Next Article in Journal
Hyperspectral Rock Classification Method Based on Spatial-Spectral Multidimensional Feature Fusion
Previous Article in Journal
Comparative Analysis of the Recovery of Cu2+ and Au from Washing Solution of Pyrite Concentrate Slag by Two Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Efficient Separation of Apatite from Dolomite Using Fucoidan as an Eco-Friendly Depressant

1
School of Resources and Safety Engineering, Wuhan Institute of Technology, Wuhan 430073, China
2
Hubei Three Gorges Laboratory, Yichang 443008, China
3
Technology Innovation Center for Comprehensive Utilization of Strategic Mineral Resources, Ministry of Natural Resources, Chengdu 610041, China
*
Authors to whom correspondence should be addressed.
Minerals 2024, 14(9), 922; https://doi.org/10.3390/min14090922
Submission received: 24 July 2024 / Revised: 6 September 2024 / Accepted: 8 September 2024 / Published: 10 September 2024
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)

Abstract

:
The aim is to explore new depressants for achieving the efficient separation of apatite and dolomite. In this work, fucoidan (FD) was examined as an eco-friendly dolomite depressant to separate dolomite from apatite. The depression ability and adsorption mechanisms were investigated. The flotation results indicated that FD selectively depressed dolomite. The flotation difference between dolomite and apatite reached 70% approximately at an FD concentration of 75 mg/L. Meanwhile, the recovery and grade of P2O5 reached 89.84% and 32.88% and that of MgO decreased to 1.64% and 34.24% in the artificially mixed minerals test. Wettability, zeta potential, and Fourier transform infrared spectroscopy (FTIR) results revealed that FD tended to adsorb onto dolomite, impeding the interaction of sodium oleate (NaOL) with dolomite, but barely affected that on apatite. Microcalorimetry analysis indicated that the adsorption heat of FD on dolomite was much higher and less time was required to achieve equilibrium. X-ray photoelectron spectroscopy (XPS) results proved that the sulfonic acid radicals within FD chemically interacted with Mg atoms on dolomite while it weakly adsorbed on apatite.

1. Introduction

As important strategic resources to extract phosphorus, phosphate ores play an indispensable role in the production of fertilizers, which have also been widely applied in many fields, such as food, defense, medicine, and new energy [1,2]. With the continuous depletion of phosphate ores, the exploration and utilization of low-grade phosphate ores are imperative to meet the constantly increasing demand for phosphorus elements, but unfortunately, most of the low-grade phosphate ores have not been explored yet due to the undeveloped reagents and techniques [3,4,5]. One of the important reasons is that the apatite generally coexists closely with carbonate and silicate minerals in these ores, in particular with dolomite, making the efficient separation difficult.
Froth flotation, as the major method to beneficiate phosphate ores, enables the selective separation of phosphates from calcium-containing gangue minerals through direct or reverse flotation [6]. Through the action of some specific surfactants, the hydrophobicity of mineral particles can be intentionally adjusted, based on which the different minerals have been separated by their hydrophobicity properties [7,8]. Fatty acid collectors are commonly used in the flotation of phosphate ores because of their low cost and good availability [9]. However, it is unable to separate them efficiently without the aid of depressants, because they possess similar surface properties and the selectivity of fatty acids for calcium-containing minerals is poor. In practice, the reverse flotation techniques are commonly adopted to upgrade phosphate ores, in which apatite depressants such as H2SO4 and H3PO4 are used to depress apatite while dolomite is floated by fatty acid collectors [10]. However, these reagents have significant disadvantages in many aspects; they not only cause serious corrosion damage to the flotation equipment and pipelines but also lead to serious environmental pollution and eutrophication.
To overcome the shortcomings of reverse flotation, much attention has been paid to exploring dolomite depressants to substitute conventional apatite depressants in direct flotation. Wang et al. proposed that tamarind seed gum (TSG) interacted with the dolomite surface by chemical bonding between the -OH group and Ca and Mg sites [11]. Du et al. [12] noted that carboxymethyl cellulose (CMC) could selectively depress dolomite without affecting apatite, which preferred to adsorb onto the dolomite surface because the two O atoms in the carboxyl group could combine with Ca and Mg atoms to generate covalent and ionic bonds, respectively, creating more stable bridge adsorption. Yang [13] discovered that hydrolytic polymaleic anhydride (HPMA) could produce stronger chelation with Mg2+ than Ca2+, which was the main reason why HPMA selectively depressed dolomite but did not influence the flotation of apatite. These novel depressants show good selective depression effect on dolomite but have not been widely used in practice yet, probably due to their limitation of availability, cost, and environmental pollution. Therefore, it is still a great challenge to explore novel depressants that are more efficient, cheap, and biocompatible for the recovery of low-grade phosphate ores.
As a water-soluble sulfated polysaccharide, FD was derived from brown algae resources, which has gained significant attention in feed processing and medical research due to its antioxidant and anticancer properties [14,15,16]. Meanwhile, because of the abundant algae resources in the ocean, the production of FD is continuously guaranteed [17]. Figure 1 illustrates the molecular structure of FD. Within molecular FD, there are many multifunction hydroxyl and sulfate groups, which may easily interact with metal ions such as calcium and magnesium. Moreover, due to its good availability and non-toxicity, it may be a good choice to depress dolomite in the flotation separation of phosphate ores.
In this work, the biodegradable polysaccharide FD has been used as a dolomite depressant to upgrade phosphate ores through direct flotation. Furthermore, the inhibition mechanism of FD in the selective flotation system was uncovered by means of wettability characterization, Zeta potential measurement, FTIR detection, microcalorimetry determination, and XPS analysis. This work aimed to develop a novel dolomite depressant in the direct flotation of phosphate ores and shed some light on the development of a dolomite depressant in the direct flotation of phosphate ores.

2. Experiment

2.1. Materials and Reagents

Dolomite and apatite mineral samples from Guizhou province were selected manually first and then ground and sieved separately. Mineral samples between 38 and 74 μm were subject to XRD and flotation tests, and mineral sample fines smaller than 38 μm were used to investigate the flotation mechanism. The apatite sample had a P2O5 grade of 38.77% determined through the volumetric titration method using ammonium phosphomolybdate, and the dolomite sample had a MgO grade of 22.42% determined via the EDTA volumetric method. XRD results (Figure 2) of the samples matched well with standard reference cards of PDF#15-0876 and PDF#36-042.
Fucoidan (Macklin, China) with a molecular weight of 242.25 was tested as a dolomite inhibitor. The analytical grade of sodium oleate (NaOL, collector), HCl, and NaOH (pH modifier) was acquired from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Ultrapure water with a resistivity of more than 18.2 MΩ·cm was used for flotation and mechanism study.

2.2. Flotation Tests

An XFG-type flotation machine equipped with a 40 mL plexiglas cell was used for micro-flotation tests. About 2.0 g samples were dispersed in 30 mL of ultrapure water for 2 min first; then, pH modifier FD and NaOL solutions were introduced sequentially at 5 min intervals. The specific flotation flowsheet is illustrated in Figure 3. After aeration, the floated minerals and sinking minerals were collected, weighted, and assayed to evaluate the flotation results. Each test was carried out three times and the average was reported as the ultimate result. In the case of mixed minerals, 1.6 g apatite and 0.4 g dolomite samples were mixed and adopted for flotation; the experiment was performed as illustrated in Figure 3.

2.3. Contact Angle Measurement

The wettability of minerals surface treated and untreated was determined by the captive bubble method using a contact angle goniometer (JC2000DM, Beijing Zhongyi Kexin Technology Co., Ltd., Beijing, China). The surfaces of apatite and dolomite lumps were first polished using sandpaper and a polishing cloth. After that, the polished surface of mineral samples was soaked in a container with 110 mL of various solutions. Then, an air bubble (around 5 μL) was generated and placed onto the mineral surface using a microsyringe. At equilibrium, the image was captured, and the contact angle results were determined. For each sample, three different locations were measured, and the mean value was reported.

2.4. Zeta Potential Determination

Malvern Zeta Sizer (Nano90, Malvern Panalytical, Malvern, UK) was employed to measure zeta potential using the Smoluchowski equation; 0.2 g of minerals were dispersed into 50 mL of 0.001 mol/L KNO3 solution first and then the pulp pH was adjusted. After that, FD or NaOL solutions were introduced and stirred for 10 min. Finally, a small portion of the pulp was taken for zeta potential determination. To ensure accuracy and reliability, each sample was tested three times.

2.5. FTIR Determination

The FTIR measurement was carried out in transmission mode using a Nicolet 6700 spectrometer (Thermo, Waltham, MA, USA); 2.0 g of sample was uniformly dispersed and stirred in 70 mL solution (75 mg/L FD or 75 mg/L FD with 3 ×10−4 mol/L NaOL) for 5 min. After adsorption, the filtered samples were freeze-dried for FTIR tests.

2.6. Microcalorimetry Analysis

Microcalorimetry analysis was conducted using a Microcalorimeter (C80). About 100 mg of pure samples (−74 μm) and 1 mL of FD solution were placed in the upper and lower calorimetric cells, which were separated by membranes, respectively. After the film was pierced, the change in heat flow after the interaction between the reagent and minerals was recorded. The measurement was terminated, and the results were obtained once the signal stabilized. The specific surface areas of dolomite and apatite samples were determined using the BET method, which were 0.1827 m2/g and 0.3411 m2/g, respectively.

2.7. XPS Analysis

Escalab 250Xi photoelectron spectrometer (Thermo, Waltham, MA, USA) was used for XPS analysis. The samples were treated in the same manner as mentioned in Section 2.5. The C1s peak at 284.80 eV was used as the standard to calibrate the spectra.

3. Results and Discussion

3.1. Micro-Flotation of Single Mineral

3.1.1. Influence of FD Concentration

The flotation performances as affected by FD dosage are illustrated in Figure 4. Without the addition of depressants, both apatite and dolomite floated well and it was hard to separate them. Once FD was employed, the floatability of dolomite dropped significantly; in contrast, the flotation of apatite remained around 90% as FD concentration increased. The flotation difference was expanded to 70% at an FD concentration of 50 mg/L, illustrating that FD may be a potential dolomite depressant capable of selectively depressing dolomite in direct flotation.

3.1.2. Effect of pH

Figure 5 shows the impact of pH on the flotation with the addition of FD. Dolomite was strongly depressed at pH less than 10, above which, it increased slightly. As for apatite, it was slightly depressed but still maintained high floatability at a weak alkaline medium. However, it regained its floatability as pH increased above 10. The optimized difference (approximately 70%) was gained at pH of 10, which provided an optimum separation window for these two minerals. These results reconfirmed that FD could be a good dolomite depressant to upgrade phosphate ores by means of direct flotation, which provided a good option for recovering apatite.

3.1.3. Effect of NaOL Concentration

Figure 6 shows the depression ability of FD as related to NaOL concentration. Dolomite was strongly depressed at NaOL dosage less than 3 ×10−4 mol/L, but it ascended significantly as NaOL concentration exceeded 3 ×10−4 mol/L. As for apatite, the flotation was almost not influenced by NaOL concentration. This phenomenon may be attributed to the fact that high NaOL concentration increased the possibility of NaOL species bonding with Ca and Mg sites on the dolomite surface, thereby weakening the depression ability of FD to dolomite. These results suggested that high collector concentration could worsen the separation performance of FD on dolomite and apatite.

3.1.4. Effect of Temperature

Figure 7 represents the influence of pulp temperature on flotation. The recovery of apatite remained around 90% in the whole temperature range. For dolomite, the floatability gradually increased as temperature increased from 10 to 50 °C, which may be attributed to the fact that high temperature promoted the adsorption of NaOL on the dolomite surface, which deteriorated the selectivity of FD [18].

3.2. Flotation of Artificially Mixed Mineral

To further confirm the selectivity of FD in flotation, the flotation of artificially mixed minerals was conducted; the results are shown in Figure 8. In the absence of FD, more than 70% dolomite was floated. But as FD was introduced, both the grade and recovery of MgO dropped significantly; meanwhile, the grade and recovery of P2O5 barely changed at an FD concentration of less than 75 mg/L. When FD concentration was 75 mg/L, the mixed ore achieved a better separation effect, at which the recovery and grade of P2O5 reached 89.84% and 32.88% and that of MgO decreased to 1.64% and 34.24%, respectively. These results suggested that FD was able to depress dolomite selectively using NaOL as a collector.

3.3. Contact Angle Results

Figure 9 indicates the wettability image of mineral surface conditioned in various solutions, and the results are shown in Table 1. For pure dolomite and apatite treated with ultrapure water, the contact angle was 28.58° and 26.75°, respectively. Upon the addition of 3 × 10−4 mol/L NaOL, the contact angle of the dolomite and apatite surface surged to 73.67° and 81.92°, respectively. After the addition of 75 mg/L FD and 3 × 10−4 mol/L NaOL, the contact angle of dolomite dramatically decreased to 30.08°. In comparison, the contact angle of apatite decreased slightly to 60.50°. The contact angle results proved that FD could greatly enlarge the hydrophobicity difference between apatite and dolomite in NaOL solutions, possibly due to the differential affinity of FD and NaOL toward the mineral surface.

3.4. Zeta Potential Results

Figure 10 presents the zeta potential of dolomite and apatite with different reagents. In Figure 10, after the introduction of FD of NaOL alone, the absolute value of zeta potential for both minerals increased dramatically, which means that both FD and NaOL could interact with apatite and dolomite individually without selectivity [19]. In Figure 10a, after the addition of both reagents, the zeta potential of dolomite approached the value conditioned by FD, which suggested that the adsorption of FD on dolomite was predominant. Compared with dolomite, the zeta potential for apatite (conditioned with FD and NaOL) was close to that conditioned with NaOL alone, as indicated in Figure 10b, which means that the adsorption of NaOL prevailed. These results reflect the preferential adsorption of FD and NaOL on dolomite and apatite surfaces, which confirmed the flotation results.

3.5. FTIR Analysis Results

Figure 11 displays the FTIR results under different experimental conditions. For dolomite and apatite treated with NaOL alone, the methylene and methyl groups originating from NaOL occurred around 2925 cm−1 and 2854 cm−1 [20]. Under the circumstance where both reagents were added, the -CH2 and -CH3 groups of NaOL vanished (Figure 11a), while they still existed on the apatite surface (Figure 11b). The results showed that FD blocked NaOL from interacting with the dolomite surface but did not influence the interaction of NaOL with apatite.

3.6. Microcalorimetry Determination

Figure 12 represents the adsorption heat of FD on apatite and dolomite. It is obvious that the reaction between reagent and mineral was exothermic. By comparing these results, the adsorption heat between FD and dolomite was −1.598 J/m2 before it reached equilibrium at 0.71 h (Figure 12a). While adsorption heat between FD and apatite was −0.794 J/m2, 8 h was required to achieve the equilibrium, as illustrated in Figure 12b. These results proved that the reaction between FD and dolomite was much stronger and faster than that of apatite [21].

3.7. XPS Analysis Results

Figure 13 displays the spectra of Ca2p, O1s, Mg1s, and S2p on dolomite surface. In Figure 13a, the peaks at 347.24 eV and 350.82 eV were associated with Ca2p 3/2 and Ca2p 1/2 in dolomite, respectively [22,23]. After the introduction of FD, a weak shift (0.03 eV) of Ca2p was observed, indicating that Ca atoms did not take part in the reaction with FD [24,25]. In Figure 13b, the peak of CO32− almost did not change after the addition of FD, indicating that CO32− in dolomite did not interact with FD. In Figure 13c, the Mg1s peak of dolomite moved downward by −0.21 eV after the addition of FD [26], which proved that Mg atoms in dolomite chemically interacted with FD [27]. In Figure 13d, the S2p 3/2 and S2p 1/2 peaks of -SO3 appeared at 168.49 eV and 169.65 eV after adsorption [28].
Figure 14 shows the high-resolution spectra of apatite with or without FD. In Figure 14a, the two peaks corresponded to Ca2p 3/2 (347.53 eV) and Ca2p 1/2 (351.06 eV) shifted just by 0.03 eV after adsorption, which means that the Ca atoms may not bond chemically with FD. In Figure 14b–d, the O1s, F1s, and P2p orbital almost did not move after the addition of FD, suggesting that FD may not interact with these atoms on the apatite surface. In addition, Figure 14e showed that the S2p signal on apatite became very vague after the treatment with FD, proving that -SO3 groups within FD had a weak affinity toward the apatite surface. Overall, the XPS results demonstrated that the -SO3 groups in FD may strongly interact with Mg sites chemically on the dolomite surface but physically adsorbed onto the apatite surface, which altered the adsorption of NaOL and resulted in separation.

4. Conclusions

The flotation performance and depression mechanism of a new dolomite depressant FD were studied in this work. The flotation results indicated that FD exhibited good selectivity in the beneficiation of apatite from dolomite using NaOL as a collector. At an FD dosage of 75 mg/L, the flotation difference between dolomite and apatite reached approximately 70% in single minerals flotation tests. Meanwhile, the recovery and grade of P2O5 reached 89.84% and 32.88% and that of MgO decreased to 1.64% and 34.24% in the artificially mixed minerals test. Wettability, zeta potential, FTIR, and microcalorimetry results confirmed that FD had a stronger affinity toward dolomite compared with apatite. XPS analysis showed that FD could chemically adsorb onto dolomite due to the strong interaction between sulfonic acid radicals and Mg sites, but it physically adsorbed onto the apatite surface.

Author Contributions

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

Funding

The authors gratefully appreciated the financial funds from the Hubei Science Foundation for Distinguished Young Scholars (2023AFA047), National Natural Science Foundation of China (52374272), Sino-German Center (GZ 1693), and Technology Innovation Center for Comprehensive Utilization of Strategic Mineral Resources, Ministry of Natural Resources (CCUM-KY-2309).

Data Availability Statement

Data are available on request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abdel-Hakeem, M.; El-Habaak, G. The potential production of rock-based fertilizer and soil conditioner from phosphate mine wastes: A case study from Abu-Tartur plateau in the Western Desert of Egypt. J. Clean. Prod. 2021, 329, 129761. [Google Scholar] [CrossRef]
  2. Lopes, C.M.; Miranda Silva, A.M.; Estrada-Bonilla, G.A.; Ferraz-Almeida, R.; Vieira, J.L.V.; Otto, R.; Vitti, G.C.; Cardoso, E.J.B.N. Improving the fertilizer value of sugarcane wastes through phosphate rock amendment and phosphate-solubilizing bacteria inoculation. J. Clean. Prod. 2021, 298, 126821. [Google Scholar] [CrossRef]
  3. Wei, Z.; Zhang, Q.; Wang, X. New insights on depressive mechanism of citric acid in the selective flotation of dolomite from apatite. Colloid Surface A 2022, 653, 130075. [Google Scholar] [CrossRef]
  4. Zeng, M.; Yang, B.; Guan, Z.; Zeng, L.; Luo, H.; Deng, B. The selective adsorption of xanthan gum on dolomite and its implication in the flotation separation of dolomite from apatite. Appl. Surf. Sci. 2021, 551, 149301. [Google Scholar] [CrossRef]
  5. Zhong, C.; Feng, B.; Zhang, W.; Zhang, L.; Guo, Y.; Wang, T.; Wang, H. The role of sodium alginate in the flotation separation of apatite and dolomite. Powder Technol. 2020, 373, 620–626. [Google Scholar] [CrossRef]
  6. Zeng, L.; Ding, K.; Zhang, X.; Zhou, Y.; Han, H. New insights into the influence of mineral surface transformation on the flotation behavior of anhydrite/apatite. Colloid Surface A 2024, 685, 133215. [Google Scholar] [CrossRef]
  7. Santos, L.H.; Santos, G.O.; Magalhães, L.F.; Silva, G.R.; Peres, A.E.C. Application of andiroba oil as a novel collector in apatite flotation. Miner. Eng. 2022, 185, 107678. [Google Scholar] [CrossRef]
  8. Jing, L.; Xu, L.; Xue, K.; Wang, D.; Ma, Z.; Meng, J.; Shi, X.; Liu, C. Selective depression by using environment-friendly depressant pectin in apatite and dolomite flotation system. Miner. Eng. 2023, 203, 108373. [Google Scholar] [CrossRef]
  9. Derhy, M.; Taha, Y.; El-Bahi, A.; Ait-Khouia, Y.; Benzaazoua, M.; Hakkou, R. Selective flotation of calcite and dolomite from apatite using bio-based alternatives to conventional collectors: Castor and mustard oils. Miner. Eng. 2024, 208, 108597. [Google Scholar] [CrossRef]
  10. Liu, X.; Ruan, Y.; Li, C.; Cheng, R. Effect and mechanism of phosphoric acid in the apatite/dolomite flotation system. Int. J. Miner. Process. 2017, 167, 95–102. [Google Scholar] [CrossRef]
  11. Wang, X.; Xie, R.; Liu, J.; Zhu, Y. The utilization of tamarind seed gum as a novel dolomite depressant in the selective flotation of apatite from dolomite. Adv. Powder Technol. 2023, 34, 104022. [Google Scholar] [CrossRef]
  12. Du, W.; Li, X. Insight into the inhibition mechanism of carboxymethyl cellulose for flotation of dolomite and fluorapatite: Experimental and DFT studies. Colloid Surface A 2023, 674, 131957. [Google Scholar] [CrossRef]
  13. Yang, B.; Yin, W.; Zhu, Z.; Sun, H.; Sheng, Q.; Fu, Y.; Yao, J.; Zhao, K. Differential adsorption of hydrolytic polymaleic anhydride as an eco-friendly depressant for the selective flotation of apatite from dolomite. Sep. Purif. 2021, 256, 117803. [Google Scholar] [CrossRef]
  14. Fawzy, M.A.; Gomaa, M. Pretreated fucoidan and alginate from a brown seaweed as a substantial carbon source for promoting biomass, lipid, biochemical constituents and biodiesel quality of Dunaliella salina. Renew. Energy 2020, 157, 246–255. [Google Scholar] [CrossRef]
  15. Cholaraj, R.; Venkatachalam, R. Investigation of antioxidant and anticancer potential of fucoidan (in-vitro & in-silico) from brown seaweed Padina boergesenii. Algal. Res. 2024, 79, 103442. [Google Scholar]
  16. Lan, Y.; Qin, K.; Wu, S. The physiological activities of fucoidan and its application in animal breeding. Fish Shellfish Immun. 2024, 147, 109458. [Google Scholar] [CrossRef]
  17. Yin, H.; Zhang, Q.; Farooqi, A.A.; Wang, J.; Yue, Y.; Geng, L.; Wu, N. Opportunities and challenges of fucoidan for tumors therapy. Carbohydr. Polym. 2024, 324, 121555. [Google Scholar]
  18. Wang, L.; Li, Z.; Zhang, H.; Huang, L.; Zhu, Y.; Li, F. Effect of Pulp Temperature on Separation of Magnesite from Dolomite in Sodium Oleate Flotation System. Miner. Eng. 2024, 212, 108718. [Google Scholar] [CrossRef]
  19. Cao, Q.; Zou, H.; Chen, X.; Wen, S. Flotation selectivity of N-hexadecanoylglycine in the fluorapatite–dolomite system. Miner. Eng. 2019, 131, 353–362. [Google Scholar] [CrossRef]
  20. Liu, M.; Li, H.; Jiang, T.; Liu, Q. Flotation of coarse and fine pyrochlore using octyl hydroxamic acid and sodium oleate. Miner. Eng. 2019, 132, 191–201. [Google Scholar] [CrossRef]
  21. Kuang, J.; Wang, X.; Yu, M.; Yuan, W.; Huang, Z.; Zhang, S.; Xiao, J. Flotation behavior and mechanism of calcium carbonate polymorphs with sodium oleate as collector. Powder Technol. 2022, 398, 117040. [Google Scholar] [CrossRef]
  22. Huang, W.; Liu, W.; Zhong, W.; Chi, X.; Rao, F. Effects of common ions on the flotation of fluorapatite and dolomite with oleate collector. Miner. Eng. 2021, 174, 107213. [Google Scholar] [CrossRef]
  23. Gong, X.; Yao, J.; Yang, B.; Guo, J.; Sun, H.; Yin, W. Study on the inhibition mechanism of guar gum in the flotation separation of brucite and dolomite in the presence of SDS. J. Mol. Liq. 2023, 380, 121721. [Google Scholar] [CrossRef]
  24. Gong, X.; Yao, J.; Yang, B.; Yin, W.; Wang, Y.; Fu, Y. Adsorption mechanism of green efficient chelating poly-L-aspartic acid in flotation separation of brucite and dolomite. Adv. Powder Technol. 2023, 34, 104207. [Google Scholar] [CrossRef]
  25. Ge, R.; Yang, B.; Martin, R.; Luo, H.; He, D.; Arroyo, M.A.C. The application of poly (sodium 4-styrene sulfonate) as an efficient dolomite depressant in the direct flotation of apatite from dolomite. Powder Technol. 2023, 425, 118583. [Google Scholar] [CrossRef]
  26. Liu, J.; Tao, Y.; Chang, T.; Ge, W.; Jiang, K.; Lv, L.; Zhu, Y.; Yuan, S. Study on flotation separation of barite fluorite by citric acid under new collector system. Colloid Surface A 2024, 692, 134058. [Google Scholar] [CrossRef]
  27. Lv, L.; Duan, H.; Liu, W.; Yue, T. Flotation separation of fluorite from calcite using bis hydroxamic acid collector. Miner. Eng. 2022, 187, 107803. [Google Scholar] [CrossRef]
  28. Ding, Z.; Li, J.; Bi, Y.; Yu, P.; Dai, H.; Wen, S.; Bai, S. The adsorption mechanism of synergic reagents and its effect on apatite flotation in oleamide-sodium dodecyl benzene sulfonate (SDBS) system. Miner. Eng. 2021, 170, 107070. [Google Scholar] [CrossRef]
Figure 1. The molecular structure of FD.
Figure 1. The molecular structure of FD.
Minerals 14 00922 g001
Figure 2. The XRD patterns of samples: dolomite (a) and apatite (b).
Figure 2. The XRD patterns of samples: dolomite (a) and apatite (b).
Minerals 14 00922 g002
Figure 3. The flotation flowsheet of single and mixed minerals.
Figure 3. The flotation flowsheet of single and mixed minerals.
Minerals 14 00922 g003
Figure 4. The flotation affected by FD concentration.
Figure 4. The flotation affected by FD concentration.
Minerals 14 00922 g004
Figure 5. The flotation performances as affected by pH in the presence of FD.
Figure 5. The flotation performances as affected by pH in the presence of FD.
Minerals 14 00922 g005
Figure 6. The flotation performance affected by NaOL dosage.
Figure 6. The flotation performance affected by NaOL dosage.
Minerals 14 00922 g006
Figure 7. The flotation performance affected by temperature.
Figure 7. The flotation performance affected by temperature.
Minerals 14 00922 g007
Figure 8. The flotation grade and recovery of P2O5 and MgO for artificially mixed minerals.
Figure 8. The flotation grade and recovery of P2O5 and MgO for artificially mixed minerals.
Minerals 14 00922 g008
Figure 9. The contact angle results of minerals treated under different conditions.
Figure 9. The contact angle results of minerals treated under different conditions.
Minerals 14 00922 g009
Figure 10. Zeta potential of dolomite (a) and apatite (b) with or without reagent.
Figure 10. Zeta potential of dolomite (a) and apatite (b) with or without reagent.
Minerals 14 00922 g010
Figure 11. FTIR results of dolomite (a) and apatite (b).
Figure 11. FTIR results of dolomite (a) and apatite (b).
Minerals 14 00922 g011
Figure 12. Microthermal results of FD interaction with dolomite (a) and apatite (b).
Figure 12. Microthermal results of FD interaction with dolomite (a) and apatite (b).
Minerals 14 00922 g012
Figure 13. XPS spectra of dolomite samples with or without FD addition: Ca2p (a), O1s (b), Mg1s (c), and S2p (d).
Figure 13. XPS spectra of dolomite samples with or without FD addition: Ca2p (a), O1s (b), Mg1s (c), and S2p (d).
Minerals 14 00922 g013
Figure 14. XPS spectra of apatite samples: Ca2p (a), O1s (b), F1s (c), P2p (d), and S2p (e).
Figure 14. XPS spectra of apatite samples: Ca2p (a), O1s (b), F1s (c), P2p (d), and S2p (e).
Minerals 14 00922 g014aMinerals 14 00922 g014b
Table 1. Contact angle results of dolomite and apatite.
Table 1. Contact angle results of dolomite and apatite.
MineralsVarious Conditions
Ultrapure Water3 × 10−4 mol/L NaOL75 mg/L FD and 3 × 10−4 mol/L NaOL
Dolomite28.58° ± 1.36°73.67° ± 1.65°30.08° ± 3.48°
Apatite26.75° ± 0.94°81.92° ± 0.72°60.50° ± 3.36°
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

Zhang, Y.; Yang, B.; Deng, B.; Luo, H.; Zhou, F. The Efficient Separation of Apatite from Dolomite Using Fucoidan as an Eco-Friendly Depressant. Minerals 2024, 14, 922. https://doi.org/10.3390/min14090922

AMA Style

Zhang Y, Yang B, Deng B, Luo H, Zhou F. The Efficient Separation of Apatite from Dolomite Using Fucoidan as an Eco-Friendly Depressant. Minerals. 2024; 14(9):922. https://doi.org/10.3390/min14090922

Chicago/Turabian Style

Zhang, Yifan, Bingqiao Yang, Bing Deng, Huihua Luo, and Fang Zhou. 2024. "The Efficient Separation of Apatite from Dolomite Using Fucoidan as an Eco-Friendly Depressant" Minerals 14, no. 9: 922. https://doi.org/10.3390/min14090922

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