smithsonite sample, which satisfies the requirements of a single-mineral flotation experiment.
Commercially pure Na2CO3, Na2S, KG-248, NaIX, ADD, and 2# oil and analytically pure DDA were used in the flotation tests of the zinc oxide ores. Na2CO3 was used to regulate the lead-flotation process. Na2S was used to sulfidize lead and zinc oxide, whereas KG-248 was used as a depressant in the zinc flotation process. NaIX, DDA, and ADD were used as collectors, and the combination of NaIX and ADD collectors was named GS-2. For the flotation of smithsonite, Na2S·9H2O (analytically pure), ADD (analytically pure), NaIX (commercially pure), ADD (commercially pure) were used as the sulfidization reagent (source of S2− ions), collector, and frother, respectively. Additionally, NaOH (analytically pure) and HCl (analytically pure) were utilized as pH regulators, and NaCl (analytically pure) was utilized as the electrolyte for zeta-potential measurements. Tap water and deionized water were used in the zinc oxide ore and smithsonite tests, respectively. Notably, KG-248 was a macromolecular organic agent with a molecular weight of about 20,000, which was a self-made agent with good water solubility, no highly toxic and harmful components, and had a good effect on inhibiting carbonate and silicate gangue minerals.

2.2. Flotation Tests

Flotation was conducted on the zinc oxide ores using an XFD flotation machine. Specifically, 400 g of zinc oxide ore with a particle size of less than 2 mm and 400 mL of tap water were added to a rod mill to obtain ground products with the desired fineness. The grinding fineness was obtained by collecting the screen undersize using the Taylor standard screen. Subsequently, the ground products were introduced into the cell of the 1.5 L XFD flotation machine, followed by the sequential addition of Na2CO3, Na2S, NaIX, and 2# oil, which were stirred for 3, 5, 3, and 2 min, respectively. The lead rough concentrate was obtained after lead flotation. Subsequently, Na2S, KG248, DDA (dodecyl amine), NaIX (sodium isoamyl xanthate), and ADD (ammonium dibutyl dithiophosphate) were added to the lead tailing pulp and reacted for 5, 3, 3, 3, and 3 min, respectively, and the zinc rough concentrate was obtained after zinc flotation. The obtained products were dried, weighed, and Atomic Absorption Spectroscopy was used for analysis; subsequently, the recovery was calculated. After determining the reagent dosage, open- and closed-circuit tests were performed based on a specified flow chart.
In the smithsonite microflotation test, 2 g of smithsonite with a particle size of −75 to +38 µm and deionized water (35 mL) were introduced into a 40 mL flotation cell. The pulp pH was adjusted to 11 using NaOH. Subsequently, Na2S, DDA, NaIX, and ADD solutions were added and stirred for 5, 3, 3, and 3 min, respectively. Subsequently, the froth (concentrate) and cell products (tailing) were obtained, and the flotation recovery of smithsonite was calculated after it was dried and weighed.

2.3. Zeta Potential

In this study, zeta potentials were measured using a Zetasizer-3000HS instrument (Malvern, Worcestershire, UK). The prepared smithsonite sample (20 mg, −5 µm) was placed inside a beaker containing NaCl solution (1 mM), the pulp pH was adjusted, and reagents were added and stirred for a certain duration depending on the flotation-test requirements. After the reaction was completed, the mixture was allowed to settle naturally for 10 min, and the supernatant was extracted using a disposable plastic straw to test the zeta potential. Each sample was measured three times, and the average value was recorded as the final value.

2.4. Adsorption Tests

Ultraviolet–visible spectroscopy was performed to evaluate the adsorption capacity of the collector on the smithsonite surface. Initially, 2 g of smithsonite (−38 µm) was introduced into a beaker containing 35 mL of deionized water. The necessary reagents were added and stirred for a certain duration using a magnetic stirrer, depending on the flotation-experiment requirements. After the reaction was completed, the mixture was allowed to precipitate. Subsequently, the supernatant was extracted using a disposable plastic straw and then transferred to a centrifuge tube. After centrifugation for 15 min, the supernatant was transferred meticulously to a quartz cuvette and analyzed using a UV-2700 spectrophotometer (Shimadzu, Kyoto, Japan). The collector concentration remaining in the solution was obtained, and the amount of collector adsorbed onto the smithsonite surface was calculated. For NaIX and ADD, UV spectroscopy was used to directly measure the residual concentration in the supernatant. Due to the lack of characteristic absorption peaks in the ultraviolet spectrum of DDA, eosin-Y was used as an indicator for the determination of DDA concentration, and the specific steps were consistent with those in the previous literature [39].

2.5. XPS Analysis

XPS analysis was performed to qualitatively and quantitatively assess the elemental composition and chemical state of the smithsonite sample. Specifically, 2 g of smithsonite (−38 µm) was mixed with 35 mL deionized water in a beaker, followed by the addition of the required reagents after the pulp pH was adjusted. The smithsonite and reagent reacted sufficiently after stirring was performed for a certain duration using a magnetic stirrer. Subsequently, the mixture product was filtrated. The filtered residue was rinsed with deionized water and then allowed to dry naturally. Subsequently, it was subjected to XPS analysis using a pHI5000 spectrometer (ULVAC-PHI Inc., Kanagawa, Japan).

2.6. Contact-Angle Test

Contact-angle measurements were performed using a JY-82 contact-angle analyzer. The rectangular smithsonite sample was polished using 400-grit and 5000-grit aluminum-oxide sandpaper and cleaned using ethanol. Subsequently, it was placed in a beaker containing 35 mL of deionized water, the pulp pH was adjusted to 11, and the required reagents were added in the specified sequence. The sample was removed after the reagents reacted with the mineral surface. Next, it was rapidly dried and transferred to a platform. Subsequently, its contact angle was measured using the liquid-droplet method.

3. Results and Discussion

3.1. Flotation of Zinc Oxide Ores

To weaken the effect of lead on the flotation recovery of zinc, we prioritized the recovery of lead from zinc oxide ores. The optimal reagent conditions were 1000 g/t Na2CO3, 2000 g/t Na2S, 250 g/t NaIX, and 40 g/t 2# oil, which resulted in 7.16% and 80.40% for the lead concentrate grade and recovery, respectively. Subsequently, the flotation of zinc, including the conditions and open-circuit tests, was conducted to increase the flotation index of the zinc concentrate.

3.1.1. Flotation Test of Zinc

Grinding fineness as well as reagent type and dosage are important parameters affecting the mineral flotation index. Based on the properties of zinc oxide ores, many exploratory experiments have been conducted to determine the reagent types for conducting zinc flotation. Ultimately, Na2S, KG-248, and the combination of DDA + NaIX + ADD were identified as the sulfidization reagent, depressant, and ternary collector, respectively, to recover zinc. Subsequently, the effects of the grinding fineness and reagent dosage on zinc flotation were investigated, and the results are presented in Figure 2 and Figure 3.
As shown in Figure 2, the grinding fineness of the ores affected the grade and recovery of the zinc rough concentrate. The grade and recovery of zinc first increased and then decreased. The best indices were achieved when 85% of the particles exhibited a grinding fineness of −0.074 mm. Meanwhile, the zinc grade and recovery were 23.74%, and 79.22%, respectively, when less than 85% of the particles measured −0.074 mm, which is attributable to the insufficient dissociation of the purpose mineral. By contrast, the ore was overground when the proportion exceeded 85%, which exacerbated the mud problem and thus adversely affected the grade and recovery index.
Na2S would produce HS and S2− ions, which could be firmly adsorbed on the mineral surface and form a film with ZnS as the main component [19]. This film made the mineral easier to combine with the flotation agent, which not only enhanced the surface hydrophobicity of the mineral but also facilitated the smooth progress of the subsequent flotation process. Because the C=S bond on the functional group of the thiol collector was very close to the polarity of the metal sulfide. This similarity allowed them to form a tight combination to form a MeS (Me: metal element) cell [14]. These crystal cells were like a bridge connecting the collector and the surface of the oxidized mineral, ensuring that the adsorption of the collector on the mineral surface could meet the needs of froth mineralization. As shown in Figure 3a, the grade and recovery of zinc initially increased and then decreased as the Na2S dosage increased from 4000 to 12,000 g/t. The optimal flotation index was obtained when the Na2S dosage was 10,000 g/t, thus indicating that a minute amount of Na2S was insufficient to completely sulfidize the zinc oxide surface, thus resulting in an undesirable flotation index. However, excessive Na2S suppressed the flotation of zinc oxide. This is attributed to the residual Na2S in the pulp solution, which resulted in competitive adsorption with the collector. Compared with the case without KG-248, the grade and recovery of zinc increased significantly when KG-248 was added (Figure 3b), thus indicating that KG-248 significantly affects dispersed mud and gangue minerals. However, the method for improving the zinc flotation index when the dosage of KG-248 exceeds 400 g/t remains ambiguous; therefore, the dosage of KG-248 was set as 400 g/t.
Figure 3c shows that although the zinc recovery increased gradually, the grade decreased as the DDA dosage increased from 50 to 250 g/t. This is because as the DDA dosage increased, DDA was adsorbed onto the surface of mud and gangue minerals, which resulted in the formation of many froth layers. Consequently, the zinc grade decreased and the flotation environment deteriorated. Some studies demonstrated that combining anions and cations as collectors effectively enhanced the stability of froth [40]. Therefore, an anion collector (GS-2) was added to effectively obtain zinc in the cation (DDA) system, and the results are shown in Figure 3d. When the DDA dosage was 200 g/t, the grade and recovery of zinc increased with the addition of 150 g/t GS-2, the flotation environment improved, and the flotation froth layer became stable; however, when the dosage of GS-2 was increased to 300 g/t, zinc recovery increased significantly but the grade deteriorated slightly. Thus, the addition of 300 g/t anionic collector to the DDA system enhanced the flotation index and improved flotation, thus facilitating the beneficiation of zinc.

3.1.2. Flotation Open-Circuit Test

After determining the optimal reagent and dosage via condition tests, the process optimization of the zinc oxide ore was carried out in order to maximize the recovery of the purpose mineral. Ultimately, a new open-circuit flotation process was performed to treat the zinc oxide ores, and verification tests were conducted based on the flow chart presented in Figure 4. Adopted the flotation process of the first lead beneficiation, followed by that of zinc. The zinc concentration was measured after one rougher step, one scavenger step, and two cleaner steps. To avoid the deterioration of the flotation index and the phenomenon caused by the sequential return of middling ore as well as by the middling ore arising from lead cleaner and zinc cleaner I, merged and treated the concentrates obtained from zinc scavenger in zinc cleaner II. The feasibility and stability of the open-circuit process were verified five times, and a favorable and stable flotation index was obtained, as listed in Table 1.
The data in Table 1 show that qualified zinc concentrate products were obtained via the new open-circuit test. The zinc concentrate grade and recovery were 28.71% and 86.24%, respectively. Furthermore, in the absence of closed-circuit processes, using the new open-circuit process obtained more advantages. The combination of anionic and cationic collectors could enhance the collection effect so that fewer collectors could be used to obtain a better flotation effect. A reasonable dosage of DDA and an open-circuit process that could discharge ore in time would effectively solve the problem of froth accumulation in flotation cells.
The cost of the reagents used in the process in Figure 4 was estimated to ensure its economic viability. The data of each reagent are taken from China, as listed in Table 2.
As can be seen from Table 2, the cost per tonne of ore treated with the reagents regime is approximately 54.51¥, and the zinc concentrate with a grade of more than 28.00% and a recovery of more than 85% can be obtained after treatment by this flotation process. Notably, there was no return stage in the open-circuit process, which meant that less equipment and energy consumption would be used, which contributed to energy saving and environmental protection. Thus, this approach is promising when used in the industrial application of zinc oxide ores.

3.2. Flotation Behavior and Surface-Adsorption Analysis of Smithsonite

The flotation results of the zinc oxide ore indicated that the open-circuit process with ternary-collector treatment (DDA + NaIX + ADD) yielded a stable flotation froth and an exceptional flotation index. The phase-analysis result of the raw ore shows that the zinc oxide ore appeared primarily in the form of smithsonite. Therefore, we investigated the effect of the ternary collector (DDA + NaIX + ADD) on the smithsonite flotation behavior as well as the adsorption mechanism on the smithsonite surface.

3.2.1. Effect of Ternary Collector on Smithsonite Flotation Behavior

The effects of pH and Na2S concentration on the floatability of smithsonite are shown in Figure 5. Based on Figure 5a, as the pH increased from 5 to 11, the floatability of smithsonite increased, and its recovery increased from 23.20% to 64.90%. However, its flotation recovery decreased slightly as the pH increased to 12, thus indicating that the optimum flotation pH for smithsonite was 11. The effect of Na2S concentration on the smithsonite floatability in different collector systems at pH 11 is depicted in Figure 5b; the recovery of smithsonite improved as the Na2S concentration increased from 1 × 10−4 to 8 × 10−4 mol/L. The optimum flotation effectiveness for smithsonite was reached at the dosage of Na2S of 8 × 10−4 mol/L. Beyond the concentration of 8 × 10−4 mol/L, the amount of Na2S continued to increase, but the recovery did not improve significantly; there was little benefit under all three systems. Notably, the excess Na2S did not significantly inhibit the co-adsorption of the collectors (DDA, NaIX, and ADD) on the surface of smithsonite under the ternary-collector system, which signifies a favorable synergistic effect.
Figure 6 shows the effect of the collector concentration on the smithsonite floatability in different systems. The floatability of the sulfidized smithsonite after the ternary-collector treatment was superior to that after the binary- and unitary-collector treatments. At low concentrations, DDA (<7 × 10−5 mol/L) did not significantly improve the recovery. However, at a DDA concentration of 7 × 10−5 mol/L, the recovery of smithsonite increased significantly. A further increase in DDA concentration did not significantly enhance the flotation effect (Figure 6a). In the binary-collector system, the addition of NaIX significantly improved the recovery of smithsonite, with better performances demonstrated at lower dosages, and the recovery increased to 74.80% (Figure 6b). In the DDA + ADD collector system (Figure 6c), ADD exhibited excellent acquisition ability, thus increasing the smithsonite recovery to 76.20% at a concentration of 7 × 10−5 mol/L. In the ternary-collector system, the maximum smithsonite recovery achieved was 94.40% (Figure 6c), which was much higher than those achieved by the binary and single collectors. Notably, the co-adsorption of the ternary collector on the smithsonite surface did not result in mutual inhibition. By contrast, collectors with opposite charges enhanced the co-adsorption effect, allowing more collectors to be adsorbed onto the smithsonite surface, thus improving the surface hydrophobicity of smithsonite. Additionally, in the presence of both anionic collectors, the half-life of the froth generated by DDA reduced significantly.

3.2.2. Zeta-Potential Measurements

The zeta potential is an important parameter for assessing the electrical properties of mineral particles at the solid–liquid interface [41,42]. The zeta potential of the sulfidized smithsonite surface under various collector systems was measured, and the results are shown in Figure 7. As shown, the zeta potential shifted in the negative direction as the pH increased from 5 to 12, which is consistent with the results reported in the literature [43]. Compared with the case of untreated sulfidized smithsonite, the zeta potential shifted in the positive direction after DDA, DDA + NaIX, and DDA + NaIX + ADD treatments, thus indicating the adsorption of positively charged ammonia species on the sulfidized smithsonite surface. However, the zeta potential shifted in the negative direction in the DDA + NaIX system as compared with that in the DDA system, which is attributable to the adsorption of negatively charged species on the sulfidized smithsonite surface. Furthermore, the addition of ADD to the pulp solution in the DDA + NaIX system resulted in a more negative zeta potential, which is attributable to the adsorption of dibutyl dithiophosphate species on the sulfidized smithsonite surface. These phenomena indicate that the ternary collector was adsorbed onto the sulfidized smithsonite surface, which enhanced the hydrophobicity of the mineral surface and benefited the flotation of smithsonite.

3.2.3. Adsorption Experiments

The adsorption behavior of the ternary-collector system on the sulfidized smithsonite surface was investigated via adsorption measurements. The initial collector concentration was denoted as C0, the residual amount of the collector in the pulp C1 was determined, and the adsorption capacity of the collector was calculated using the formula τ = v(C0 − C1)/m (Figure 8). Figure 8a illustrates the changes in the adsorption capacity of DDA in the different systems. Within the DDA concentration range of 1 × 10−5 to 5 × 10−5 mol/L, the adsorption capacities of all three systems varied only slightly, thus indicating low adsorption efficiency at low DDA concentrations. However, the adsorption of DDA increased significantly as the DDA concentration increased to 7 × 10−5 mol/L, particularly under the ternary-collector system. Meanwhile, the adsorption capacity increased from 5.31 × 10−6 to 9 × 10−7 mol/g, thus suggesting that the adsorption of NaIX and ADD enhanced the adsorption of DDA on the sulfidized smithsonite surface. Similarly, as shown in Figure 8b,c, the adsorption capacity of NaIX and ADD increased with the amounts of NaIX and ADD. In particular, the ternary system exhibited a greater increase than the binary and unitary systems, thus indicating that the ternary collector not only promoted adsorption but also synergistic adsorption, thus resulting in the absorption of more collectors onto the sulfidized smithsonite surface, which consequently enhanced its floatability.

3.2.4. XPS Analysis

To further examine the synergistic adsorption mechanism of the ternary collector on the sulfidized smithsonite surface, the elemental composition and chemical state of the mineral surface were analyzed using XPS [44,45]. Figure 9 and Figure 10 illustrate the elemental composition and high-resolution N 1s XPS spectra of the smithsonite surface after treatment with different reagents.
Only C, O, S, and Zn were detected on the sulfidized smithsonite surface (Figure 9a) at proportions of 35.99%, 43.36%, 0.65%, and 20.00%, respectively. However, with the addition of DDA (Figure 9b), a new N element was discovered on the sulfidized smithsonite surface at a proportion of 0.75%. Identically, the data in Figure 10a show that no N 1s peak was detected on the sulfidized smithsonite surface. Additionally, with the addition of DDA (Figure 10b), a new N 1s XPS spectrum comprising two split peaks with binding energies of 399.94 and 405.39 eV were detected on the mineral surface, which corresponded to -NH2 and R-NH3+ groups in the DDA species, respectively [46,47]. Meanwhile, the distributions of -NH2 and R-NH3+ were 67.43% and 32.57%, respectively, thus indicating that DDA was adsorbed onto the mineral surface. After DDA + NaIX treatment, the binding energies of -NH2 and R-NH3+ groups were 399.90 and 405.40 eV, respectively; the distribution of -NH2 increased from 67.43% to 73.68%, the proportion of R-NH3+ decreased by 6.25% (Figure 10c), and the proportion of N increased to 1.05% (Figure 9c). This shows that a higher amount of DDA was transferred to the mineral surface from the pulp solution and that the presence of NaIX enhanced the adsorption of DDA on the sulfidized smithsonite surface. In the DDA + NaIX + ADD ternary-collector system (Figure 10d), a new peak appeared at a binding energy of 402.02 eV, which originated from the C-N species in the ADD molecules, thus indicating the adsorption of ADD on the sulfidized smithsonite surface. Additionally, the increase in the peak area and intensity indicates that the adsorption of DDA increased, thus resulting in an increase N to 1.65% (Figure 9d). These results indicate a more intense adsorption reaction on the sulfidized smithsonite surface under the synergistic action of the DDA + NaIX + ADD ternary-collector system. The XPS analysis results are consistent with the test results of the zeta potential and adsorption capacity.

3.3. Contact-Angle Test

The contact angles of the smithsonite surface under various conditions are shown in Figure 11. The contact angle of the untreated smithsonite surface was 33.12° (Figure 11a), which is consistent with the result in the literature [48], thus indicating that the smithsonite surface is highly hydrophilic. The contact angle of the sulfidized smithsonite surface treated with DDA increased to 60.17° (Figure 11b), thus indicating that DDA adsorbed onto the sulfidized smithsonite surface and increased the hydrophobicity of the mineral surface. In the binary-collector system (DDA + NaIX), the contact angle further increased to 62.92° (Figure 11c). Based on the result above as well as the adsorption capacity and XPS analysis, the addition of NaIX improved the adsorption of DDA on the mineral surface. After treatment with the ternary-collector system (DDA + NaIX + ADD), the contact angle of the sulfidized smithsonite surface increased significantly to 79.05° (Figure 11d), which may have promoted the synergistic adsorption of the ternary collector, thus enhancing the hydrophobicity of the mineral surface and improving the floatability of the sulfidized smithsonite.

4. Conclusions

In this study, a new open-circuit process using a ternary-collector system (DDA + NaIX + ADD) was developed to treat low-grade refractory zinc oxide ore. The floatability of smithsonite and the synergistic adsorption mechanism of the ternary collector on the sulfidized smithsonite surface were investigated via flotation experiments and surface analysis. The following conclusions were obtained:
  • By adopting the new open-circuit process and using a ternary-collector system to treat low-grade zinc oxide ore, favorable flotation indices of 28.71% and 86.24% were obtained for the zinc grade and recovery, respectively. Additionally, the open-circuit process effectively solved the problem of froth buildup caused by amine collectors during flotation.
  • The results of microflotation experiments revealed that the floatability of smithsonite enhanced significantly by employing the ternary-collector treatment. Specifically, the recovery of smithsonite increased to 94.40%.
  • The zeta-potential measurements and adsorption-capacity test results indicated that the ternary collector can be co-adsorbed onto the sulfidized smithsonite surface and thus reduce the total consumption of reagents.
  • XPS results showed that the addition of NaIX promoted DDA adsorption on the mineral surface and that the adsorption of ADD increased the contact angle of the sulfidized smithsonite surface treated with the ternary collector, thus enhancing the hydrophobicity and floatability of the smithsonite.

Author Contributions

Conceptualization, S.W.; writing—original draft, Z.L.; writing—review and editing, Q.Z.; supervision, Q.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Yunnan Science and Technology Leading Talent Project grant number 202305AB350005.

Data Availability Statement

All relevant data are within the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Majid, E.; Mahdi, G.; Mehdi, I. A review of zinc oxide mineral beneficiation using flotation method. Adv. Colloid. Inter. Sci. 2014, 206, 68–78. [Google Scholar] [CrossRef]
  2. Hosking, N.C.; Ström, M.A.; Shipway, P.H.; Rudd, C.D. Corrosion resistance of zinc-magnesium coated steel. Corros. Sci. 2007, 49, 3669–3695. [Google Scholar] [CrossRef]
  3. Zhao, H.B.; Gan, X.W.; Wang, J.; Tao, L.; Qiu, W.Q.; Qiu, G.Z. Stepwise bioleaching of Cu-Zn mixed ores with comprehensive utilization of silver-bearing solid waste through a new technique process. Hydrometallurgy 2017, 171, 374–386. [Google Scholar] [CrossRef]
  4. Mudd, G.M.; Jowitt, S.M.; Werner, T.T. The world’s lead-zinc mineral resources: Scarcity, data, issues and opportunities. Ore Geol. Rev. 2017, 80, 1160–1190. [Google Scholar] [CrossRef]
  5. Irannajad, M.; Ejtemaei, M.; Gharabaghi, M. The effect of reagents on selective flotation of smithsonite–calcite–quartz. Miner. Eng. 2009, 22, 766–771. [Google Scholar] [CrossRef]
  6. Moradi, S.; Monhemius, A.J. Mixed sulphide-oxide lead and zinc ores: Problems and solutions. Miner. Eng. 2011, 24, 1062–1076. [Google Scholar] [CrossRef]
  7. Navidi Kashani, A.H.; Rashchi, F. Separation of oxidized zinc minerals from tailings: Influence of flotation reagents. Miner. Eng. 2008, 21, 967–972. [Google Scholar] [CrossRef]
  8. Li, B.B.; Zhang, G.F.; Shi, Q. Atomic insights into the interaction mechanism underpinning carbonate ion adsorption in smithsonite flotation. Process Saf. Environ. Prot. 2024, 190, 11–21. [Google Scholar] [CrossRef]
  9. Marabini, A.M.; Ciriachi, M.; Plescia, P.; Barbaro, M. Chelating reagents for flotation. Miner. Eng. 2007, 20, 1014–1025. [Google Scholar] [CrossRef]
  10. Li, G.; Zhang, Z.; Wei, Q.; Jiao, F.; Yang, C.; Li, W.; Qin, W. Study on flotation recovery of typical carbon-bearing lead-zinc sulphide ore in Guizhou with pre-decarbonization. Geochemistry 2024, 84, 126096. [Google Scholar] [CrossRef]
  11. Hosseini, S.H.; Forssberg, E. Smithsonite flotation using potassium amyl xanthate and hexylmercaptan. Trans. Inst. Min. Metall. Sect. C Miner. Process. Extr. Metall. 2006, 115, 107–112. [Google Scholar] [CrossRef]
  12. Nanda, S.; Kumar, S.; Mandre, N.R. Flotation behavior of a complex lead-zinc ore using individual collectors and its blends for lead sulfide. J. Dispers. Sci. Technol. 2023, 44, 1703–1710. [Google Scholar] [CrossRef]
  13. Wang, J.G.; Ji, Y.H.; Cheng, S.Y.; Liu, S.; Cao, J.; Chen, P. Selective flotation separation of galena from sphalerite via chelation collectors with different nitrogen functional groups. Appl. Surf. Sci. 2021, 568, 150956. [Google Scholar] [CrossRef]
  14. Shen, Z.H.; Zhang, Q.; Fang, J. Research Progress in Surface Sulfidization of Smithsonite. Nonferr. Met. Mieral Process. Sect. 2021, 01, 37–46+59. [Google Scholar] [CrossRef]
  15. Zhu, H.L.; Qin, W.Q.; Chen, C.; Chai, L.Y.; Li, L.S.; Liu, S.J.; Zhang, T. Selective flotation of smithsonite, quartz and calcite using alkyldiamineether as collector. Trans. Nonferr. Metals Soc. China 2018, 28, 163–168. [Google Scholar] [CrossRef]
  16. Luo, B.; Nie, W.L.; Dong, J.S. Effect of Lead Ions on the Sulfidization Flotation of Smithsonite Using Sodium Butyl Xanthate as a Collector. Miner. Eng. 2022, 185, 107710. [Google Scholar] [CrossRef]
  17. Ceylan, Y.; Altınışık, Y.; Cebeci, H.; Sis, L.K. A Process Mineralogy Approach to the Flotation of Complex Lead–Zinc Ores from Görgü (Malatya) Region. Min. Met. Explor. 2022, 39, 1219–1232. [Google Scholar] [CrossRef]
  18. Mehdilo, A.; Zarei, H.; Irannajad, M.; Arjmandfar, H. Flotation of zinc oxide ores by cationic and mixed collectors. Miner. Eng. 2012, 36, 331–334. [Google Scholar] [CrossRef]
  19. Feng, Q.C.; Wen, S.M. Formation of zinc sulfide specieson smithsonite surfaces and its response to flotation performance. J. Alloys Compd. 2017, 709, 602–608. [Google Scholar] [CrossRef]
  20. Mehdilo, A.; Irannajad, M.; Zarei, H. Smithsonite Flotation from Zinc Oxide Ore using Alkyl Amine Acetate Collectors. Sep. Sci. Technol. 2014, 49, 445–457. [Google Scholar] [CrossRef]
  21. Zhu, Y.H.; Huang, K.H.; Yang, S.Y.; Liang, Z.A. Studies of benzyl hydroxamic acid/calcium lignosulphonate addition order in the flotation separation of smithsonite from calcite. Int. J. Min. Sci. Technol. 2021, 31, 1153–1158. [Google Scholar] [CrossRef]
  22. Barbery, G. Complex sulphide ores: Processing options. Miner. Process. A Crossroads 1986, 117, 157–194. [Google Scholar] [CrossRef]
  23. Hosseini, S.H.; Forssberg, E. Adsorption studies of smithsonite flotation using dodecylamine and oleic acid. Min. Metall. Explor. 2006, 23, 87–96. [Google Scholar] [CrossRef]
  24. Liu, C.; Feng, Q.M.; Zhang, G.F.; Chen, W.W.; Chen, Y.F. Effect of sodium carbonate on flotation behavior of smithsonite and its mechanism in presence of sodium oleate. Chin. J. Nonferrous Met. 2017, 27, 2379–2384. [Google Scholar] [CrossRef]
  25. Ferreira, P.H.T.; Fernandes Lima, R.M. Concentration of oxidized Brazilian zinc ore by flotation: Comparative study between anionic and cationic routes. Sep. Sci. Technol. 2022, 57, 2625–2634. [Google Scholar] [CrossRef]
  26. Sun, H.F.; Niu, F.S.; Zhang, J.X. Investigation on the flotation separation of smithsonite from calcite using calcium lignosulphonate as depressant. Colloids Surf. A Physicochem. Eng. Asp. 2021, 630, 127571. [Google Scholar] [CrossRef]
  27. Shi, Q.; Feng, Q.M.; Zhang, G.F.; Deng, H. Electrokinetic properties of smithsonite and its floatability with anionic collector. Colloids Surf. A Physicochem. Eng. Asp. 2012, 410, 178–183. [Google Scholar] [CrossRef]
  28. Hosseini, S.H.; Forssberg, E. Physicochemical studies of smithsonite flotation using mixed anionic/cationic collector. Miner. Eng. 2007, 20, 624. [Google Scholar] [CrossRef]
  29. Hosseini, S.H.; Forssberg, E. Studies on selective flotation of smithsonite from silicate minerals using mercaptans and one stage desliming. Miner. Process. Extr. Metall. 2011, 120, 79–84. [Google Scholar] [CrossRef]
  30. Majid, E.; Mehdi, I.; Mahdi, G. Influence of important factors on flotation of zinc oxide mineral using cationic, anionic and mixed (cationic/anionic) collectors. Miner. Eng. 2011, 24, 1402–1408. [Google Scholar] [CrossRef]
  31. Wang, L.; Hu, G.Y.; Sun, W.; Khoso, S.A.; Liu, R.Q.; Zhang, X.F. Selective flotation of smithsonite from dolomite by using novel mixed collector system. Trans. Nonferr. Met. Soc. China 2019, 29, 1082–1089. [Google Scholar] [CrossRef]
  32. Wang, Z.; Xu, L.H.; Wang, J.M.; Wang, L.; Xiao, J.H. A comparison study of adsorption of benzohydroxamic acid and amyl xanthate on smithsonite with dodecylamine as co-collector. Appl. Surf. Sci. 2017, 426, 1141–1147. [Google Scholar] [CrossRef]
  33. Gao, Y.; Fu, X.Z.; Khoso, S.A.; Pan, Z.J.; Han, H.S.; Sun, W.; Yue, T. A quantitative innovation perspective on synergism and selectivity mechanism of mixed collectors in flotation. Miner. Eng. 2024, 206, 108474. [Google Scholar] [CrossRef]
  34. Deng, R.R.; Wang, Y.; Duan, W.T.; Xing, D.Q.; Hu, Y. Induced crystallization of Pb2+ on smithsonite surface during sulfidation-xanthate flotation. Colloids Surf. 2022, 650, 129576. [Google Scholar] [CrossRef]
  35. Wen, S.M. Thermodynamic theory of flotation for a complex multiphase solid–liquid system and high-entropy flotation. Int. J. Min. Met. Mater. 2024, 31, 1177–1197. [Google Scholar] [CrossRef]
  36. Jia, K.; Feng, Q.M.; Zhang, G.F.; Ji, W.Y.; Zhang, W.K.; Yang, B.Q. The role of S(II) and Pb (II) in xanthate flotation of smithsonite: Surface properties and mechanism. Appl. Surf. Sci. 2018, 442, 92–100. [Google Scholar] [CrossRef]
  37. Wang, M.T.; Zhang, G.F.; Chen, Y.F.; Zhao, L. Effect of surface oxidization on quartz slime coating in the sulfidization-amine flotation of smithsonite. Miner. Eng. 2022, 188, 107847. [Google Scholar] [CrossRef]
  38. Cao, S.M.; Zhang, X.L.; Zeng, S.Q. The present situation and progress of zinc oxide flotation. Adv. Mater. Res. 2013, 756, 68–71. [Google Scholar] [CrossRef]
  39. Liao, R.P.; Wen, S.M.; Liu, J.; Bai, S.J.; Feng, Q.C. Synergetic adsorption of dodecylamine and octyl hydroxamic acid on sulfidized smithsonite: Insights from experiments and molecular dynamics simulation. Sep. Purif. Technol. 2024, 329, 125106. [Google Scholar] [CrossRef]
  40. Krystian, C.; Cezary, R.; Grazyna, P. Flotation of zinc and lead oxide minerals from Olkusz region calamine ores. E3S Web Conf. 2016, 8, 01042. [Google Scholar] [CrossRef]
  41. Sun, Q.; Feng, Q.M.; Shi, Q. Adsorption mechanism of smithsonite by dodecyl phosphate ester potassium. J. Cent. South Univ. 2018, 49, 1845–1850. [Google Scholar] [CrossRef]
  42. Arturo, B.T.; Roberto, P.G.; Diego, M.C. Zeta potential of air bubbles conditioned with typical froth flotation reagents. Int. J. Min. Process 2015, 140, 50–57. [Google Scholar] [CrossRef]
  43. Wang, M.L.; Zhao, W.J.; Han, G.; Feng, Q.C. Utilization of lead ions to improve surface hydrophobicity and flotation recovery of sulfidized smithsonite. Colloids Surf. 2023, 663, 131126. [Google Scholar] [CrossRef]
  44. Zhang, G.F.; Cui, M.M.; Zhu, Y.G.; Shi, Q.; Luo, N. Effect of water glass on flotation separation of smithsonite and quartz. Chin. J. Nonferrous Met. 2012, 22, 3535–3541. [Google Scholar]
  45. Luo, B.; Liu, J.; Liu, Q.; Song, C.; Yu, L.; Li, S.; Lai, H. A Mechanism for the Adsorption of 2-(Hexadecanoylamino)Acetic Acid by Smithsonite: Surface Spectroscopy and Microflotation Experiments. Minerals 2019, 9, 15. [Google Scholar] [CrossRef]
  46. Zeng, Y.; Liu, J.; Dong, W.C.; Hao, J.M.; Wang, Y. Study on Sulfide Layer Attenuation Behavior of Smithsonite During Sulfidization Flotation. Front. Mater. 2019, 6, 347. [Google Scholar] [CrossRef]
  47. Zeng, Y.H.; Yao, X.; Liu, G.X.; He, G.M.; Yu, X.Y.; He, G.C.; Huang, Z.Q.; Zhang, R.R.; Cheng, C. Flotation behavior and mechanism of phenylpropenyl hydroxamic acid for the separation of smithsonite and calcite. J. Mol. Liq. 2021, 339, 116893. [Google Scholar] [CrossRef]
  48. Zhao, L.; Liu, W.G.; Liu, W.B.; Zhou, S.J.; Peng, X.Y. Investigation on matching relationship between surface characters and collector properties: Achieving flotation separation of zinc oxide minerals from quartz. Colloids Surf. A 2021, 617, 126392. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of smithsonite sample.
Figure 1. XRD pattern of smithsonite sample.
Minerals 14 00902 g001
Figure 2. Effect of grinding fineness on zinc flotation.
Figure 2. Effect of grinding fineness on zinc flotation.
Minerals 14 00902 g002
Figure 3. Effect of reagent dosage on zinc flotation: (a) Na2S; (b) KG-248; (c) DDA; (d) GS-2.
Figure 3. Effect of reagent dosage on zinc flotation: (a) Na2S; (b) KG-248; (c) DDA; (d) GS-2.
Minerals 14 00902 g003
Figure 4. Flow chart of the new flotation open-circuit test.
Figure 4. Flow chart of the new flotation open-circuit test.
Minerals 14 00902 g004
Figure 5. Effects of (a) pH and (b) Na2S concentration on smithsonite floatability.
Figure 5. Effects of (a) pH and (b) Na2S concentration on smithsonite floatability.
Minerals 14 00902 g005
Figure 6. Effect of collector concentration on smithsonite floatability: (a) DDA; (b) NaIX; (c) ADD.
Figure 6. Effect of collector concentration on smithsonite floatability: (a) DDA; (b) NaIX; (c) ADD.
Minerals 14 00902 g006
Figure 7. Zeta potentials of smithsonite (Sm) under different collector systems.
Figure 7. Zeta potentials of smithsonite (Sm) under different collector systems.
Minerals 14 00902 g007
Figure 8. Adsorption capacity of collectors in different systems vs. concentrations of (a) DDA, (b) NaIX, and (c) ADD.
Figure 8. Adsorption capacity of collectors in different systems vs. concentrations of (a) DDA, (b) NaIX, and (c) ADD.
Minerals 14 00902 g008
Figure 9. Proportion of elements on smithsonite surface treated with (a) Na2S, (b) Na2S + DDA, (c) Na2S + DDA + NaIX, and (d) Na2S + DDA + NaIX + ADD.
Figure 9. Proportion of elements on smithsonite surface treated with (a) Na2S, (b) Na2S + DDA, (c) Na2S + DDA + NaIX, and (d) Na2S + DDA + NaIX + ADD.
Minerals 14 00902 g009aMinerals 14 00902 g009b
Figure 10. N 1s XPS spectra of smithsonite surface treated with (a) Na2S; (b) Na2S + DDA; (c) Na2S + DDA + NaIX; and (d) Na2S + DDA + NaIX + ADD.
Figure 10. N 1s XPS spectra of smithsonite surface treated with (a) Na2S; (b) Na2S + DDA; (c) Na2S + DDA + NaIX; and (d) Na2S + DDA + NaIX + ADD.
Minerals 14 00902 g010
Figure 11. Contact angles of smithsonite surface treated with (a) deionized water, (b) Na2S + DDA, (c) Na2S + DDA + NaIX, and (d) Na2S + DDA + NaIX + ADD.
Figure 11. Contact angles of smithsonite surface treated with (a) deionized water, (b) Na2S + DDA, (c) Na2S + DDA + NaIX, and (d) Na2S + DDA + NaIX + ADD.
Minerals 14 00902 g011
Table 1. Results of the new flotation open-circuit test.
Table 1. Results of the new flotation open-circuit test.
ProductYield, %Zn Grade, %Zn Recovery, %
Pb-c4.894.962.97
Zn-c24.5228.7186.24
Tailings70.591.2510.79
Sum100.008.16100.00
Table 2. Economic estimates for reagents.
Table 2. Economic estimates for reagents.
NamePrice, ¥/tDosage, g/tPrice, ¥
Na2CO380010000.8
Na2S220015,00033
NaIX10,0003003
KG-24815,5004607.13
DDA17,8003005.34
GS-212,5004005
2# oil6000400.24
Sum--54.51
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

Li, Z.; Feng, Q.; Zhang, Q.; Wen, S. Open-Circuit Technology of Zinc Oxide Ore Flotation with Ternary Collector and Its Adsorption Characteristics on Smithsonite Surface. Minerals 2024, 14, 902. https://doi.org/10.3390/min14090902

AMA Style

Li Z, Feng Q, Zhang Q, Wen S. Open-Circuit Technology of Zinc Oxide Ore Flotation with Ternary Collector and Its Adsorption Characteristics on Smithsonite Surface. Minerals. 2024; 14(9):902. https://doi.org/10.3390/min14090902

Chicago/Turabian Style

Li, Zhiwei, Qicheng Feng, Qian Zhang, and Shuming Wen. 2024. "Open-Circuit Technology of Zinc Oxide Ore Flotation with Ternary Collector and Its Adsorption Characteristics on Smithsonite Surface" Minerals 14, no. 9: 902. https://doi.org/10.3390/min14090902

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