Next Article in Journal
The Effects of Sex, Women’s Body Composition and Monthly Cycle Phases on the Sensory Threshold of Upper Limb to Transcutaneous Electrical Nerve Stimulation in Healthy Subjects
Next Article in Special Issue
Analysis of Feed Inlet and Optimal Feeding Amount of Waste Ground Film Impurity Removal Equipment
Previous Article in Journal
OpenFog-Compliant Application-Aware Platform: A Kubernetes Extension
Previous Article in Special Issue
Application of Fly Ash Obtained from the Incineration of Municipal Solid Waste in Agriculture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rare Earth Elements Recovery from Primary and Secondary Resources Using Flotation: A Systematic Review

by
Pongsiri Julapong
1,2,
Apisit Numprasanthai
1,
Ladda Tangwattananukul
3,
Onchanok Juntarasakul
1,
Palot Srichonphaisarn
1,
Kosei Aikawa
4,
Ilhwan Park
4,
Mayumi Ito
4,
Carlito Baltazar Tabelin
5,6 and
Theerayut Phengsaart
1,*
1
Department of Mining and Petroleum Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
2
Department of Mining and Materials Engineering, Faculty of Engineering, Prince of Songkla University, Songkhla 90110, Thailand
3
Department of Earth Sciences, Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
4
Division of Sustainable Resources Engineering, Faculty of Engineering, Hokkaido University, Sapporo 060-8628, Japan
5
Department of Materials and Resources Engineering Technology, College of Engineering, Mindanao State University–Iligan Institute of Technology, Iligan City 9200, Philippines
6
Resource Processing and Technology Center, Research Institute of Engineering and Innovative Technology, Mindanao State University–Iligan Institute of Technology, Iligan City 9200, Philippines
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(14), 8364; https://doi.org/10.3390/app13148364
Submission received: 26 June 2023 / Revised: 16 July 2023 / Accepted: 18 July 2023 / Published: 19 July 2023
(This article belongs to the Special Issue Advances in Waste Treatment and Resource Utilization)

Abstract

:
Rare earth minerals (REMs) contain rare earth elements (REEs) that are important in modern technologies due to their unique magnetic, phosphorescent, and catalytic properties. However, REMs are not only non-renewable resources but also non-uniformly distributed on the Earth’s crust, so the processing of REE-bearing secondary resources via recycling is one potential route to ensure the long-term sustainability of REE supply. Flotation—a method that separates materials based on differences in their surface wettability—is a process applied for both mineral processing and recycling of REEs, especially when the particles are fine and/or a high-purity product is required. In this review, studies about rare earth flotation from 2012 to 2021 were systematically reviewed using the PRISMA guideline. It was found that most REM flotation research works focused on finding better collectors and depressants while, for recycling, studies on advanced flotation techniques like froth flotation, ion flotation, solvent sublation, electroflotation, and adsorbing colloid flotation with an emphasis on the recovery of dissolved REEs from aqueous solutions dominated.

1. Introduction

Rare earth elements (REEs)—a group composed of 15 lanthanides, scandium, and yttrium—are essential in numerous modern technologies because of their unique magnetic, phosphorescent, and catalytic properties. The bulk of REEs produced worldwide are used as catalysts (75%), while the rest are consumed by various industries, including ceramics and glass manufacturing (6%), polishing agents (5%), metallurgical applications and alloys (4%), and others (e.g., computer memory, DVDs, rechargeable batteries, super magnets, mobile phones, LED lighting, superconductors, fluorescent materials, phosphate-binding agents, solar panels, magnetic resonance imaging (MRI) agents) (10%) [1,2,3].
Although REEs occur in nature only in trace quantities, they are often concentrated in 250 minerals called rare earth minerals (REMs), which are classified into carbonates, phosphates, silicates, and oxides [1,4]. Despite this large number of REMs, however, there are only a few economically viable sources of REEs, including bastnaesite ((La, Ce, Y)CO3F), monazite ((La, Ce, Nd)PO4), and xenotime (YPO4), which occur in various geological ore deposits like (i) carbonatites, (ii) peralkaline igneous systems, (iii) magmatic magnetite-hematite bodies, (iv) iron oxide copper–gold (IOCG) deposits, (v) ion-adsorption clay deposits, and (vi) monazite-xenotime-bearing placer deposits [5,6].
In carbonatite deposits, the primary REMs are bastnaesite, parisite, synchysite, ancylite, and monazite, while the predominant gangue minerals are carbonate minerals (e.g., calcite, dolomite, ankerite, siderite), which may coexist with aegirine-augite, barite, fluorite, fluorapatite, hematite, magnetite, phlogopite, quartz, and/or oxide copper–gold strontianite. The world’s largest carbonatite deposit, which is found in China, is called the Bayan Obo iron-carbonatite deposit. In comparison, peralkaline igneous rock deposits make up a large group of rock types generally characterized as deficient in silicon dioxide (SiO2) relative to Na2O, K2O, and calcium oxide (CaO). Important REMs in this deposit include apatite, eudialyte, loparite, gittinsite, xenotime, gadolinite, monazite, bastnaesite, kainosite, mosandrite, britholite, allanite, and zircon associated with peralkaline intrusion. Meanwhile, REMs in magmatic magnetite-hematite body deposits are hosted by magnetite and hematite, most of which could be recovered as a by-product during iron ore mining. Iron oxide copper–gold deposits are similar to magmatic magnetite-hematite bodies because they are formed under magmatic–hydrothermal conditions. These iron-bearing deposits host not only copper and gold but also REMs, which are associated both with iron oxides and small carbonatites [7]. Aside from magmatic intrusive bodies, REEs can be naturally concentrated via prolonged leaching, weathering, adsorption/ion exchange, and precipitation processes, which lead to preferential accumulation of REEs like those found in ion-adsorption clay deposits and monazite-xenotime-bearing placer deposits. The largest deposit of the former is found in China, while the latter type of deposit is also classified as a heavy-mineral sand and is an important source of monazite and xenotime [5,6].
In 2021, the total global reserve of REEs was estimated at 120 million tons (Mt), reported as rare earth oxide (REO) equivalent content [2]. China accounted for the largest slice of this reserve at around 44 Mt (37%), followed by Vietnam, Brazil, and Russia at 18% each. Meanwhile, the production of REEs in 2020 was around 240,000 t REO equivalent, which was dominated by China (58%) and the USA (16%) [2].
The mining methods applied for REMs are dependent on the type of deposit; that is, hard rock mining techniques are used for carbonatite, peralkaline-igneous, magmatic magnetite-hematite and IOCG deposits, while, for clay and placer deposits, in-situ leaching and placer mining are used, respectively [8]. Beneficiation processes—gravity separation, magnetic/electrical separation, and flotation—are usually applied to separate REMs from their associated gangue minerals and concentrate them in the final product. At larger particle sizes, gravity and magnetic separation techniques have been reported as effective ways of concentrating REMs. Park et al. [9], for example, used multi-stage wet high-intensity magnetic (WHIM) separation to upgrade a low-grade peralkaline igneous REE deposit from Mongolia with a particle size distribution between 75 µm and 106 µm and reported a recovery of 80% and enrichment ratio of 5.5.
For finer particle size distribution (<75 µm) and high purity requirements, however, more advanced techniques like flotation—a method that separates target materials either by promoting or suppressing their attachment to bubbles [10,11,12]—are required to concentrate REMs and REE-bearing materials [1,13,14,15,16]. Because of advancements in flotation technologies in recent years, like agglomeration-flotation [17,18], carrier flotation [19], and nano-bubble flotation [20], the size range limit where flotation operates has been lowered from 50–75 µm to 5–20 µm. For example, Hornn and coworkers [21,22,23] highlighted the improved flotation efficiency of ultrafine chalcopyrite (<10 µm) after agglomeration using kerosene as a bridging agent. Once sufficiently concentrated, REEs in REMs are typically extracted via hydrometallurgy, which involves multiple iterations of leaching, solvent extraction, ion exchange, and precipitation [14,24].
Similar to other critical elements like lithium, nickel, and cobalt, however, REEs are non-renewable resources that would eventually be depleted if their sustainable utilization and management were not considered. By assuming that the current REE reserve of primary resources and production remain constant, while consumption increases at an annual rate of 5% [25], the global REE reserve is forecasted to be depleted within 127 years (i.e., by 2148). This means that the processing of secondary resources (e.g., wastes), especially end-of-life products and electronic wastes (E-wastes), via recycling—also called urban mining—is a potential way of reducing supply risks and prolonging the limited REE reserves. Binnemans et al. [26], for example, estimated that there are 5600–10,700 t of recoverable REEs from end-of-life products like magnets, nickel metal hydride batteries, and phosphors.
This systematic review was carried out to summarize significant findings about REE processing and recycling from 2012 to 2021. Specifically, research works on REE recovery by flotation from primary and secondary resources were reviewed using the PRISMA guidelines. The contents of this work are divided into five sections: (i) introduction, (ii) review methodology, (iii) rare earth flotation from primary resources, (iv) rare earth flotation from secondary resources, and (v) conclusions.

2. Review Methodology

The literature was systematically reviewed using the research question “What is the state of conventional technologies and research studies related to REE recovery from primary and secondary resources using flotation for the last 10 years (2012–2021)?” and the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines [27]. The guideline recommended by Andrews [28] for peer-reviewed journal identification was followed, and the keywords “flotation”, “rare earth”, and the names of 17 REEs [3] and 84 REMs [29] were used, as summarized in Table S1 [28].
Web of Science and Scopus were selected as the databases for this systematic review, and the publication dates were limited to between 2012 and 2021. Based on these initial criteria, a total of 1251 articles were obtained after removing duplicates (Figure S1). In the screening step, titles, highlights, abstracts, and keywords were checked to remove studies not focusing on “rare earth”. This step removed 920 papers, and the remaining 331 papers were further evaluated in the eligibility step. In this step, full-text articles were checked, and, based on the systematic selection criteria [30], 127 papers qualified for the review. The selected papers were categorized into two main sections: (i) “rare earth flotation from primary resources” (101 papers), and (ii) “rare earth flotation from secondary resources” (26 papers).

3. Rare Earth Flotation from Primary Resources

The separation of REMs via flotation involves the following: (i) the addition of a collector and pH modifiers to make the REMs more hydrophobic and suitable for bubble attachment, (ii) the stabilization of bubbles using frothers, and (iii) the depression of gangue minerals by inhibiting unnecessary collector adsorption (Figure 1).

3.1. Flotation Reagents

3.1.1. Collectors

Between 2012 and 2021, three collector types—carboxylates, hydroxamates, and phosphates/phosphonates—were frequently reported. The typical chemical structures of these three classes of collectors are illustrated in Figure S2.

Carboxylate Collectors

Carboxylates, also known as oxhydryl collectors, are conventional collectors in REM flotation because they are cheaper than the other collector types [31]. The mechanism of carboxylate attachment on REMs can be classified as physisorption and/or chemisorption. Espiritu et al. [32] proposed that sodium oleate adsorbs on REMs via physisorption when the pulp pH is lower than the point of zero charge (PZC). In contrast, chemisorption was observed to dominate when the pulp pH was higher than the PZC and was attributed to interactions between the negatively charged surface of the REMs and anionic oleate ions [32,33].
The main disadvantage of carboxylates noted from previous works is their generally low selectivity for REMs because they also tend to adsorb to gangue minerals like calcite, fluorite, dolomite, barite, and quartz [34,35,36,37,38]. To enhance the REM grade when carboxylate is used, depressants are needed to inhibit collector adsorption to gangue minerals [39] (more details will be provided in the later subsection on depressants).
Sodium oleate is the most frequently used carboxylate collector in REM flotation, especially for bastnaesite, monazite, and xenotime (Table 1). The best pH for this collector is around 6–10, where REO recovery of 95% could be obtained when single REMs were evaluated [32,33,40,41]. Unfortunately, this high recovery decreased substantially in the presence of gangue minerals, as highlighted by Espiritu and Waters [34], who obtained only 32% recovery with monazite–dolomite mixtures.
To improve recovery in real REMs, depressants—xanthan gum, lignin sulfonate, sodium silicate, and sodium silicate mixed with starch—are used. Flotation of REMs from low-grade sources using sodium oleate as a collector and sodium silicate as a depressant has reported 40–80% recoveries and obtained 0.3–0.5% grade of products [36,42,43,44,45].
For REM flotation using fatty acids as collectors, the optimum pH is around 7–11.5. The effects of temperature on REM flotation with fatty acids were evaluated by Satur et al. [46], and they reported negligible improvement in flotation recovery even after increasing the temperature to 40 °C, due to the low selectivity of this collector at a high temperature. It was also found that, among the fatty acids, Sylfat FA2 was better than Aero 704 at pH 7 to 11.5. Meanwhile, oleic acid, an unsaturated fatty acid, was effectively applied as a collector in REMs with associated silicates and iron-bearing minerals [47].
Table 1. A summary of recent studies on REM flotation using carboxylate collectors.
Table 1. A summary of recent studies on REM flotation using carboxylate collectors.
CollectorsSamplespHDepressantsRecovery [%]Grade [%]References
Sodium oleateBastnaesite6–895[32,33]
995[40,41]
Bastnaesite and calcite8Xanthan gum72.5[48]
Bastnaesite, calcite, and barite9–10Lignin sulfonate86–9011–33.4[39]
Bastnaesite, monazite, stetindite, brannerite, florencite, stillwellite, hematite, and quartz9Sodium silicate and starch632.3 [49]
Monazite 795[32,33]
Monazite and dolomite732[34]
Monazite, hematite, and quartz9Sodium silicate and starch616.3[50,51]
Xenotime, zircon, schorl, staurolite, and ilmenite9Sodium silicate71–9636.3–51.1[52]
Ammonium lignosulfonate33–5252.3–59.6[52]
Allanite and quartz[53]
Fergusonite, quartz, and galena5–7[54]
Heavy mineral sands; containing monazite, ilmenite, rutile, zircon, leucoxene, garnet, and pyroxene10Cyquest400020–3243–66[55]
Pre-concentrates of iron oxide–silicate-rich tailings 9Sodium silicate and starch121.9[56]
Kaolin by-products containing monazite10Sodium silicate80.20.54[42]
Fine coal refuse 90.47 [44]
Fire clay seam coal source, calcite, dolomite, illite, kaolinite, pyrite, quartz, and rutile0.05–0.1[45]
Tailings from Kankberg mine8Sodium silicate40–550.28–0.3[43]
Sodium oleate; mixed with sorbitan monooleateFergusonite, zircon, quartz, and feldspar5Sodium metasilicate, oxalic acid, and citric acid96[57]
Fatty acidPhosphate flotation tailing containing REMs640.04[58]
Fatty acid (i.e., oleic acid)Monazite, hematite, and quartz7Sodium silicate66–9632.9–62.8[59]
Starch72–8232.9–37.7[59]
Fatty acid (i.e., sylfat FA2)Xenotime monazite, silicate minerals, and hematite10–12Sodium metasilicate731.7[46]
Sodium lignin sulfonate812.0[46]
Sodium fluoride911.6[46]
Soluble starch842.1[46]
Fatty acid; mixed with fuel oilPhosphatic clay9Sodium silicate880.4[60]
Alaskan coal samples7[61]
Fatty acid (i.e., Aero704); mixed with sulfosuccinamate (i.e., Aero845)Bastnaesite, parisite, synchysite-Y, gittinsite, britholite, elpidite, zircon, pyrochlor, K-feldspar, albite, quartz, arfevdsonite, echemannite, calcite, fluorite, rutile, hematite, and chlorite9Sodium silicate and starch52–541.1–1.2[62]

Hydroxamate Collectors

Hydroxamates are a group of chelating collectors popular in REM flotation because of their high selectivity [34,36,63]. The adsorption of hydroxamate-type collectors on mineral surfaces can occur via chemisorption and/or surface reactions (Figure S3). In chemisorption, hydroxamate interacts with the mineral surfaces without the movement of surface atoms [31,32,64,65,66,67]. This type of adsorption was proposed when hydroxamate concentration was low and monolayer adsorption predominated. Meanwhile, surface reactions—combinations of chemisorption and multilayer physisorption—are dominant at high concentrations of hydroxamate and involve (i) hydrolysis of metal atoms, (ii) formation of hydroxyl-complexes, and (iii) re-adsorption to or precipitation on the mineral surface [34,65,68].
There are many types of hydroxamate collectors used for REM flotation (Table 2), and the most common and frequently used are octanohydroxamic acid, benzohydroxamic acid, salicylhydroxamic acid, H205, Aero 6493, and AEROFLOAT 6494.
Wanhala et al. [65] studied the adsorption mechanism of octanohydroxamic acid on bastnaesite and proposed that, at low concentrations, monodentate adsorption was dominant while, at higher collector concentrations, the mechanism shifts to bidentate adsorption. These authors also observed that the binding enthalpies of octanohydroxamic acid interacting with La- and Ce-bastnaesite surfaces were stronger than on calcite. Moreover, the selectivity of hydroxamate was enhanced by maintaining a low collector concentration that limited hydroxamate adsorption to gangue minerals. The optimum pH for pure bastnaesite flotation using hydroxamate collectors is around pH 8–9, with 67–75% recovery [31,32,69,70,71]. Similarly, the optimum pH for pure monazite flotation using hydroxamate collectors is around 7–9, with 61–95% recovery [32,70,72]. It should be noted that when common gangue minerals are present, a lower grade and recovery were obtained.
Another notable feature of monazite flotation using hydroxamate was described by Zhang and Honaker [72,73]. These authors observed the high affinity of monazite to float at pH 9, likely due to hydrolyzed species (e.g., Ce(OH)2+, Ce(OH)2+, La(OH)2+, and La(OH)2+, which increased the active sites on monazite. However, the flotation response of monazite was very similar to its associated gangue minerals like apatite, ilmenite, quartz, rutile, and zircon when octanohydroxamic acid was used [74].
Table 2. A summary of recent studies on REM flotation using hydroxamate collectors.
Table 2. A summary of recent studies on REM flotation using hydroxamate collectors.
CollectorsSamplespHDepressantsRecovery [%]Grade [%]References
Octanohydroxamic acidBastnaesite8–967[71]
Bastnaesite and calcite99036.4[75]
Bastnaesite and fluorite9EDTA80–8868[76]
Bastnaesite and parisite10Sodium silicate65[77]
Bastnaesite, calcite, and barite9–10Lignin sulfonate28–8523–33[39]
Monazite995[72]
780[78]
Monazite and calcite990[79]
Sodium silicate80[80]
Sodium hexametaphosphate62[80]
Monazite, apatite, ilmenite, quartz, rutile, and zircon8–10> 90[74]
Xenotime, zircon, schorl, staurolite, and ilmenite9Sodium silicate86–9335.1–49.8[52]
Ammonium lignosulfonate79–8834.8–50.1[52]
Ancylite, strontianite, and calcite885[81]
976–7910.7–10.9 [82]
Benzohydroxamic acidBastnaesite968–75[32,70]
Bastnaesite and calcite95722.5[83]
Bastnaesite and quartz9Sodium silicate68[84]
Monazite961[70]
868[32]
Monazite and dolomite737[34]
Allanite and quartz420[53]
Bastnaesite, synchysite, allanite, monazite, fergusonite, feldspar, quartz, and others925–30[85]
Niobec REE ore9Sodium silicate61–8025–34[86]
Salicylhydroxamic acidBastnaesite872[69]
Bastnaesite and calcite96733.4[83]
Bastnaesite, barite, and calcite7–9> 80[87]
Bastnaesite, feldspar, calcite, quartz, fluorite, and biotiteSodium silicate67–738.3–8.8[88]
Tailings from Bayan Obo mine8Sodium silicate[89]
3-hydroxy-2-naphthyl hydroxamic acid (H205)Bastnaesite, calcite, and fluorite8–9Oxalic acid80–8242.7–43.1[90]
9Sodium silicate70–8216–20[91]
Bastnaesite, fluorite, apatite, hematite, and quartz7–10> 80[92]
Bastnaesite and (sulfide, silicate, and carbonate) mineralsSodium silicate55–6313.0–14.5[93]
Aero-6493Bastnaesite and monazite7–960–90[94]
Bastnaesite, monazite, calcite, dolomite, and quartz962–6660.7[95]
AEROFLOAT 6494Monazite791–95[59,96]
Monazite, quartz, and hematite7Sodium silicate and starch931.0[96]
Bastnaesite, monazite, stetindite, brannerite, florencite, stillwellite hematite, and quartz7Sodium silicate and starch602.0[49]
Bastnaesite, parisite, synchysite-Y, gittinsite, britholite, elpidite, zircon, pyrochlor, K-feldspar, albite, quartz, arfevdsonite, echemannite, calcite, fluorite, rutile, hematite, and chlorite9Sodium silicate and starch51–530.3[62]
Decanedioic hydroxamic acid (DDHA)Bastnaesite and calcite88686.2[97]
N-[3-(dodecylamino)] hydroxamic acid (DAHA)Bastnaesite and calcite899[98]
N-[3-(octylamino)] hydroxamic acid (OAHA)Bastnaesite and calcite895[98]
N-[3-(hexylamino)] hydroxamic acid (HAHA)Bastnaesite and calcite893[98]
N-[(3-hydroxyamino)-propoxy]-N-octyl dithiocarbamateBastnaesite, calcite, and quartz897[99]
Octylmalondihydroxamic acidBastnaesite and calcite99069.3[75]
Octylamino-bis-(isobutanohydroxamic acid) (OIBHA)Bastnaesite and fluorite7Sodium fluorosilicate9270.1[100]
Octylamino-bis-(butanohydroxamic acid) (OBHA)Bastnaesite, barite, and calcite78968.1[101]
3-carboxy-2-naphthylhydroxamic acid; mixed with C5-C9 hydroxamic acidBastnaesite, monazite, hematite, magnetite, iron silicate, and pyrite9Sodium silicate and “sodium carbonate; mixed with fine mud”5051.9[102]
1-hydro-2-naphthyl hydroxamic acid (LF8#)Bastnaesite, monazite, fluorite,
barite, and silicate minerals
9Sodium silicate7545–50[103]
Benzylhydroxamic; mixed with naphthalenic hydroxamate (LF-P81)Thor lake deposit9Sodium silicate11.4[104]
Hydroxamate (AC-3)Kaolin by-products containing monazite10Sodium silicate880.3[42]
Hydroxamic acidPreconcentrates of iron oxide–silicate-rich tailings 9Sodium silicate and starch91–931.3[56]
Bastnaesite, monazite, niobite, and iron-bearing minerals 874[105]

Organo-Phosphoric Acid Collectors

Table 3 summarizes the studies applying organo-phosphoric acids as collectors. The adsorption mechanism of this collector type is similar to carboxylates and hydroxamates, and their optimum pH is around 4–9. The (ROPO3H-) and (ROPO32−) are the dominant species of phosphate collectors [32,106], which interact with REEs (e.g., Ce, La) on REMs via monolayer chemisorption. Surface precipitation of REE phosphate complexes was also reported at high collector concentrations, along with the presence of Ce3+, CeF2+, CeF2+, and CeCO3+ [106,107,108,109]. Lauryl phosphate and Flotino 1682 are the two most frequently studied organo-phosphoric acid collectors.

Others

A combination of kerosene and fatty acid as alternative collectors was investigated by Fang et al. [116] and Zhou et al. [117], which obtained 60% recovery in pure bastnaesite flotation at pH 6 (Table S2). Xu et al. [118,119] used octanohydroxamic acid and sodium oleate and reported 94–97% recovery in bastnaesite flotation with fluorite and calcite.
Ionic liquids as novel collectors were investigated by Azizi et al. [95] and Li et al. [120], which obtained 65–90% recovery. Other collectors like sodium dodecyl sulphate, Flotigam EDA, R845N, and fuel oil have also been investigated [50,121,122], and these works reported recoveries of 75–98%.

3.1.2. Frothers

In flotation, frothers are used to stabilize bubbles to make them strong enough to carry hydrophobic particles to the surface of the flotation pulp. Compared with other flotation reagents, studies on the effects of frothers on REM flotation remain limited. The most widely used frother in REM flotation studies is methyl isobutyl carbinol (MIBC), while other frothers like F150, EM-312, and Aerofroth-88 are utilized in industrial-scale processes [34,93,122].

3.1.3. Depressants

Depressants are needed because common gangue minerals found in REM deposits like carbonates, silicates, and iron-bearing minerals have similar adsorption affinities to collectors as REM during flotation [62,96,123,124].
Zhang and Anderson [52] and Boulanger et al. [86] successfully depressed silicate minerals (e.g., zircon, schorl, staurolite) and carbonates (at 75 °C) using sodium silicate. The depression mechanism proposed by these authors was the physical adsorption of negatively charged silicate to positively charged alkaline hydroxylated complexes, which led to hydrophilic film formation on silicate and carbonate minerals. In addition, combining sodium silicate and starch was suggested by Abaka-Wood et al. [49,50,51,56,96], which lowered sodium silicate consumption without affecting efficiency. Oxalic acid could also depress quartz, dolomite, and calcite at around pH 3.8–7.2 due to the formation of Mg/Ca oxalate precipitates on gangue mineral surfaces that inhibit collector adsorption [123].
Other depressants have also been investigated, but their mechanisms of depression were not reported. Ammonium lignosulfonate could depress iron-bearing minerals (ilmenite) [52], while sodium fluorosilicate [100] and/or EDTA [76] suppressed fluorite floatability. Meanwhile, sodium hexametaphosphate [80] and/or xanthan gum [48] were found to depress calcite and sodium metasilicate, preventing the flotation of quartz and feldspar [57]. Finally, Satur et al. [46] reported the depression of iron-bearing minerals by sodium lignin sulfonate.

3.2. Roles of Dissolved Ions in Rare Earth Mineral Flotation

Coexisting dissolved ions and their complexes are also important factors in REM flotation. They can be categorized based on their origin as (i) inherently from REMs and associated minerals, and (ii) artificially from flotation reagents.
REM deposits contain slightly soluble common associated gangue minerals (e.g., calcite, fluorite, barite, and dolomite) and highly soluble salt gangue minerals (e.g., strontianite, celestite) [40,41,66,70,107]. When in contact with water, these gangue minerals partly dissolve and release ionic species (e.g., Ca2+, Ba2+, Mg2+, Sr2+, CO32−, SO42−) into the flotation pulp [3,33,38,69,70,125]. These ions in solution could form various complexes (e.g., CaOH+, BaOH+, MgOH+, SrOH+, Ca(OH)2, Ba(OH)2, Mg(OH)2, CaCO3, BaCO3, MgCO3, and BaSO4) that adsorb and/or precipitate on the surface of REMs [40,41,69,79]. Their adsorption and/or precipitation generally inhibit collector adsorption on REMs, which complicates the separation of REMs from associated gangue minerals (Figure 2a).
Furthermore, Ca2+, Ba2+, and Mg2+ may react with anionic collectors, especially with carboxylate, which lower collector concentration in the flotation pulp and reduce REM floatability. One workaround to suppress this reaction is the use of citric acid and EDTA, both of which limit the availability of interfering cations to react with carboxylate by forming calcium–citric acid complex (CaCit) and calcium–EDTA complex (CaEdta2−) [80].
Other coexisting cations like La3+, Pb2+, Al3+, and Fe3+ may improve the flotation efficiency of REMs, as noted by many researchers [43,71,78,85,104,126,127,128,129]. These cations are artificially added into the flotation pulp to make the surface of REMs more positively charged (e.g., PbOH+ adsorption) and improve collector adsorption. This technique works especially well for light rare earth elements (LREEs) (Figure 2b).

4. Rare Earth Flotation from Secondary Resources

REEs in secondary resources are generally found in two forms: solid and liquid. Conventional froth flotation is a promising approach for solid REE-bearing wastes (e.g., phosphors), but more advanced flotation techniques are required for liquid wastes (e.g., industrial wastewater). There are four unique flotation technologies—ion flotation (Figure 3a), solvent sublation (Figure 3b), electroflotation (Figure 3c), and absorbing colloid flotation (Figure 3d)—recently developed for recovering REEs from industrial wastewater even at low concentrations.

4.1. Froth Flotation

As described earlier, froth flotation (Figure 1) is a classical method to separate solid particles based on their different surface wettability. To date, the only solid waste amenable to flotation is phosphor, which comes from waste fluorescent lamps. Phosphors in fluorescent lamps can be classified into three groups: (i) red phosphors ((Y0.95Eu0.05)2O3), (ii) green phosphors ((Ce0.67Tb0.33)MgAl11O19), and (iii) blue phosphors ((Ba0.9Eu0.1)Mg2Al16O27) [130].
Yu et al. [130] carried out the successful separation of phosphors and glass from waste fluorescence lamps using reverse flotation. In contrast to conventional direct flotation, reverse flotation removes impurities in the froth product while target REEs-bearing materials are recovered in the tailings. Yu et al. [130] used dodecyl ammonium acetate (DDA) as a collector—usually called “surfactant” in resources recycling—to physically adsorb on glass particles while alkaline starch was utilized as a depressant for waste phosphors. The results showed that glass particles had high floatability and were recovered in the froth products while phosphors were mostly retained in the pulp. Among the three phosphors, red phosphors had the lowest floatability, followed by blue phosphors and green phosphors. The highest purity of TREO was obtained at pH 9, and the purity and recovery of model samples were 27.0% and 71.4%, respectively, while, in real wastes, the purity was slightly lower at 26.0% but with a slightly higher recovery of 73.3%.

4.2. Ion Flotation

For the REEs in the form of ions in liquid wastes like industrial wastewater, advanced flotation can be applied. Ion flotation (Figure 3a)—sometimes called “foam flotation” and “ion foam flotation (IFF)”—is the most widely used flotation method for REE recycling from wastewater in the last ten years. Similar to direct flotation, surfactants are used; however, adsorption in ion flotation occurs on the REE ions rather than the particle surface. The liquid–gas adsorption between the “sublate”, REE ions with adsorbed surfactants, and bubbles occurs. The bubbles then bring the sublate up to the water surface, foam is generated, and REE ions are recovered as “foamate products” [131].
Compared with the purification methods in hydrometallurgy like chemical precipitation, ion exchange, and solvent extraction, ion flotation is more environmentally friendly because of the low toxicity and amounts of chemical reagents used [132,133]. Also, other advantages of this technique include simplicity, low operational costs, rapid reactions, flexibility, low energy consumption, and applicability for substances of different physicochemical properties [133,134,135].
Justeau et al. [132] found that Ce ions could be recovered using ion flotation without any surfactants but using F-Hexane gas to generate microbubbles. However, the majority of ion flotation studies used sodium dodecyl sulfate (NaDDS) as the surfactant (Table 4). Previous studies have shown that this technique could recover low-concentration REEs (10–6 to 10–3 M) with efficiencies of 57–98% [133,134,135,136,137,138,139,140]. Aside from NaDDS, other surfactants like monorhamnolipid [131,140], AKYPO RO 90 VG [141], and cetylpyridinium chloride (CPC) [142] have also been used. Although frothers are typically not used in ion flotation, several authors have reported improved efficiency when using frothers like ethyl alcohol (EtOH) [131,142].

4.3. Solvent Sublation

Solvent sublation (Figure 3b) is an advanced flotation technique combining the principles of ion flotation and solvent extraction. This technique consists of two liquid components: (i) aqueous phases containing REE ions, and (ii) an organic phase. It should be noted that the organic substances used in this process must have a lower density than water to facilitate separation. They also need to have low solubility and a high ability to accumulate sublate, so isooctyl alcohol and 2-octanol are usually used [138,139,144,145].
Separation occurs when sublate-bearing gas bubbles (usually using nitrogen gas) float and transfer rare earth ions to the organic phase. The gas flow rate used in this process must be very low (~3.3 × 10−4 m3/h) to ensure that the bubbles do not disturb the organic phase. Similar to ion flotation, NaDDS is frequently used as the surfactant in solvent sublation. Several authors have obtained 99% separation efficiency using this technique at pH 7–8 [138,139,144,145].

4.4. Electroflotation

Electroflotation (Figure 3c) is an advanced flotation technique that uses electrolysis to generate microbubbles. Stainless steel and aluminum are the most common and inexpensive cathode materials used, while titanium with ruthenium-titanium oxide coating is preferred as the anode [146,147,148].
In the past ten years, electroflotation has often been applied to higher concentration REEs (50–250 M) and suitable for the leachate of the hydrometallurgical processes in chloride, carbonate, sulfate, and nitrate media [147]. It is also important to note that in electroflotation, the formation of poorly soluble compounds of metal salts (e.g., metal-hydroxides, metal-sulfides, metal-phosphates) is essential before floating REE ions to the surface via bubbles [149].
Surfactants used in electroflotation include cationic (Septapav and DDAC), anionic (NaDDS), and non-ionic (PEO-1500) types (Table 5). It has been reported, however, that electroflotation efficiency depends on the type of electrolyte, which could be related to the effects of REE ionic charges. In a chloride electrolyte, anionic surfactant (NaDDS) is outperformed by other surfactants, especially for the recovery of Sc at neutral pH. Meanwhile, cationic surfactant is better when REE ions are found in carbonate and nitrate electrolytes. In some cases, flocculants (i.e., C494, C496, N300, M345, and A137) were used instead of surfactants to enlarge the precipitated metal compounds and enhance their probability of collision with bubbles [146,150,151].

4.5. Absorbing Colloid Flotation

Another advanced flotation technique capable of extracting low quantities of REE ions from an aqueous solution is absorbing colloid flotation (ACF; Figure 3d). This method utilizes colloidal Fe(OH)3 by adding ferric chloride (FeCl3) into the flotation system. Separation occurs when REE ions adsorb to colloids via physisorption called “hydroxide flocs”, which have a positive charge. An anionic surfactant (NaDDS) is then added to collect the hydroxide flocs through their attachment to rising bubbles. Rezgar et al. [154], for example, investigated the application of ACF to recover Ce from an aqueous solution and found that maximum recovery reached 99.8% under acidic conditions.
Based on the concept of ACF, Perlova and Chernetskaya [155] investigated the application of phosphine oxide as a carrier to recover La and Ce from an aqueous solution. In this previous work, the recovery of REE ions reached 90% under alkaline pH because of higher La3+- and Ce3+-hydroxocomplexes interacting with polar groups on phosphine oxide.

5. Conclusions

Flotation technology is widely employed in mineral processing, particularly for the concentration of REMs, because of its high efficiency in separating particles based on their contrasting surface properties (hydrophilic and hydrophobic). The findings of this systematic review indicate that flotation technology as applied to REEs can be categorized into two distinct groups: (i) flotation technology for primary resources, and (ii) flotation technology applied to secondary resources. For the first group, flotation proves to be advantageous in the separation of REMs from associated gangue minerals, especially at finer liberation sizes. This technique also offers higher selectivity for REMs recovery than other separation methods and ensures that the products have higher grades. In the second group, conventional and advanced flotation techniques were employed to recover REE-bearing materials from secondary sources, such as wastewater generated by hydrometallurgical processes and other industries, including end-of-life products like phosphors in fluorescent lamps. The application of flotation technology in these secondary resources allows for the effective extraction and recovery of REEs, contributing to not only resource recycling but also environmental conservation.
Recent research works highlight that in the flotation of primary resources, significant attention was given to the selection of flotation reagents, including collectors, frothers, and depressants. Among the various collectors used, three types are frequently employed: (i) carboxylate, (ii) hydroxamate, and (iii) organo-phosphoric acid. It was observed that, during flotation, hydroxamate and organo-phosphoric acid collectors exhibited higher affinities for REMs than carboxylate collectors. Hydroxamate, however, has the highest price among the three common collectors for REMs. Meanwhile, organo-phosphoric acid collector was the most sensitive in terms of flotation conditions and was reported to be effective only under acidic conditions. When economics is considered for the selection of a collector for REMs flotation, carboxylate is the best candidate, due to its lower cost compared with the other collector groups.
Over the past decade, numerous studies have investigated the performance-enhancing effects of hydroxamate collectors. Additionally, alternative collectors, such as ionic liquids, sodium dodecyl sulphate, flotigam EDA, R845N, combinations of kerosene and fatty acids, and combinations of octanohydroxamic acid and sodium oleate, have been introduced to further improve flotation efficiency. The literature consistently demonstrates that the most effective approach to enhance flotation recovery and grade is through the depression of associated gangue minerals. Several depressants have been employed in REMs flotation. Notably, the combination of sodium silicate and starch was beneficial for depressing all types of gangue minerals, particularly silicate gangue minerals. Meanwhile, ammonium lignosulfonate was suitable for inhibiting iron-bearing minerals. Dissolved ions were also highlighted as having significant effects on REMs flotation. Extensive research works have elucidated the consumption of collectors when semi-soluble gangue minerals like calcite, fluorite, dolomite, and barite were present in the flotation pulp. The inevitable presence of alkaline earth metal ions (e.g., Ca2+, Mg2+, Ba2+, Sr2+) inhibited the adsorption of collectors on REM surfaces, resulting in lower floatability. In contrast, certain metal cations such as Pb2+, La3+, Fe3+, and Al3+ have been applied to enhance the floatability of REMs. These cations exhibited a strong affinity to activate REM surfaces by providing more active sites for collector adsorption.
For secondary resources, flotation technologies for REE recovery can be classified into five main techniques: (i) froth flotation, (ii) ion flotation, (iii) solvent sublation, (iv) electroflotation, and (v) absorbing colloid flotation. It is worth noting that froth flotation, identical to conventional flotation of primary resources, remains a highly effective method for isolating end-of-life products containing REEs, particularly for separating phosphors from glass in spent fluorescent lamps. Meanwhile, the other flotation techniques were primarily applied to isolate REEs from wastewater, showcasing their effectiveness in separating bulk REEs. These flotation technologies not only contributed to wastewater treatment but also facilitated the recovery of REEs, aligning with the principles of sustainable development and supporting the UN Sustainable Development Goals (Goal 12: Responsible consumption and production). However, it is important to highlight that the separation of individual REE ions through flotation techniques still lacks extensive study. This represents a challenging topic for future research, focusing on enhancing flotation technologies by investigating the effects of important parameters such as surfactant selection, solution pH, and electrolyte conditions. By addressing this research gap, advancements can be made in optimizing flotation processes for individual REE separation.
We believe that this review is beneficial for stakeholders across various disciplines, including those working in fundamental research studies, policy development, mineral processing, and extractive metallurgy, as well as waste management and recycling. It offers a fundamental understanding of conventional and advanced flotation, presents key concepts, and provides numerous ideas for developing future technologies in the REMs industry. By utilizing the concepts and insights presented in this systematic review, stakeholders can drive advancements and innovations in REMs processing and REE recycling, ultimately contributing to the growth and sustainability of the industry.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app13148364/s1, Figure S1: Study selection flow diagram.; Figure S2: Chemical structure of typical three collector types mostly used in REMs flotation (a) carboxylate, (b) hydroxamate, and (c) organo-phosphoric acid.; Figure S3: (a) Adsorption mechanism of collector on REM surface; (a-1) chemisorption and (a-2) surface reaction, and (b) chemisorption modes; (b-1) bridge, (b-2) mondentate, and (b-3) bidentate.; Table S1: Lists of 17 REEs and 84 REMs used as searching keywords.; Table S2: Summary of recent studies on REM flotation using other collectors [3,29,31,32,33,50,63,64,65,66,95,101,106,116,117,118,119,120,121,122].

Author Contributions

Conceptualization, P.J., A.N., L.T., O.J. and T.P.; methodology, P.J., O.J., P.S., C.B.T. and T.P.; software, P.J. and P.S.; validation, P.J., C.B.T. and T.P.; formal analysis, P.J. and P.S.; data curation, P.J. and P.S.; writing—original draft preparation, P.J., A.N., L.T., O.J., P.S., C.B.T. and T.P.; writing—review and editing, P.J., K.A., I.P., M.I., C.B.T. and T.P.; visualization, P.J. and T.P.; supervision, A.N., M.I., C.B.T. and T.P.; project administration, P.J. and T.P.; funding acquisition, T.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by Chulalongkorn University (ReinUni_65_02_21_06).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to thank the members of the Faculty of Engineering, Chulalongkorn University, for their support to this project and gratefully acknowledge the editors and anonymous reviewers for their valuable inputs to this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Balaram, V. Rare earth elements: A review of applications, occurrence, exploration, analysis, recycling, and environmental impact. Geosci. Front. 2019, 10, 1285–1303. [Google Scholar] [CrossRef]
  2. U.S. Geological Survey. Mineral Commodity Summaries 2021; U.S. Geological Survey: Reston, VA, USA, 2021; 200p. [CrossRef]
  3. Xu, L.; Wang, Z.; Shu, K.; Wu, H.; Hu, Y. Surface chemistry considerations of gangue dissolved species in the bastnaesite flotation system. Fundam. Res. 2021, 2, 748–756. [Google Scholar] [CrossRef]
  4. Dostal, J. Rare earth element deposits of alkaline igneous rocks. Resources 2017, 6, 34. [Google Scholar] [CrossRef] [Green Version]
  5. Van Gosen, B.S.; Verplanck, P.L.; Seal, R.R., II; Long, K.R.; Gambogi, J. Rare-Earth Elements. In Critical Mineral Resources of the United States—Economic and Environmental Geology and Prospects for Future Supply; Klaus, J.H., Schulz, J., Eds.; U.S. Geological Survey: Reston, VA, USA, 2017. [Google Scholar] [CrossRef]
  6. Verplanck, P.L.; Van Gosen, B.S.; Seal, R.R.; McCafferty, A.E. A Deposit Model for Carbonatite and Peralkaline Intrusion-Related Rare Earth Element Deposits; U.S. Geological Survey Scientific Investigations Report 2010–5070-J; U.S. Geological Survey: Reston, VA, USA, 2014. [CrossRef] [Green Version]
  7. Tabelin, C.B.; Park, I.; Phengsaart, T.; Jeon, S.; Villacorte-Tabelin, M.; Alonzo, D.; Yoo, K.; Ito, M.; Hiroyoshi, N. Copper and critical metals production from porphyry ores and E-wastes: A review of resources availability, processing/recycling challenges, socio-environmental aspects, and sustainability issues. Resour. Conserv. Recy. 2021, 170, 105610. [Google Scholar] [CrossRef]
  8. Cincinnati, O.H. Rare Earth Elements: A Review of Production, Processing, Recycling, and Associated Environmental Issues; EPA 600/R-12/572; United States Environmental Protection Agency: Washington, DC, USA, 2012.
  9. Park, I.; Kanazawa, Y.; Sato, N.; Galtchandmani, P.; Jha, M.K.; Tabelin, C.B.; Jeon, S.; Ito, M.; Hiroyoshi, N. Beneficiation of Low-Grade Rare Earth Ore from Khalzan Buregtei Deposit (Mongolia) by Magnetic Separation. Minerals 2021, 11, 1432. [Google Scholar] [CrossRef]
  10. Aikawa, K.; Ito, M.; Segawa, T.; Jeon, S.; Park, I.; Tabelin, C.B.; Hiroyoshi, N. Depression of lead-activated sphalerite by pyrite via galvanic interactions: Implications to the selective flotation of complex sulfide ores. Miner. Eng. 2020, 152, 106367. [Google Scholar] [CrossRef]
  11. Jeon, S.; Ito, M.; Tabelin, C.B.; Pongsumrankul, R.; Kitajima, N.; Hiroyoshi, N. Gold recovery from shredder light fraction of E-waste recycling plant by flotation-ammonium thiosulfate leaching. Waste Manag. 2018, 77, 195–202. [Google Scholar] [CrossRef]
  12. Phengsaart, T.; Srichonphaisan, P.; Kertbundit, C.; Soonthornwiphat, N.; Sinthugoot, S.; Phumkokrux, N.; Juntarasakul, O.; Maneeintr, K.; Numprasanthai, A.; Park, I.; et al. Conventional and recent advances in gravity separation technologies for coal cleaning: A systematic and critical review. Heliyon 2023, 9, e13083. [Google Scholar] [CrossRef]
  13. Abaka-Wood, G.B.; Addai-Mensah, J.; Skinner, W. Magnetic Separation of Monazite from Mixed Minerals; Chemeca: Adelaide, Australia, 2016. [Google Scholar]
  14. Jordens, A.; Cheng, Y.P.; Waters, K.E. A review of the beneficiation of rare earth element bearing minerals. Miner. Eng. 2013, 41, 97–114. [Google Scholar] [CrossRef]
  15. Kim, K.; Jeong, S. Separation of Monazite from Placer Deposit by Magnetic Separation. Minerals 2019, 9, 149. [Google Scholar] [CrossRef] [Green Version]
  16. Khodakarami, M.; Alagha, L.; Burnett, D.J. Probing surface characteristics of rare earth minerals using contact angle measurements, atomic force microscopy, and inverse gas chromatography. ACS Omega 2019, 4, 13319–13329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Hornn, V.; Ito, M.; Shimada, H.; Tabelin, C.B.; Jeon, S.; Park, I.; Hiroyoshi, N. Agglomeration-Flotation of Finely Ground Chalcopyrite and Quartz: Effects of Agitation Strength during Agglomeration Using Emulsified Oil on Chalcopyrite. Minerals 2020, 10, 380. [Google Scholar] [CrossRef] [Green Version]
  18. Julapong, P.; Ekasin, J.; Katethol, P.; Srichonphaisarn, P.; Juntarasakul, O.; Numprasanthai, A.; Tabelin, C.B.; Phengsaart, T. Agglomeration–flotation of microplastics using kerosene as bridging liquid for particle size enlargement. Sustainability 2022, 14, 15584. [Google Scholar] [CrossRef]
  19. Bilal, M.; Ito, M.; Akishino, R.; Bu, X.; Hassan, F.U.; Park, I.; Jeon, S.; Aikawa, K.; Hiroyoshi, N. Heterogenous carrier flotation technique for recovering finely ground chalcopyrite particles using coarse pyrite particles as a carrier. Miner. Eng. 2022, 180, 107518. [Google Scholar] [CrossRef]
  20. Xiao, Z.; Aftab, T.B.; Li, D. Applications of micro–nano bubble technology in environmental pollution control. Micro Nano Lett. 2019, 14, 782–787. [Google Scholar] [CrossRef]
  21. Hornn, V.; Ito, M.; Shimada, H.; Tabelin, C.B.; Jeon, S.; Park, I.; Hiroyoshi, N. Agglomeration–Flotation of Finely Ground Chalcopyrite using Emulsified Oil Stabilized by Emulsifiers: Implications for Porphyry Copper Ore Flotation. Metals 2020, 10, 912. [Google Scholar] [CrossRef]
  22. Hornn, V.; Ito, M.; Yamazawa, R.; Shimada, H.; Tabelin, C.B.; Jeon, S.; Park, I.; Hiroyoshi, N. Kinetic Analysis for Agglomeration-Flotation of Finely Ground Chalcopyrite: Comparison of First Order Kinetic Model and Experimental Results. Mater. Tran. 2020, 61, 1940–1948. [Google Scholar] [CrossRef]
  23. Hornn, V.; Park, I.; Ito, M.; Shimada, H.; Suto, T.; Tabelin, C.B.; Jeon, S.; Hiroyoshi, N. Agglomeration-flotation of finely ground chalcopyrite using surfactant-stabilized oil emulsions: Effects of co-existing minerals and ions. Miner. Eng. 2021, 171, 107076. [Google Scholar] [CrossRef]
  24. Dahan, A.M.E.; Alorro, R.D.; Pacaña, M.L.C.; Baute, R.M.; Silva, L.C.; Tabelin, C.B.; Resabal, V.J.T. Hydrochloric Acid Leaching of Philippine Coal Fly Ash: Investigation and Optimisation of Leaching Parameters by Response Surface Methodology (RSM). Sustain. Chem. 2022, 3, 76–90. [Google Scholar] [CrossRef]
  25. Dutt, T.; Kim, K.H.; Uchimiya, M.; Kwon, E.E.; Jeon, B.H.; Deep, A.; Yun, S.T. Global demand for rare earth resources and strategies for green mining. Environ. Res. 2016, 150, 182–190. [Google Scholar] [CrossRef]
  26. Binnemans, K.; Jones, P.T.; Blanpain, B.; Van Gerven, T.; Yang, Y.; Walton, A.; Buchert, M. Recycling of rare earths: A critical review. J. Clean. Prod. 2013, 51, 1005702. [Google Scholar] [CrossRef]
  27. Moher, D.; Liberati, A.; Tetzlaff, J.; Altman, D.G. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. BMJ 2009, 339, b2535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Andrews, R. The place of systematic reviews in education research. Br. J. Educ. Stud. 2005, 53, 399–416. [Google Scholar] [CrossRef]
  29. Orris, G.J.; Grauch, R.I. Rare Earth Element Mines, Deposits, and Occurrences, Open-File; Report 02-189; U.S. Geological Survey: Reston, VA, USA, 2002.
  30. Jindal-Snape, D.; Hannah, E.F.S.; Cantali, D.; Barlow, W.; MacGillivray, S. Systematic literature review of primary-secondary transitions: International research. Rev. Educ. 2020, 8, 526–566. [Google Scholar] [CrossRef]
  31. Liu, W.; Wang, X.; Wang, Z.; Miller, J. Flotation chemistry features in bastnaesite flotation with potassium lauryl phosphate. Miner. Eng. 2016, 85, 17–22. [Google Scholar] [CrossRef] [Green Version]
  32. Espiritu, E.; Naseri, S.; Waters, K. Surface chemistry and flotation behavior of dolomite, monazite and bastnäsite in the presence of benzohydroxamate, sodium oleate and phosphoric acid ester collectors. Colloids Surf. A Physicochem. Eng. Asp. 2018, 546, 254–265. [Google Scholar] [CrossRef]
  33. Espiritu, E.; da Silva, G.; Azizi, D.; Larachi, F.; Waters, K. Flotation behavior and electronic simulations of rare earth minerals in the presence of dolomite supernatant using sodium oleate collector. J. Rare Earths 2019, 37, 101–112. [Google Scholar] [CrossRef]
  34. Espiritu, E.; Waters, K. Flotation studies of monazite and dolomite. Miner. Eng. 2018, 116, 101–106. [Google Scholar] [CrossRef]
  35. Chen, W.; Zhou, F.; Wang, H.; Zhou, S.; Yan, C. The occurrence states of rare earth elements bearing phosphorite ores and rare earth enrichment through the selective reverse flotation. Minerals 2019, 9, 698. [Google Scholar] [CrossRef] [Green Version]
  36. McEvoy, J.G.; Thibault, Y. Impact of crystal chemistry properties on the collector-mineral interactions observed for REE orthophosphates and oxides. Appl. Surf. Sci. 2019, 466, 970–981. [Google Scholar] [CrossRef]
  37. Srinivasan, S.G.; Shivaramaiah, R.; Kent, P.R.C.; Stack, A.G.; Navrotsky, A.; Riman, R.; Anderko, A.; Bryantsev, V.S. Crystal structures, surface stability, and water adsorption energies of La-bastnäsite via density functional theory and experimental studies. J. Phys. Chem. C 2016, 120, 16767–16781. [Google Scholar] [CrossRef]
  38. Srinivasan, S.G.; Shivaramaiah, R.; Kent, P.R.C.; Stack, A.G.; Riman, R.; Anderko, A.; Navrotsky, A.; Bryantsev, V.S. A comparative study of surface energies and water adsorption on Ce-bastnasite, La-bastnaasite, and calcite via density functional theory and water adsorption calorimetry. Phys. Chem. Chem. Phys. 2017, 19, 7820–7832. [Google Scholar] [CrossRef] [PubMed]
  39. Pradip; Fuerstenau, D.W. Design and development of novel flotation reagents for the beneficiation of Mountain Pass rare-earth ore. Min. Metall. Explor. 2013, 30, 1–9. [Google Scholar] [CrossRef]
  40. Wang, Z.; Wu, H.; Xu, Y.; Shu, K.; Fang, S.; Xu, L. The effect of dissolved calcite species on the flotation of bastnaesite using sodium oleate. Miner. Eng. 2020, 145, 106095. [Google Scholar] [CrossRef]
  41. Wang, Z.; Wu, H.; Xu, Y.; Shu, K.; Yang, J.; Luo, L.; Xu, L. Effect of dissolved fluorite and barite species on the flotation and adsorption behavior of bastnaesite. Sep. Purif. Technol. 2020, 237, 116387. [Google Scholar] [CrossRef]
  42. Filippov, L.; Dehaine, Q.; Filippova, I. Rare earths (La, Ce, Nd) and rare metals (Sn, Nb, W) as by-products of kaolin production—Part 3: Processing of fines using gravity and flotation. Miner. Eng. 2016, 95, 96–106. [Google Scholar] [CrossRef]
  43. Geneyton, A.; Filippov, L.; Heinig, T.; Buaron, N.; Menad, N.-E. Towards the efficient flotation of monazite from silicate-rich tailings with fatty acids collectors using a lanthanum salt as a selective phosphate activator. Miner. Eng. 2021, 160, 106704. [Google Scholar] [CrossRef]
  44. Zhang, W.; Honaker, R.; Groppo, J. Concentration of rare earth minerals from coal by froth flotation. Miner. Metall. Process. 2017, 34, 132–137. [Google Scholar] [CrossRef]
  45. Zhang, W.; Yang, X.; Honaker, R.Q. Association characteristic study and preliminary recovery investigation of rare earth elements from Fire Clay seam coal middlings. Fuel 2018, 215, 551–560. [Google Scholar] [CrossRef]
  46. Satur, J.V.; Calabia, B.P.; Hoshino, M.; Morita, S.; Seo, Y.; Kon, Y.; Takagi, T.; Watanabe, Y.; Mutele, L.; Foya, S. Flotation of rare earth minerals from silicate–hematite ore using tall oil fatty acid collector. Miner. Eng. 2016, 89, 52–62. [Google Scholar] [CrossRef]
  47. Wang, X.; Zhang, Q.; Mao, S.; Cheng, W. A theoretical study on the electronic structure and floatability of rare earth elements (La, Ce, Nd and Y) bearing fluorapatite. Minerals 2019, 9, 500. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, Z.; Wu, H.; Yang, J.; Tang, Z.; Luo, L.; Shu, K.; Xu, Y.; Xu, L. Selective flotation separation of bastnaesite from calcite using xanthan gum as a depressant. Appl. Surf. Sci. 2020, 512, 145714. [Google Scholar] [CrossRef]
  49. Abaka-Wood, G.B.; Zanin, M.; Addai-Mensah, J.; Skinner, W. The upgrading of rare earth oxides from iron-oxide silicate rich tailings: Flotation performance using sodium oleate and hydroxamic acid as collectors. Adv. Powder Technol. 2018, 29, 3163–3172. [Google Scholar] [CrossRef]
  50. Abaka-Wood, G.B.; Addai-Mensah, J.; Skinner, W. A study of flotation characteristics of monazite, hematite, and quartz using anionic collectors. Int. J. Miner. Process. 2017, 158, 55–62. [Google Scholar] [CrossRef]
  51. Abaka-Wood, G.B.; Fosu, S.; Addai-Mensah, J.; Skinner, W. Flotation recovery of rare earth oxides from hematite–quartz mixture using sodium oleate as a collector. Miner. Eng. 2019, 141, 105847. [Google Scholar] [CrossRef]
  52. Zhang, Y.; Anderson, C. A comparison of sodium silicate and ammonium lignosulfonate effects on xenotime and selected gangue mineral microflotation. Miner. Eng. 2017, 100, 1–8. [Google Scholar] [CrossRef] [Green Version]
  53. Jordens, A.; Marion, C.; Kuzmina, O.; Waters, K.E. Physicochemical aspects of allanite flotation. J. Rare Earths 2014, 32, 476–486. [Google Scholar] [CrossRef]
  54. Malas, K.; Jordens, A.; Mirnezami, M.; Chu, P.; Gauvin, R.; Waters, K. Surface characterisation of fergusonite. Can. Metall. Q. 2013, 52, 278–284. [Google Scholar] [CrossRef]
  55. Tranvik, E.; Becker, M.; Pålsson, B.; Franzidis, J.-P.; Bradshaw, D. Towards cleaner production—Using flotation to recover monazite from a heavy mineral sands zircon waste stream. Miner. Eng. 2017, 101, 30–39. [Google Scholar] [CrossRef]
  56. Abaka-Wood, G.B.; Zanin, M.; Addai-Mensah, J.; Skinner, W. Recovery of rare earth elements minerals from iron oxide–silicate rich tailings—Part 2: Froth flotation separation. Miner. Eng. 2019, 142, 105888. [Google Scholar] [CrossRef]
  57. Fawzy, M.M. Surface characterization and froth flotation of fergusonite from Abu Dob pegmatite using a combination of anionic and nonionic collectors. Physicochem. Probl. Miner. Process. 2018, 54, 677–687. [Google Scholar] [CrossRef]
  58. Al-Thyabat, S.; Zhang, P. Extraction of rare earth elements from upgraded phosphate flotation tailings. Miner. Metall. Process. 2016, 33, 23–30. [Google Scholar] [CrossRef]
  59. Abaka-Wood, G.B.; Addai-Mensah, J.; Skinner, W. Selective flotation of rare earth oxides from hematite and quartz mixtures using oleic acid as a collector. Int. J. Miner. Process. 2017, 169, 60–69. [Google Scholar] [CrossRef]
  60. Zhang, P.; Liang, H.; Jin, Z.; DePaoli, D. The ultimate mineral processing challenge: Recovery of rare earths, phosphorus and uranium from Florida phosphatic clay. Miner. Metall. Process. 2017, 34, 183–188. [Google Scholar] [CrossRef]
  61. Gupta, T.; Ghosh, T.; Akdogan, G.; Srivastava, V. Characterizing rare earth elements in Alaskan coal and ash. Miner. Metall. Process. 2017, 34, 138–145. [Google Scholar] [CrossRef]
  62. Yang, X.; Satur, J.V.; Sanematsu, K.; Laukkanen, J.; Saastamoinen, T. Beneficiation studies of a complex REE ore. Miner. Eng. 2015, 71, 55–64. [Google Scholar] [CrossRef]
  63. Cui, J.; Hope, G.A.; Buckley, A.N. Spectroscopic investigation of the interaction of hydroxamate with bastnaesite (cerium) and rare earth oxides. Miner. Eng. 2012, 36–38, 91–99. [Google Scholar] [CrossRef]
  64. Owens, C.L.; Schach, E.; Heinig, T.; Rudolph, M.; Nash, G.R. Surface nanobubbles on the rare earth fluorcarbonate mineral synchysite. J. Colloid Interface Sci. 2019, 552, 66–71. [Google Scholar] [CrossRef]
  65. Wanhala, A.K.; Doughty, B.; Bryantsev, V.S.; Wu, L.; Mahurin, S.M.; Jansone-Popova, S.; Cheshire, M.C.; Navrotsky, A.; Stack, A.G. Adsorption mechanism of alkyl hydroxamic acid onto bastnäsite: Fundamental steps toward rational collector design for rare earth elements. J. Colloid Interface Sci. 2019, 553, 210–219. [Google Scholar] [CrossRef]
  66. Zhang, X.; Du, H.; Wang, X.; Miller, J.D. Surface chemistry considerations in the flotation of rare-earth and other semisoluble salt minerals. Miner. Metall. Process. 2013, 30, 24–37. [Google Scholar] [CrossRef]
  67. Zhang, J.; An, D.; Withers, J. A Micro-scale investigation of the adsorption of collectors on bastnaesite. Min. Metall. Explor. 2019, 36, 957–966. [Google Scholar] [CrossRef]
  68. Sime, M.; Das, A.; Galt, G.; Hope, G.; Young, C. Studies on the nature of salicyl hydroxamate adsorption at the surface of neodymium oxide. J. Dispers. Sci. Technol. 2019, 40, 1488–1498. [Google Scholar] [CrossRef]
  69. Cao, S.; Cao, Y.; Liao, Y.; Ma, Z. Depression mechanism of strontium ions in bastnaesite flotation with salicylhydroxamic acid as collector. Minerals 2018, 8, 66. [Google Scholar] [CrossRef] [Green Version]
  70. Espiritu, E.; da Silva, G.; Azizi, D.; Larachi, F.; Waters, K. The effect of dissolved mineral species on bastnäsite, monazite and dolomite flotation using benzohydroxamate collector. Colloids Surf. A Physicochem. Eng. Asp. 2018, 539, 319–334. [Google Scholar] [CrossRef]
  71. Lin, Y.; Chen, C.; Wang, W.; Jiang, Y.; Cao, X. Beneficial effects and mechanism of lead ions for bastnaesite flotation with octyl hydroxamic acid collector. Miner. Eng. 2020, 148, 106199. [Google Scholar] [CrossRef]
  72. Zhang, W.; Honaker, R. A fundamental study of octanohydroxamic acid adsorption on monazite surfaces. Int. J. Miner. Process. 2017, 164, 26–36. [Google Scholar] [CrossRef]
  73. Zhang, W.; Honaker, R. Surface charge of rare earth phosphate (monazite) in aqueous solutions. Powder Technol. 2017, 318, 263–271. [Google Scholar] [CrossRef]
  74. Nduwa-Mushidi, J.; Anderson, C. Surface chemistry and flotation behaviors of monazite–apatite–ilmenite–quartz–rutile–zircon with octanohydroxamic acid. J. Sustain. Metall. 2017, 3, 62–72. [Google Scholar] [CrossRef]
  75. Duan, H.; Liu, W.; Wang, X.; Liu, W.; Zhang, N.; Zhou, S. Flotation separation of bastnaesite from calcite using novel octylmalon dihydroxamic acid as collector. J. Mol. Liq. 2020, 312, 113484. [Google Scholar] [CrossRef]
  76. Cao, Z.; Cao, Y.; Qu, Q.; Zhang, J.; Mu, Y. Separation of bastnäsite from fluorite using ethylenediamine tetraacetic acid as depressant. Miner. Eng. 2019, 134, 134–141. [Google Scholar] [CrossRef]
  77. Wang, J.; Cheng, Z.; Hu, Y.; Cao, Y.; Wang, P.; Cao, Z. Depression behavior and mechanism of sodium silicate on bastnaesite and parisite flotation. Colloids Surf. A Physicochem. Eng. Asp. 2021, 631, 127691. [Google Scholar] [CrossRef]
  78. Zheng, Q.; Qian, Y.; Zou, D.; Wang, Z.; Bai, Y.; Dai, H. Surface mechanism of Fe3+ ions on the improvement of fine monazite flotation with octyl hydroxamate as the collector. Front. Chem. 2021, 9, 700347. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, W.; Honaker, R.Q.; Groppo, J.G. Flotation of monazite in the presence of calcite part I: Calcium ion effects on the adsorption of hydroxamic acid. Miner. Eng. 2017, 100, 40–48. [Google Scholar] [CrossRef]
  80. Zhang, W.; Honaker, R.Q. Flotation of monazite in the presence of calcite part II: Enhanced separation performance using sodium silicate and EDTA. Miner. Eng. 2018, 127, 318–328. [Google Scholar] [CrossRef]
  81. Cui, H.; Anderson, C.G. Fundamental studies on the surface chemistry of ancylite, calcite, and strontianite. J. Sustain. Metall. 2017, 3, 48–61. [Google Scholar] [CrossRef]
  82. Cui, H.; Anderson, C.G. Alternative flowsheet for rare earth beneficiation of Bear Lodge ore. Miner. Eng. 2017, 110, 166–178. [Google Scholar] [CrossRef]
  83. Chapleski, R.C., Jr.; Chowdhury, A.U.; Wanhala, A.K.; Bocharova, V.; Roy, S.; Keller, P.C.; Everly, D.; Jansone-Popova, S.; Kisliuk, A.; Sacci, R.L.; et al. A molecular-scale approach to rare-earth beneficiation: Thinking small to avoid large losses. iScience 2020, 23, 101435. [Google Scholar] [CrossRef]
  84. Jordens, A.; Marion, C.; Kuzmina, O.; Waters, K.E. Surface chemistry considerations in the flotation of bastnäsite. Miner. Eng. 2014, 66–68, 119–129. [Google Scholar] [CrossRef]
  85. Jordens, A.; Marion, C.; Grammatikopoulos, T.; Hart, B.; Waters, K.E. Beneficiation of the Nechalacho rare earth deposit: Flotation response using benzohydroxamic acid. Miner. Eng. 2016, 99, 158–169. [Google Scholar] [CrossRef]
  86. Boulanger, J.-F.; Bazin, C.; Turgeon, K. Effect of depressants and temperature on bastnaesite and monazite flotation separation from a Canadian rare earth element (REE) ore. Minerals 2019, 9, 225. [Google Scholar] [CrossRef] [Green Version]
  87. Xiong, W.; Deng, J.; Zhao, K.; Wang, W.; Wang, Y.; Wei, D. Bastnaesite, barite, and calcite flotation behaviors with salicylhydroxamic acid as the collector. Minerals 2020, 10, 282. [Google Scholar] [CrossRef] [Green Version]
  88. Jiao, Y.; Qiu, K.-H.; Zhang, P.-C.; Li, J.-F.; Zhang, W.-T.; Chen, X.-F. Process mineralogy of Dalucao rare earth ore and design of beneficiation process based on AMICS. Rare Metals 2020, 39, 959–966. [Google Scholar] [CrossRef]
  89. Chang, S.; Zhang, J.; Liu, Z.; Gao, K.; Zhang, D.; Huang, J. Activation mechanisms of apatite by purifying reagent polyacrylamide for backwater from mineral processing in flotation of rare earths. Adv. Powder Technol. 2021, 32, 4447–4460. [Google Scholar] [CrossRef]
  90. Zhu, X.; Huang, Y.; Zhu, Y.; Sun, N.; Wang, W. Investigating the performance of oxalic acid for separating bastnaesite from calcium-bearing gangue minerals based on experiment and theoretical calculation. Miner. Eng. 2021, 170, 107047. [Google Scholar] [CrossRef]
  91. Yu, B.; Che, X.; Zheng, Q. Flotation of ultra-fine rare-earth minerals with selective flocculant PDHA. Miner. Eng. 2014, 60, 23–25. [Google Scholar] [CrossRef]
  92. Yang, Z.; Wu, W.; Bian, X. Synthesis of 3-hydroxy-2-naphthyl hydroxamic acid collector: Flotation performance and adsorption mechanism on bastnaesite. J. S. Afr. Inst. Min. Metall. 2017, 117, 593–598. [Google Scholar] [CrossRef] [Green Version]
  93. Xiong, W.; Deng, J.; Chen, B.; Deng, S.; Wei, D. Flotation-magnetic separation for the beneficiation of rare earth ores. Miner. Eng. 2018, 119, 49–56. [Google Scholar] [CrossRef]
  94. Sarvaramini, A.; Azizi, D.; Larachi, F. Hydroxamic acid interactions with solvated cerium hydroxides in the flotation of monazite and bastnäsite—Experiments and DFT study. Appl. Surf. Sci. 2016, 387, 986–995. [Google Scholar] [CrossRef]
  95. Azizi, D.; Larachi, F.; Latifi, M. Ionic-liquid collectors for rare-earth minerals flotation—Case of tetrabutylammonium bis(2-ethylhexyl)-phosphate for monazite and bastnäsite recovery. Colloids Surf. A Physicochem. Eng. Asp. 2016, 506, 74–86. [Google Scholar] [CrossRef]
  96. Abaka-Wood, G.B.; Addai-Mensah, J.; Skinner, W. A study of selective flotation recovery of rare earth oxides from hematite and quartz using hydroxamic acid as a collector. Adv. Powder Technol. 2018, 29, 1886–1899. [Google Scholar] [CrossRef]
  97. Yu, X.; Zhang, R.; Yang, S.; Liu, C.; He, G.; Wang, H.; Wang, J. A novel decanedioic hydroxamic acid collector for the flotation separation of bastnäsite from calcite. Miner. Eng. 2020, 151, 106306. [Google Scholar] [CrossRef]
  98. Qi, J.; Fan, H.; Liu, G. β-Amino-hydroxamate surfactants: Preparation, and adsorption mechanism in bastnaesite flotation. Sep. Purif. Technol. 2020, 240, 116634. [Google Scholar] [CrossRef]
  99. Qi, J.; Liu, G.; Dong, Y. Probing the hydrophobic mechanism of N-[(3-hydroxyamino)-propoxy]-N-octyl dithiocarbamate toward bastnaesite flotation by in situ AFM, FTIR and XPS. J. Colloid Interface Sci. 2020, 572, 179–189. [Google Scholar] [CrossRef]
  100. Duan, H.; Liu, W.; Wang, X.; Zhao, L.; Fang, P.; Gu, X. Preparation of a novel bis hydroxamic collector and its impact on bastnaesite flotation. Miner. Eng. 2020, 156, 106496. [Google Scholar] [CrossRef]
  101. Duan, H.; Liu, W.; Wang, X.; Gu, X.; Sun, W.; Peng, X.; Yue, H. Investigation on flotation separation of bastnaesite from calcite and barite with a novel surfactant: Octylamino-bis-(butanohydroxamic acid). Sep. Purif. Technol. 2021, 256, 117792. [Google Scholar] [CrossRef]
  102. Guo, W.; Cang, D.; Zhang, L.; Guo, J. Mechanism of recovery processes for rare earth and iron from Bayan Obo tailings. Int. J. Chem. React. Eng. 2020, 18, 20200077. [Google Scholar] [CrossRef]
  103. Li, M.; Gao, K.; Zhang, D.; Duan, H.; Ma, L.; Huang, L. The influence of temperature on rare earth flotation with naphthyl hydroxamic acid. J. Rare Earths 2018, 36, 99–107. [Google Scholar] [CrossRef]
  104. Xia, L.; Hart, B.; Loshusan, B. A Tof-SIMS analysis of the effect of lead nitrate on rare earth flotation. Miner. Eng. 2015, 70, 119–129. [Google Scholar] [CrossRef]
  105. Zhang, Y.; Lin, H.; Dong, Y.-B.; Xu, X.-F.; Wang, X.; Gao, Y.-J. Coupling relationship between multicomponent recovery of rare earth tailings. Rare Metals 2017, 36, 220–228. [Google Scholar] [CrossRef]
  106. Liu, W.; Wang, X.; Xu, H.; Miller, J. Lauryl phosphate adsorption in the flotation of bastnaesite, (Ce,La)FCO3. J. Colloid Interface Sci. 2017, 490, 825–833. [Google Scholar] [CrossRef] [Green Version]
  107. Cao, S.; Cao, Y.; Ma, Z.; Liao, Y.; Zhang, X. Structural and electronic properties of bastnaesite and implications for surface reactions in flotation. J. Rare Earths 2020, 38, 332–338. [Google Scholar] [CrossRef]
  108. Chelgani, S.C.; Hart, B.; Xia, L. A TOF-SIMS surface chemical analytical study of rare earth element minerals from micro-flotation tests products. Miner. Eng. 2013, 45, 32–40. [Google Scholar] [CrossRef]
  109. Sutton, J.E.; Roy, S.; Chowdhury, A.U.; Wu, L.; Wanhala, A.K.; Silva, N.D.; Jansone-Popova, S.; Hay, B.P.; Cheshire, M.C.; Windus, T.L.; et al. Molecular recognition at mineral interfaces: Implications for the beneficiation of rare earth ores. ACS Appl. Mater. Interfaces 2020, 12, 16327–16341. [Google Scholar] [CrossRef]
  110. Xia, L.; Hart, B.; Douglas, K. The role of citric acid in the flotation separation of rare earth from the silicates. Miner. Eng. 2015, 74, 123–129. [Google Scholar] [CrossRef]
  111. Liu, W.; Wang, X.; Xu, H.; Miller, J. Physical chemistry considerations in the selective flotation of bastnaesite with lauryl phosphate. Miner. Metall. Process. 2017, 34, 116–124. [Google Scholar] [CrossRef]
  112. Liu, W.; McDonald, L.W., IV; Wang, X.; Miller, J.D. Bastnaesite flotation chemistry issues associated with alkyl phosphate collectors. Miner. Eng. 2018, 286–295. [Google Scholar] [CrossRef]
  113. Fan, H.; Yang, X.; Qi, J.; Liu, G.; Qin, J. A comparative investigation into floatability of bastnaesite with three di/trialkyl phosphate surfactants. J. Rare Earths 2021, 39, 1442–1449. [Google Scholar] [CrossRef]
  114. Fan, H.; Tan, W.; Liu, G. 1-Hydroxydodecylidene-1,1-diphosphonic acid flotation of bastnäsite: Performance and mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2021, 609, 125623. [Google Scholar] [CrossRef]
  115. Yang, Z.; Bian, X.; Wu, W. Flotation performance and adsorption mechanism of styrene phosphonic acid as a collector to synthetic (Ce,La)2O3. J. Rare Earths 2017, 35, 621–628. [Google Scholar] [CrossRef]
  116. Zhou, F.; Wang, L.; Xu, Z.; Liu, Q.; Chi, R. Interaction of reactive oily bubble in flotation of bastnaesite. J. Rare Earths 2014, 32, 772–778. [Google Scholar] [CrossRef]
  117. Zhou, F.; Wang, L.; Xu, Z.; Liu, Q.; Deng, M.; Chi, R. Application of reactive oily bubbles to bastnaesite flotation. Miner. Eng. 2014, 64, 139–145. [Google Scholar] [CrossRef]
  118. Xu, Y.; Xu, L.; Wu, H.; Wang, Z.; Shu, K.; Fang, S.; Zhang, Z. Flotation and co-adsorption of mixed collectors octanohydroxamic acid/sodium oleate on bastnaesite. J. Alloys Compd. 2020, 819, 152948. [Google Scholar] [CrossRef]
  119. Xu, Y.; Xu, L.; Wu, H.; Tian, J.; Wang, Z.; Gu, X. The effect of citric acid in the flotation separation of bastnaesite from fluorite and calcite using mixed collectors. Appl. Surf. Sci. 2020, 529, 147166. [Google Scholar] [CrossRef]
  120. Li, R.; Marion, C.; Espiritu, E.; Multani, R.; Sun, X.; Waters, K. Investigating the use of an ionic liquid for rare earth mineral flotation. J. Rare Earths 2021, 39, 866–874. [Google Scholar] [CrossRef]
  121. Kursun, I.; Terzi, M.; Ozdemir, O. Determination of surface chemistry and flotation properties of rare earth mineral allanite. Miner. Eng. 2019, 132, 113–120. [Google Scholar] [CrossRef]
  122. Gupta, T.; Ghosh, T.; Akdogan, G.; Bandopadhyay, S. Maximizing REE enrichment by froth flotation of Alaskan coal using Box-Behnken design. Min. Metall. Explor. 2019, 36, 571–578. [Google Scholar] [CrossRef]
  123. Chelgani, S.C.; Hart, B. Explaining surface interactions for common associated gangues of rare earth minerals in response to the oxalic acid. Int. J. Min. Sci. Technol. 2018, 28, 343–346. [Google Scholar] [CrossRef]
  124. Zhang, Y.; Anderson, C. Surface chemistry and microflotation of xenotime and selected gangue minerals using octanohydroxamic acid as the collector. J. Sustain. Metall. 2017, 3, 39–47. [Google Scholar] [CrossRef]
  125. Li, C.C.; Fuerstenau, D.W. Surface chemical characterization of bastnaesite through electrokinetics. KONA Powder Part. J. 2015, 32, 176–183. [Google Scholar] [CrossRef] [Green Version]
  126. Cao, S.; Cao, Y.; Ma, Z.; Liao, Y. The adsorption mechanism of Al(III) and Fe(III) ions on bastnaesite surfaces. Physicochem. Probl. Miner. Process. 2019, 55, 97–107. [Google Scholar] [CrossRef]
  127. Geneyton, A.; Filippov, L.; Renard, A.; Mallet, M.; Menad, N.-E. Advances in carboxylate collectors adsorption on monazite surface: Part 1—Assessment of the hydroxylation and carbonation of surface lanthanide ions. Appl. Surf. Sci. 2019, 485, 283–292. [Google Scholar] [CrossRef]
  128. Geneyton, A.; Foucaud, Y.; Filippov, L.; Menad, N.-E.; Renard, A.; Badawi, M. Synergistic adsorption of lanthanum ions and fatty acids for efficient rare earth phosphate recovery: Surface analysis and ab initio molecular dynamics studies. Appl. Surf. Sci. 2020, 526, 146725. [Google Scholar] [CrossRef]
  129. Zhu, X.; Lin, Y.; Huang, Y.; Zhu, Y.; Shi, C.; Wang, W. Adsorption of ferric ions on the surface of bastnaesite and its significance in flotation. Miner. Eng. 2020, 158, 106588. [Google Scholar] [CrossRef]
  130. Yu, M.; Mei, G.; Li, Y.; Liu, D.; Peng, Y. Recovering rare earths from waste phosphors using froth flotation and selective flocculation. Miner. Metall. Process. 2017, 34, 161–169. [Google Scholar] [CrossRef]
  131. Hogan, D.E.; Curry, J.E.; Maier, R.M. Ion flotation of La3+, Cd2+, and Cs+ using monorhamnolipid collector. Colloids Interfaces 2018, 2, 43. [Google Scholar] [CrossRef] [Green Version]
  132. Justeau, C.; Vela-Gonzalez, A.V.; Jourdan, A.; Riess, J.G.; Krafft, M.P. Adsorption of cerium salts and cerium oxide nanoparticles on microbubbles can be induced by a fluorocarbon gas. ACS Sustain. Chem. Eng. 2018, 6, 11450–11456. [Google Scholar] [CrossRef]
  133. Khatir, M.Z.; Abdollahy, M.; Khalesi, M.R.; Rezai, B. Selective separation of neodymium from synthetic wastewater by ion flotation. Sep. Sci. Technol. 2021, 56, 1802–1810. [Google Scholar] [CrossRef]
  134. Khatir, M.Z.; Abdollahy, M.; Khalesi, M.R.; Rezai, B. Role of D2EHPA in ion flotation of neodymium for achieving a high selectivity over base metal impurities. JOM 2021, 74, 240–248. [Google Scholar] [CrossRef]
  135. Lobacheva, O.; Dzhevaga, N.; Danilov, A. The method of removal yttrium (III) and ytterbium (III) from dilute aqueous solutions. J. Ecol. Eng. 2016, 17, 38–42. [Google Scholar] [CrossRef] [Green Version]
  136. Dzhevaga, N.; Lobacheva, O. Reduction in technogenic burden on the environment by flotation recovery of rare earth elements from diluted industrial solutions. Appl. Sci. 2021, 11, 7452. [Google Scholar] [CrossRef]
  137. Lobacheva, O.L. Ion flotation of ytterbium water-salt systems—An innovative aspect of the modern industry. Water 2021, 13, 3493. [Google Scholar] [CrossRef]
  138. Lobacheva, O.L.; Berlinskii, I.V.; Chere, O.V. Solvent sublation and ion flotation in aqueous salt solutions containing Ce(III) and Y(III) in the presence of a surfactant. Russ. J. Appl. Chem. 2014, 87, 1863–1867. [Google Scholar] [CrossRef]
  139. Lobacheva, O.L.; Dzhevaga, N.V. The experimental study of innovative methods regarding the removal of Sm (III). Appl. Sci. 2021, 11, 7726. [Google Scholar] [CrossRef]
  140. Shetty, S.; Chernyshova, I.V.; Ponnurangam, S. Foam flotation of rare earth elements by conventional and green surfactants. Miner. Eng. 2020, 158, 106585. [Google Scholar] [CrossRef]
  141. Micheau, C.; Diat, O.; Bauduin, P. Ion foam flotation of neodymium: From speciation to extraction. J. Mol. Liq. 2018, 253, 217–227. [Google Scholar] [CrossRef]
  142. Mahmoud, M.R.; Soliman, M.A.; Rashad, G.M. Competitive foam separation of rare earth elements from aqueous solutions using a cationic collector. Sep. Sci. Technol. 2019, 54, 2374–2385. [Google Scholar] [CrossRef]
  143. Lobacheva, O.L.; Dzhevaga, N.V.; Chirkst, D.E. Solvent sublation of cerium ions from dilute aqueous solutions. Russ. Chem. Bull. 2012, 61, 962–965. [Google Scholar] [CrossRef]
  144. Lobacheva, O.; Berlinskii, I. Solvent sublation of the Tb (III) from aqueous solutions with sodium dodecyl sulfate. Int. J. Appl. Eng. Res. 2016, 11, 6350–6354. [Google Scholar]
  145. Chirkst, D.E.; Lobacheva, O.L.; Dzhevaga, N.V. Ion flotation of lanthanum (III) and holmium (III) from nitrate and nitrate-chloride media. Russ. J. Appl. Chem. 2012, 85, 25–28. [Google Scholar] [CrossRef]
  146. Gaidukov, E.N.; Kolesnikov, A.V.; Moshkina, D.S.; Kolesnikov, V.A. Electroflotation recovery of poorly soluble lanthanum compounds from highly concentrated salt systems. Russ. J. Appl. Chem. 2018, 91, 70–77. [Google Scholar] [CrossRef]
  147. Gaydukova, A.; Kolesnikov, A.; Gubin, A.; Kolesnikov, V. Selective separation of rare-earth metal ions from aqueous solutions on the example of slightly soluble cerium and lanthanum compounds using electroflotation method in a flow-through plant. Sep. Purif. Technol. 2019, 223, 260–263. [Google Scholar] [CrossRef]
  148. Kolesnikov, A.V.; Gaidukov, E.N.; Kolesn, V.A. Studying the efficiency of an electroflotation process for the extraction of low-soluble scandium compounds from aqueous media in the presence of surfactants. Theor. Found. Chem. Eng. 2016, 50, 678–683. [Google Scholar] [CrossRef]
  149. Gaydukova, A.M.; Brodskiy, V.A.; Volkova, V.V.; Kolesnikov, V.A. Selective separation and extraction of copper (II), iron (II, III), and Cerium (III, IV) ions from aqueous solutions by electroflotation method. Russ. J. Appl. Chem. 2017, 90, 1252–1257. [Google Scholar] [CrossRef]
  150. Kolesnikov, V.A.; Gaidukova, A.M.; Kolesnikov, A.V.; Gubin, A.F.; Vetlugin, N.A. Electroflotation extraction of sparingly soluble rare-earth compounds in a multicomponent mixture from aqueous solutions containing chloride ions. Theor. Found. Chem. Eng. 2020, 54, 856–862. [Google Scholar] [CrossRef]
  151. Meshalkin, V.P.; Kolesnikov, A.V.; Gaidukova, A.M.; Achkasov, M.G.; Kolesnikov, V.A.; Belozerskii, A.Y.; Men’shova, I.I. Experimental investigation of the efficiency of the electroflotation recovery of a mixture of a poorly soluble cerium group metal compounds from aqueous solutions. Phys. Chem. 2019, 487, 99–102. [Google Scholar] [CrossRef]
  152. Zhurinov, M.; Kolesnikov, V.; Bayeshov, A.; Abduvaliyeva, U.A.; Gaydukova, A.; Kolesnikov, A.; Kuznetsov, V. Electroflotation extraction of lanthanum and scandium hydroxides from aqueous solutions. News Natl. Acad. Sci. Repub. Kazakhstan Ser. Geol. Tech. Sci. 2020, 4, 94–101. [Google Scholar] [CrossRef]
  153. Kolesnikov, A.V.; Gaidukov, E.N.; Kolesnikov, V.A. The role of surfactants in intensifying and increasing the efficiency of the electroflotation extraction of sparingly soluble lanthanum compounds. Theor. Found. Chem. Eng. 2016, 50, 142–147. [Google Scholar] [CrossRef]
  154. Rezgar, B.; Mohamad, N.; Mahdi, G. Cerium extraction from solution by adsorbing colloid flotation (ACF). Iran. J. Chem. Chem. Eng. 2018, 37, 133–145. [Google Scholar] [CrossRef]
  155. Perlova, O.V.; Chernetskaya, V.V. Flotation recovery of rare-earth metal compounds from dilute aqueous solutions. J. Water Chem. Technol. 2015, 37, 90–95. [Google Scholar] [CrossRef]
Figure 1. A schematic illustration of a mechanical flotation cell.
Figure 1. A schematic illustration of a mechanical flotation cell.
Applsci 13 08364 g001
Figure 2. Schematic illustrations of (a) effects of Ca2+ inhibition, and (b) effects of Pb2+ activation of bastnaesite.
Figure 2. Schematic illustrations of (a) effects of Ca2+ inhibition, and (b) effects of Pb2+ activation of bastnaesite.
Applsci 13 08364 g002
Figure 3. Schematic illustrations of advanced flotation techniques used in REEs recycling: (a) ion flotation, (b) solvent sublation, (c) electroflotation, and (d) absorbing colloid flotation.
Figure 3. Schematic illustrations of advanced flotation techniques used in REEs recycling: (a) ion flotation, (b) solvent sublation, (c) electroflotation, and (d) absorbing colloid flotation.
Applsci 13 08364 g003
Table 3. A summary of recent studies on REM flotation using organo-phosphoric acid collectors.
Table 3. A summary of recent studies on REM flotation using organo-phosphoric acid collectors.
CollectorsSamplespHDepressantsRecovery [%]Grade [%]References
Phophoric acid ester (Flotinor 1682)Bastnaesite4–9> 95[32]
Monazite 4–9> 95[32]
Monazite and dolomite731[34]
Flotinor SM15Bastnaesite and quartz9Sodium silicate> 90[84]
Flotinor 1682; mixed with A845REMs7Citric acid6[110]
Lauryl phosphateBastnaesite4–589–91[31,106]
Bastnaesite, calcite, and quartz4–790[111]
58052[112]
2-ethylhexyl phosphateBastnaesite, calcite, and quartz57595[112]
Di(2-ethylhexyl) phosphateBastnaesite896[113]
Dibutyl phosphateBastnaesite837[113]
Tributyl phosphateBastnaesite822[113]
1-Hydroxydodecylidene-1,1-diphosphonicBastnaesite and calcite896[114]
Styrene phosphonic acidSynthetic (Ce,La)2O3, synthetic calcium fluorite, synthetic fluorapatite59072[115]
Table 4. A summary of recent studies on ion flotation and solvent sublation for REE ion recovery.
Table 4. A summary of recent studies on ion flotation and solvent sublation for REE ion recovery.
Flotation TechniquesREEsREE Concentration [M] pHSurfactantsFrothersEff. * [%]References
Ion flotationCe1 [mg/L][132]
La, Ce, Gd, and Yb(0.01–5.4) × 10–43NaDDS57–61[140]
Nd7 × 10–43NaDDS85–97[133,134]
Ce and Y1 × 10–36NaDDS[138]
Y and Yb1 × 10–3NaDDS[135]
Yb1 × 10–3NaDDS [137]
La and Y1 × 10–3NaDDS[136]
La and Ho1 × 10–3NaDDS[143]
Sm1 × 10–3NaDDS[139]
La, Ce, Gd, and Yb (0.01–5.4) × 10–49Monorhamnolipid90–95[140]
La(0.3–1.6) × 10–5MonorhamnolipidEtOH[131]
Eu and Tb2.5 × 10–5Cetylpyridinium chlorideEtOH[142]
Nd10–7–10–6AKYPO RO 90 VG[141]
Solvent sublationCe and Y1 × 10–37–8NaDDSIsooctyl alcohol99[138]
Ce1 × 10–38NaDDSIsooctyl alcohol99[144]
Tb1 × 10–37NaDDS2-octanol99[145]
Sm1 × 10–38NaDDS2-octanol98[139]
Note: NaDDS = sodium dodecyl sulfate; AKYPO® RO 90 VG = trade name of nonaoxyethylene oleylether carboxilic acid; EtOH = ethanol. * = separation efficiency.
Table 5. A summary of recent studies on electroflotation for REE ions recovery.
Table 5. A summary of recent studies on electroflotation for REE ions recovery.
ElectrolyteREEsREE Concentration [M] pHSurfactants; (Type)Eff. * [%]References
ChlorideSc507NaDDS; (−)99[148,152]
Septapav; (+)88[148]
DDAC; (+)88[152]
PEO-1500; (0)90[148,152]
La5010NaDDS; (−)83–98[146,148,152]
Septapav; (+)74–88[146,148]
PEO-1500; (0)53–95[146,148,152]
CarbonateSc507NaDDS; (−)91[148,152]
Septapav; (+)77[148]
DDAC; (+)77[152]
PEO-1500; (0)58[148,152]
La507NaDDS; (−)12–20[146]
Septapav; (+)29–34
DDAC; (+)3–70
Ce, La, Nd, Pr, and Sm3–1377Septapav; (+)97–99[151]
Carbonate-chlorideCe, La, Nd, Pr, Sm, and Gd200–2507NaDDS; (−)86–87[150]
Septapav; (+)85–90
PEO-1500; (0)83–90
NitrateSc507NaDDS; (−)64[148,152]
Septapav; (+)97
PEO-1500; (0)80
La5010NaDDS; (−)11–96[146,152,153]
Septapav; (+)14–98[146,153]
DDAC; (+)98[152]
PEO-1500; (0)9–96[146,152,153]
SulfateSc507NaDDS; (−)22[148,152]
Septapav; (+)45[148,152]
DDAC; (+)41[152]
PEO-1500; (0)88[148,152]
La5010NaDDS; (−)18–97[152,153]
Septapav; (+)38–97[153]
DDAC; (+)97[152]
PEO-1500; (0)14–56[152,153]
Note: (+) = cationic surfactant; (−) = anionic surfactant; (0) = nonionic surfactant; NaDDS = sodium dodecyl sulfate; Septapav = didecyldimethylammonium chloride; DDAC = dodecyldimethylammonium chloride; PEO-1500 = polyethyleneoxide; * = separation efficiency.
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

Julapong, P.; Numprasanthai, A.; Tangwattananukul, L.; Juntarasakul, O.; Srichonphaisarn, P.; Aikawa, K.; Park, I.; Ito, M.; Tabelin, C.B.; Phengsaart, T. Rare Earth Elements Recovery from Primary and Secondary Resources Using Flotation: A Systematic Review. Appl. Sci. 2023, 13, 8364. https://doi.org/10.3390/app13148364

AMA Style

Julapong P, Numprasanthai A, Tangwattananukul L, Juntarasakul O, Srichonphaisarn P, Aikawa K, Park I, Ito M, Tabelin CB, Phengsaart T. Rare Earth Elements Recovery from Primary and Secondary Resources Using Flotation: A Systematic Review. Applied Sciences. 2023; 13(14):8364. https://doi.org/10.3390/app13148364

Chicago/Turabian Style

Julapong, Pongsiri, Apisit Numprasanthai, Ladda Tangwattananukul, Onchanok Juntarasakul, Palot Srichonphaisarn, Kosei Aikawa, Ilhwan Park, Mayumi Ito, Carlito Baltazar Tabelin, and Theerayut Phengsaart. 2023. "Rare Earth Elements Recovery from Primary and Secondary Resources Using Flotation: A Systematic Review" Applied Sciences 13, no. 14: 8364. https://doi.org/10.3390/app13148364

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