Next Article in Journal
A Cippus from Turris Libisonis: Evidence for the Use of Local Materials in Roman Painting on Stone in Northern Sardinia
Next Article in Special Issue
Mineral Liberation and Concentration Characteristics of Apatite Comminuted by High-Pressure GRU
Previous Article in Journal
Control of Seepage Characteristics in Loose Sandstone Heap Leaching with Staged Particle Sieving-Out Method
Previous Article in Special Issue
Review on the Challenges of Magnesium Removal in Nickel Sulfide Ore Flotation and Advances in Serpentinite Depressor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Particle Size on Flotation Separation of Ilmenite and Forsterite

1
Institute of Multipurpose Utilization of Mineral Resources, Chinese Academy of Geological Sciences (CAGS), Chengdu 610041, China
2
Technology Innovation Center for Comprehensive Utilization of Strategic Mineral Resources, Ministry of Natural Resources, Chengdu 610041, China
3
Key Laboratory of Solid Waste Treatment and Resource Recycle, Ministry of Education, Southwest University of Science and Technology, Mianyang 621010, China
*
Authors to whom correspondence should be addressed.
Minerals 2024, 14(10), 1041; https://doi.org/10.3390/min14101041
Submission received: 25 September 2024 / Revised: 14 October 2024 / Accepted: 16 October 2024 / Published: 17 October 2024

Abstract

:
In addition to bubble–particle interaction, particle–particle interaction also has a significant influence on mineral flotation. Fine particles that coat the mineral surface prevent direct contact with collectors and/or air bubbles, thereby lowering flotation recovery. Calculating the particle interaction energy can help in evaluating the interaction behavior of particles. In this study, the floatability of coarse ilmenite (−151 + 74 μm) and different particle sizes (−45 + 25, −25 + 19, −19 μm) of forsterite with NaOL as a collector was investigated. The results showed that forsterite sizes of −45 + 25 and −25 + 19 μm had no effect on the ilmenite floatability, whereas −19 μm forsterite significantly reduced ilmenite floatability. A particle size analysis of artificially mixed minerals and a scanning electron microscopy (SEM) analysis of the flotation products showed that heterogeneous aggregation occurred between ilmenite and −19 μm forsterite particles. The extended DLVO (Derjaguin–Landau–Verwey–Overbeek) theory was applied to calculate the interaction energy between mineral particles using data from zeta potential and contact angle measurements. The results showed that the interaction barriers between ilmenite (−151 + 74 μm) and forsterite (−45 + 25, −25 + 19, and −19 μm) were 11.94 × 103 kT, 8.23 × 103 kT and 4.09 × 103 kT, respectively. Additionally, the interaction barrier between forsterite particles smaller than 19 μm was 0.51 × 103 kT. The strength of the barrier decreased as the size of the forsterite decreased. Therefore, fine forsterite particles and aggregated forsterite can easily overcome the energy barrier, coating the ilmenite particle surface. This explains the effect of different forsterite sizes on the floatability of ilmenite and the underlying mechanism of particle interaction.

1. Introduction

Titanium is widely distributed around the world and is used in strategic emerging fields such as aerospace, navigation, energy, construction, transportation, medicine, and materials [1,2,3]. Ilmenite is an important titanium-bearing mineral with the molecular formula FeTiO3. It has been reported that more than 90% of the world’s titanium resources come from ilmenite [4]. China’s titanium reserves rank among the largest in the world, and the country has developed into the world’s largest producer of titanium concentrate. Titanium resources in China are mainly found in the Panxi area, and it is reported that this area contains up to 93% of China’s total titanium reserves [5,6]. However, most of the ores in the Panxi area are difficult to recover ilmenite from [7]. As the mining process progresses deeper, the original gabbro-type iron ore is gradually transformed into forsterite (Mg2SiO4) pyroxene-type iron ore. This transformation results in a finer particle size distribution of ilmenite, a notable increase in forsterite content, and a more intricate mineral composition [8]. Therefore, it is necessary to investigate the influence of forsterite particle size on ilmenite recovery.
Increasing attention has been paid to the effect of fine particles on mineral flotation separation. Fine particles have a relatively high surface area, and more reagents are needed for processing. Prior studies showed that the collision probability between fine particles and bubbles is low [9], and fine particles exhibit a reduced tendency of adhesion to bubbles, necessitating greater kinetic energy and a prolonged induction time [10,11,12]. Therefore, flotation recovery and selectivity are deteriorated under normal conditions [13,14]. This circumstance is unfavorable for flotation separation from non-target minerals. On the other hand, fine gangue particles can coat the surface of useful minerals, preventing agents or bubbles from acting on them, thereby reducing flotation recovery [15,16]. The influence of fine particles on the flotation process and improvement methods has been studied in slime coating [17,18], carrier flotation [19,20], and flocculation flotation [21].
Zhu et al. [22] studied the autogenous-carrier flotation of ilmenite. They measured the particle size distribution before and after conditioning with NaOL at a concentration of 1 × 10⁻4 M, an ilmenite autogenous-carrier particle size of −38 + 74 µm, and a pH range of 4.5–5.5 using a laser particle size analyzer. They found that fine and coarse ilmenite particles can aggregate, thereby improving the recovery rate of fine ilmenite. Yang et al. [23] studied the influence of pyroxene and olivine particle size on the flotation recovery of ilmenite from the Panxi area. They found that the fine particle content has a significant deterioration effect on the flotation recovery of ilmenite, and they suggested that a narrow particle size distribution of minerals is beneficial to flotation recovery. Fan and Cao [24] calculated the interaction force between ilmenite and titanaugite particles through extended DLVO. Their results showed that aggregation occurs easily between ilmenite and fine titanaugite particles, thus reducing the recovery rate of ilmenite. To date, there has been extended research on the interaction between titanaugite and ilmenite particles. However, the interaction between fine forsterite, another main gangue mineral, and ilmenite particles, along with its mechanism, has not been studied.
Lange et al. [25] used an online particle size distribution technique to obtain the change in particle size in conditioning and used optical microscopy to observe the flotation products. Their results showed that there is aggregation between coarse and fine particles. Due to the difference in elemental composition between valuable minerals and gangue minerals, SEM/EDS is used as a combined technique involving scanning electron microscopy and energy-dispersive X-ray spectroscopy to distinguish these minerals. Many researchers have observed the aggregation of flotation products using SEM-EDS analysis [15,16,26,27]. Derjaguin–Landau–Verwey–Overbeek theory [28,29,30] (E-DLVO theory), which considers hydrophobic and hydrophilic interactions, can be applied to calculate the interaction energies between mineral particles. It can explain the dispersion and aggregation between mineral particles and the flotation performance of minerals [31,32,33]. The above research methods have been widely used to study particle interactions in the flotation process, such as those for coal [34,35,36], hematite [37,38], magnesite [15,16], and phosphate rock [39,40]. However, these methods have not been used in the flotation system of fine forsterite and ilmenite.
In this study, the aggregation/dispersion behavior of forsterite and ilmenite of different particle sizes was studied, and the interaction mechanism was explored by calculating the interaction potential energy profiles. Based on the contact angle and zeta potential data, E-DLVO theory was applied to calculate the interaction energy between mineral particles before and after collector treatment. Particle size and SEM analyses were used to observe the particle aggregation directly. In our previous work [23], the influence of different particle sizes of forsterite on ilmenite flotation recovery was investigated, explaining the mechanism in terms of agent adsorption. In this paper, the intrinsic mechanism is explained by means of the particle interaction microscopic force.

2. Materials and Methods

2.1. Materials

After the repeated application of magnetic separation and gravity separation to vanadium–titanium magnetite ore from Panzhihua (Sichuan province, China), purified ilmenite and forsterite samples were obtained. These purified samples were subjected to grinding and wet screening to obtain different particle sizes (−151 + 74, −45 + 25, −25 + 19, −19 μm). Then, these samples were washed three times using deionized (DI) water to remove impurity ions such as Ca2+ and Mg2+, produced during the preparation process, and dried at room temperature. Table 1 and Figure 1 show the multi-element chemical analysis and X-ray diffraction (XRD) characterization results, respectively, for the purified samples. The theoretical TiO2 content in ilmenite is 52.7%, indicating a purity of 96.09% for the ilmenite sample. Additionally, the theoretical (Fe+Mg)O content in forsterite is 57.2%, indicating a purity of 93.96% for the forsterite sample.
Sodium oleate (NaOL, analytical grade) obtained from ChengDu Chron Chemicals Co, Ltd. (Chengdu, China), was used as an anionic collector in ilmenite flotation. Sulfuric acid and sodium hydroxide (H2SO4, NaOH, analytical grade), obtained from ChengDu Chron Chemicals Co, Ltd. (Chengdu, China), were used to regulate the pH. Deionized (DI) water with a resistivity of 18.25 MΩ·cm was used in the related experiments.

2.2. Methods

2.2.1. Micro-Flotation Experiments

Micro-flotation experiments with pure and artificial mixed minerals were performed using an XFGCII flotation machine (Jitan Machinery Co., Ltd., Changchun, China). Flotation was operated at an airflow rate of 0 L/min. The speed of the impeller was kept constant at 1602 rpm. In the purified mineral micro-flotation experiments, 2 g of purified mineral was suspended in 40 mL of DI water. Before the addition of the reagents, the slurry was conditioned for 1 min. The prepared H2SO4 solution or NaOH solution was added to adjust the pH to different levels (ranging from 2 to 12) with 3 min of agitation. Subsequently, a NaOL solution was added to form suspensions with varying NaOL concentrations (0, 1, 2, 3, 4, 5 × 10−4 M), followed by an additional 3 min of stirring. The froth product was then collected by scraping every 5 s for 5 min. Finally, the separated froth product and tailing were filtered and dried in a drying oven at 60 °C.
In the artificial mixed mineral experiments, 2 g samples of purified ilmenite and forsterite were mixed in different ratios (ilmenite/forsterite = 9:1, 8:2, 7:3, and 5:5). After the same operation following the purified mineral micro-flotation experiments, the froth product was assessed by means of chemical analysis. Each experiment was operated three times under the same conditions, and the reported result was the average of standard deviations in three experiments. The standard deviation for each test was calculated and presented as the error bar to illustrate the deviation of the three flotation results in the test relative to the average value.

2.2.2. Zeta Potential Measurements

The zeta potential values of the minerals were measured using a Nano-ZS90 Micro-Electrophoresis Apparatus (Malvern, UK). In each measurement, 30 mg of powdered single-mineral sample (−19 μm) and 40 mL of DI water were mixed in a beaker. A NaOL solution was introduced to achieve a NaOL concentration of 3 × 10−4 M in the suspension, followed by 10 min of stirring, and the pH was adjusted to different levels (ranging from 2 to 12). The suspension was allowed to settle for an additional 10 min, after which the zeta potential values of the particles in the liquid supernatant were measured. The reported result was the average of standard deviations in three repeated tests. The standard deviation for each test was calculated and presented as an error bar to describe the deviation of the three flotation results in the test relative to the average value.

2.2.3. Contact Angle Measurements

A Surface Analyzer LSA 100 LSA100 (Lauda, German) was used for the contact angle measurements. The single-mineral powdered samples (−19 μm) treated with NaOL solution (3 × 10−4 M) were filtered, dried, and tableted with a pressure of 10 MPa. Then, the liquid drops were dropped onto the tablet using a microsyringe, with a droplet volume of 4 μL. In each test, a treated sample was used to capture the contact process image of the liquid drop. At least three different locations of samples were measured, and the average value of the contact angle was reported. The error range of the average was controlled within ±2 degrees to ensure the validity of the data.

2.2.4. Particle Size Measurements

The particle size distribution of the minerals was determined using an LS 13 320 Laser Diffraction Particle Size Analyzer (Beckman Coulter, Indianapolis, Indiana, USA). Forsterite with different particle sizes (−45 + 25, −25 + 19, −19 μm) and coarse ilmenite (−151 + 74 μm) were mixed at a mass ratio of 1:1. After the same procedures were performed as for the micro-flotation experiments, the samples were directly taken from in the mixed slurry for particle size measurement.

2.2.5. SEM and EDS Analysis

A Sigma300 (Zessie, Oberkochen, Germany) scanning electron microscope (SEM) was used to observe the aggregation of the fine forsterite on the coarse ilmenite particles. An XFlash 6160 (Bruker, Berlin, Germany) energy-dispersive X-ray spectroscope (EDS) was used to analyze the elementary composition to distinguish the mineral particles in the SEM images.

2.2.6. DLVO Theoretical

The classical DLVO theory was developed by Derjaguin and Landau in Russia in 1941 and Verwey and Overbeek in 1948 [41], respectively. It is based on the sum ( V T D ) of the electrostatic energy ( V E ) and van der Waals energy ( V W ) between mineral particles to predict the aggregation/dispersion behavior.
V T D = V W + V E
However, flotation reagents can make the surface of mineral particles more hydrophilic or hydrophobic, resulting in polar interfacial interaction between mineral particles. To explain this phenomenon, the extended DLVO theory has been put forward. Compared with the classical DLVO theory, the extended DLVO theory considers the hydration or hydrophobic interactions ( V H ) that play a decisive role in the aggregation/dispersion behavior between mineral particles [42]:
V T ED = V W + V E + V H
(a) Van der Waals interaction, V W [43,44]
V W = A 132 6 H R 1 R 2 R 1 + R 2
Here,
A 132 = A 11 A 33 A 22 A 33
(b) Electrical interaction, V E [45,46]
For the same particle surface charge ball–ball models:
V E = 2 ε 0 ε r R Ψ 2 ln 1 + exp κ H
For ball–flat models:
V E = ε 0 ε r R 1 R 2 R 1 + R 2 k T Z e 2 Ψ 1 2 + Ψ 2 2 2 φ 1 φ 2 φ 1 2 + φ 2 2 p + q
where
p = ln 1 + exp κ H 1 exp κ H
q = ln 1 exp 2 κ H
κ = 2 e 2 n 0 Z 2 ε 0 ε r k T 1 / 2 = 2 e 2 N A C Z 2 ε 0 ε r k T 1 / 2
(c) Polar interfacial interaction, V H [47,48]
V H = 2 R 1 R 2 R 1 + R 2 h 0 V H 0 exp H 0 H h 0
In order to calculate the hydrophobic force V H , the “decay length” h 0 must be determined. According to the general literature, h 0 = 1~10 nm; in this study, the h 0 value was assumed to be 2 nm. H 0 is the minimum equilibrium distance; according to the literature, H 0 = 0.1677 nm in a water medium [49]. V H 0 is the polar interface interaction energy constant, which can be calculated using Equation (11):
V H 0 = 2 γ 3 + γ 1 + γ 2 γ 3 + γ 3 γ 1 + + γ 2 + γ 3 + γ 1 + γ 2 γ 1 γ 2 +
where γ + is for the electron-acceptor parameter and γ is for the electron-donor parameter of the surface energy of the substance.
According to Van Oss et al. [50,51,52,53,54], the contact angle is related to the polar (Lewis acid-base, γ + or γ ) and apolar (Lifshitz–van der Waals, γ L W ) components of the surface energy of solids, as well as the solid–liquid interfacial energy. They derived Equation (12):
1 + cos θ γ L + γ L γ L + = 2 γ S L W γ L L W + γ S + γ L + γ S γ L +
Most oxidized ores can be considered as monopolar surfaces, for which γ S + 0 ; Equation (13) can be simplified as follows:
1 + cos θ γ L + γ L γ L + = 2 γ S L W γ L L W + γ S γ L +
Through Equation (13), we only need to know the contact angles of the two polar liquids; then, we can calculate the values of γ S L W and γ S of the different minerals. By substituting them into Equation (11), the polar interfacial interaction energy constants V H 0 can be calculated. In the end, the polar interfacial interaction V H can be calculated using Equation (10).
In addition, the Hamaker constant and γ S L W of minerals have a corresponding relationship. γ S L W can be calculated using the polar interfacial interaction V H , and V H can be calculated from γ S L W of the minerals using Equation (14) [55]:
A = 1.51 × 10 21 γ S L W

3. Results and Discussion

3.1. Micro-Flotation Experiments

Figure 2a shows the results of the effects of different NaOL concentrations on the recovery of ilmenite and forsterite at a pH of 7.0 ± 0.5. Figure 2a illustrates that the recovery of coarse ilmenite (−151 + 74 μm) and forsterite at different sizes reached the maximum with a NaOL concentration of 3 × 10−4 M. Even at a high concentration of NaOL, the recovery did not increase significantly. This trend in flotation recovery is consistent with our previous work report [23].
Figure 2b shows the results of the effect of pH on the recovery of ilmenite and forsterite when the concentration of NaOL was 2 × 10−4 M. It shows that the floatability of coarse ilmenite (−151 + 74 μm) and different-sized forsterite had different variations within a wide pH range. The coarse ilmenite recovery first increased and then decreased with the increase in pH, and the maximum recovery was 86.35% at pH 10. In a highly alkaline environment, the chemical interaction between NaOL and Fe3⁺ weakened, while the interaction with Ca2⁺ and Mg2⁺ strengthened [56]. As a result, the ilmenite recovery decreased rapidly, and the forsterite recovery increased at pH 12.
Considering the results of the micro-flotation experiments and cost efficiency, the experiment of the artificial mixed mineral was carried out under the conditions of pH = 7.0 ± 0.5 and NaOL = 3 × 10−4 M. The experimental results are shown in Figure 3. The results show that the ilmenite recovery remains unchanged with the increase in the forsterite (−45 + 25, −25 + 19 μm) content. This indicates that conventional-sized forsterite (−45 + 25, −25 + 19 μm) had no effect on the floatability of coarse ilmenite. The reason for the decrease in ilmenite recovery was that conventional-sized forsterite (−45 + 25, −25 + 19 μm) has good floatability and entered the concentrate. However, with an increase in the fine forsterite (−19 μm) content, the recovery of coarse ilmenite decreased gradually. Aggregation between fine forsterite and coarse ilmenite may have occurred, resulting in lower ilmenite recovery.

3.2. Zeta Potential Measurements

The zeta potential is an important parameter for calculating the electrostatic interactions of particles in solution; it also can indicate the degree of adsorption between mineral particles and flotation agents [57,58]. Figure 4 shows the pH effect on the zeta potential of ilmenite and forsterite in the presence and absence of NaOL (3 × 10−4 M). The results show that the pHzpc values of ilmenite and forsterite are approximately 4.1 and 2.8, respectively. The results are similar to those reported in [59]. After mixing with NaOL, the zeta potential of both of them shifted towards negative, and the degree of the negative shift was similar. This indicates that there was a slight difference in the adsorption amount of NaOL between them, and it was difficult to separate them with flotation. In addition, ilmenite and forsterite are negatively charged in the pH range of 2–12, indicating that there is electrostatic repulsion between them.

3.3. Contact Angle Measurements

The wettability of a solid surface is usually characterized by the contact angle; thus, the value of the contact angle of a mineral can directly reflect its floatability [60,61]. Table 2 presents the contact angles of the samples with different polar liquids in the presence and absence of NaOL (3 × 10−4 M). The results showed that the contact angle of ilmenite was 30.22° and that of forsterite was 34.72° in DI water, indicating that there was no natural floatability. After NaOL treatment, the contact angle of ilmenite and forsterite increased significantly, and there was slight difference in the extent of the contact angle increase between ilmenite and forsterite. This suggests that there is little difference in the degree of NaOL adsorption, which is consistent with the zeta potential measurements.
According to the contact angle data in the two polar liquids and the known surface energy parameters of water and glycerin, the components of surface energy parameters and the Hamaker constants of the minerals can be calculated using Equations (13) and (14). The calculated results for A, γ S L W , and γ S are shown in Table 3.

3.4. Particle Size Measurements

Variations in particle sizes provide compelling evidence when assessing changes in the agglomeration or dispersion state of particles in a solution. Figure 5 shows the experimental results of the influence of fine forsterite on the particle size of mixed minerals. When conventional-sized forsterite particles (−45 + 25, −25 + 19 μm) were added, the volume-average particle size of coarse ilmenite remained unchanged, at approximately 103.68 μm. However, upon the addition of fine forsterite (−19 μm), the volume-average particle size increased to 108.86 μm, with an increase in the proportion of particles exceeding 74 μm. The data suggest that the addition of conventional-sized forsterite particles has no impact on the particle size of coarse ilmenite. However, the addition of fine forsterite (−19 μm) led to an increase in the volume-average size of coarse ilmenite particles. This indicates that fine forsterite can adsorb onto the surface of coarse ilmenite, resulting in the enlargement of ilmenite particles, thereby affecting the flotation recovery of ilmenite. These findings are consistent with the results of artificial mixed mineral flotation tests.

3.5. SEM and EDS Analysis

SEM images of flotation products can used to observe the dispersion/aggregation state between particles [62]. Figure 6a–c show SEM images of flotation tailings with coarse ilmenite particles and −45 + 25, −25 + 19, and −19 μm forsterite particles, respectively. In addition, Figure 6d shows a partial enlargement image of (c). Figure 6a,b show that the conventional-sized forsterite particles (−45 + 25, −25 + 19 μm) and coarse ilmenite particles were dispersed, and the surface of the ilmenite particles was clean. However, Figure 6c,d show that aggregation occurred not only between fine forsterite (−19 μm) and coarse ilmenite but also between fine forsterite itself. The ilmenite surface adhered to fine forsterite, and there were also aggregates of fine forsterite present. This is consistent with the results of the flotation experiment and the particle size measurement.

3.6. DLVO Theoretical Calculation

With the zeta potential and mineral surface energy parameters, the interaction potential energy profiles between forsterite and ilmenite particles can be calculated according to Equations (2)–(11); the calculation results are shown in Figure 7. The volume-average radius of ilmenite in the −151 + 74 μm range was 56.18 μm, while the volume-average radius of −19 μm forsterite was 4.98 μm. Additionally, the volume-average radii of the −45 + 25 and −25 + 19 μm forsterite were 17.5 μm and 11 μm, respectively. The Hamaker constant for water (A33) is 3.7 × 10−20 J, while the values for ilmenite and forsterite are presented in Table 3. Utilizing Equation (4), A132 was calculated to be 4.20 × 10−20 J. The van der Waals interaction energy, Vw, could subsequently be determined. The electrolyte concentration in the flotation process was uniformly 10−3 M; then, κ = 0.104   nm 1 was calculated from Equation (10). According to Figure 5, under the condition of pH 7, the zeta potentials of ilmenite and forsterite were −25.62 mV and −28.44 mV without NaOL and −58.46 mV and −49.39 mV with NaOL (3 × 10−4 M), respectively. The electrostatic interaction energy, VE, could be calculated based on these values. Using the data from Table 3 and the surface energy parameters of water ( γ S = γ S + = 25.5 mJ·m−2), the polar interface interaction energy constant, V H 0 , was computed. VH was also calculated. The total interaction potential energy, V T ED , was obtained by summing Vw, VE, and VH.
Figure 7a–c show the interaction potential energy profiles of DLVO and E-DLVO between coarse ilmenite (−151 + 74 μm) and forsterite particles of various particle sizes. At pH 7 and a NaOL concentration of 3 × 10−4 M, an energy barrier was observed in the interaction potential energy between coarse ilmenite (−151 + 74 μm) and forsterite particles of sizes −45 + 25 μm, −25 + 19 μm, and −19 μm, with energy barrier values of 11.94 × 103 kT, 8.23 × 103 kT, and 4.09 × 103 kT, respectively. The strength of the barrier diminished as the size of forsterite decreased, suggesting that fine forsterite (−19 μm) might more easily overcome the barrier and aggregate with coarse ilmenite during flotation agitation. Thus, the content of conventional-sized forsterite (−45 + 25, −25 + 19 μm) has no effect on the floatability of ilmenite, whereas the fine forsterite (−19 μm) will gradually affect the floatability of coarse ilmenite with the increasing content.
Figure 7d presents the interaction potential energy profiles of DLVO and E-DLVO for fine forsterite (−19 μm). The energy barrier between fine forsterite was calculated at 0.51 × 103 kT, which is lower compared to the barriers calculated between ilmenite and forsterite. This indicates that fine forsterite tends to aggregate more readily, another significant factor contributing to the lower recovery of ilmenite. Fine forsterite will aggregate with each other to form agglomerates of fine forsterite particles. This agglomerate still retains stronger micro-forces between the fine forsterite and ilmenite compared to the conventional-sized forsterite (−45 + 25, −25 + 19 μm), because the surface of the agglomerate is still ultrafine forsterite. As a result, these agglomerates adhere to ilmenite particles, leading to more forsterite entering the concentrate and resulting in reduced ilmenite recovery.

4. Conclusions

In this study, according to the significant changes in the properties of magmatic ilmenite ore, forsterite with increasing content in the ore was selected as the gangue mineral, and a mineral flotation experiment was carried out. The interaction phenomenon between forsterite and ilmenite particles with different particle sizes was discussed, and the underlying interaction mechanism was explained by applying E-DLVO theory. The results show the following:
  • The content of conventional-sized forsterite (−45 + 25, −25 + 19 μm) has little effect on the floatability of ilmenite in a neutral environment. However, the ilmenite recovery decreases with an increase in the fine forsterite (−19 μm) content. In the mining industry, where the content of forsterite is gradually increasing, particular attention should be paid to variations in the content of fine forsterite (−19 μm) during the flotation production process.
  • According to particle size measurements, scanning electron microscopy (SEM), and E-DLVO theoretical calculations, fine forsterite (−19 μm) directly covers the surface of coarse ilmenite (−151 + 74 μm) particles, preventing the ilmenite from being collected and reducing its floatability. On the other hand, aggregates of fine forsterite (−19 μm) enter the ilmenite concentrate together with the ilmenite, resulting in lower ilmenite recovery.
Further research should be conducted to explore the correlations between various particle sizes and contact angles, as well as to reduce the influence of fine particles on the flotation process, considering the potential limitations highlighted in this study.

Author Contributions

Conceptualization, S.Z.; data curation, S.Z., W.Y. and W.L.; funding acquisition, Y.Y.; investigation, W.Y. and W.L.; methodology, S.Z. and Y.Y.; resources, Y.Y.; supervision, D.W.; writing—original draft, S.Z.; writing—review and editing, D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Open Foundation of Technology Innovation Center for Comprehensive Utilization of Strategic Mineral Resources, Ministry of Natural Resources (Grant No. CCUM-KY-2307); the National Key Research and Development Program of China (Grant No. 2021YFC2900800); and the China Geological Survey, Ministry of Natural Resources (Grant No. DD20230039).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

V T D total interaction energy by DLVO theory, J
V T ED total interaction energy by E-DLVO theory, J
V W interaction energy due to van der Waals forces, J
V E interaction energy due to electrical double layer effects, J
V H interaction energy due to hydration/hydrophobic effects, J
A Hamaker constant, J
A 132 effective Hamaker constant of materials 1 and 2 in medium 3, J
R 1 , R 2 radius of particles 1 and 2, m
ε 0 permittivity of free space, 8.854 × 10−12 F·m−1
ε r relative permittivity, for water ε r = 81
Ψ dimensionless surface potential, Ψ = Z e φ k T
φ 1 , φ 2 zeta potential of particles 1 and 2, V
κ Debye–Hückel parameter, m−1
H minimum separation distance between two spheres, m
e elementary charge, 1.602 × 10−19 C
N A Avogadro number, 6.023 × 1023 mol−1
C concentration, mol·m−3
Z charge number
k Boltzmann constant, 1.38 × 10−23 J·K−1
T temperature, K
γ i parameter of polar component of the surface tension of compound i, donating electron or accepting proton
γ i + parameter of polar component of the surface tension of compound i, donating proton or accepting electron
γ i L W parameter of a polar (Lifshitz–van der Waals) component of the surface tension of compound i

References

  1. Fang, S.; Xu, L.; Wu, H.; Tian, J.; Lu, Z.; Sun, W.; Hu, Y. Adsorption of Pb(II)/benzohydroxamic acid collector complexes for ilmenite flotation. Miner. Eng. 2018, 126, 16–23. [Google Scholar] [CrossRef]
  2. Vásquez, R.; Molina, A. Effects of thermal preoxidation on reductive leaching of ilmenite. Miner. Eng. 2012, 39, 99–105. [Google Scholar] [CrossRef]
  3. Meng, Q.; Yuan, Z.; Yu, L.; Xu, Y.; Du, Y.; Zhang, C. Selective depression of titanaugite in the ilmenite flotation with carboxymethyl starch. Appl. Surf. Sci. 2018, 440, 955–962. [Google Scholar] [CrossRef]
  4. USGS. Mineral commodity summaries 2023. US Geological Survey; USGS: Reston, VA, USA, 2023; 210p. [Google Scholar] [CrossRef]
  5. Zhao, J.H.; Zhou, M.F. Geochemistry of Neoproterozoic mafic intrusions in the Panzhihua district (Sichuan Province, SW China): Implications for subduction-related metasomatism in the upper mantle. Precambrian Res. 2007, 152, 27–47. [Google Scholar] [CrossRef]
  6. Zhou, M.F.; Chen, W.T.; Wang, C.Y.; Prevec, S.A.; Liu, P.P.; Howarth, G.H. Two stages of immiscible liquid separation in the formation of Panzhihua-type Fe-Ti-V oxide deposits, SW China. Geosci. Front. 2013, 4, 481–502. [Google Scholar] [CrossRef]
  7. Yang, Y.; Hui, B.; Yan, W.; He, B.; Wang, W. Research on process mineralogy of fine ilmenite in Panxi area. Multipurp. Util. Miner. Resour. 2020, 3, 131–135. [Google Scholar] [CrossRef]
  8. Hui, B.; Yang, Y.H. Properties of olive-pyroxene vanadium-titanium magnetite ore in Hongge mining area of Panxi research and influence on mineral processing technology. Multipurp. Util. Miner. Resour. 2020, 4, 126–129. [Google Scholar] [CrossRef]
  9. Gomez-Flores, A.; Solongo, S.K.; Heyes, G.W.; Ilyas, S.; Kim, H. Bubble−particle interactions with hydrodynamics, XDLVO theory, and surface roughness for flotation in an agitated tank using CFD simulations. Miner. Eng. 2020, 152, 106368. [Google Scholar] [CrossRef]
  10. Nikolaev, A.A.; Batkhuyag, A.; Goryachev, B.E. Mineralization Kinetics of Air Bubble in Pyrite Slurry under Dynamic Conditions. J. Min. Sci. 2018, 54, 840–844. [Google Scholar] [CrossRef]
  11. Goryachev, B.E.; Nikolaev, A.A.; Ils’ina, E.Y. Analysis of flotation kinetics of particles with the controllable hydrophobic behavior. J. Min. Sci. 2010, 46, 72–77. [Google Scholar] [CrossRef]
  12. Nikolaev, A.A. Flotation Recovery of Sphalerite in Sea Water: A Feasibility Study. Resources 2023, 12, 51. [Google Scholar] [CrossRef]
  13. Jameson, G.J. The effect of surface liberation and particle size on flotation rate constants. Miner. Eng. 2012, 36–38, 132–137. [Google Scholar] [CrossRef]
  14. Miettinen, T.; Ralston, J.; Fornasiero, D. The limits of fine particle flotation. Miner. Eng. 2010, 23, 420–437. [Google Scholar] [CrossRef]
  15. Yao, J.; Xue, J.; Li, D.; Fu, Y.; Gong, E.; Yin, W. Effects of fine–coarse particles interaction on flotation separation and interaction energy calculation. Part. Sci. Technol. 2018, 36, 11–19. [Google Scholar] [CrossRef]
  16. Yao, J.; Yin, W.; Gong, E. Depressing effect of fine hydrophilic particles on magnesite reverse flotation. Int. J. Miner. Process. 2016, 149, 84–93. [Google Scholar] [CrossRef]
  17. Forbes, E.; Davey, K.J.; Smith, L. Decoupling rehology and slime coatings effect on the natural flotability of chalcopyrite in a clay-rich flotation pulp. Miner. Eng. 2014, 56, 136–144. [Google Scholar] [CrossRef]
  18. Yu, Y.; Ma, L.; Cao, M.; Liu, Q. Slime coatings in froth flotation: A review. Miner. Eng. 2017, 114, 26–36. [Google Scholar] [CrossRef]
  19. Eckert, K.; Schach, E.; Gerbeth, G.; Rudolph, M. Carrier Flotation: State of the Art and its Potential for the Separation of Fine and Ultrafine Mineral Particles. Mater. Sci. Forum 2019, 959, 125–133. [Google Scholar] [CrossRef]
  20. Zhang, X.; Hu, Y.; Sun, W.; Xu, L. The Effect of Polystyrene on the Carrier Flotation of Fine Smithsonite. Minerals 2017, 7, 52. [Google Scholar] [CrossRef]
  21. Yang, H.; Tang, Q.; Wang, C.; Zhang, J. Flocculation and flotation response of Rhodococcus erythropolis to pure minerals in hematite ores. Miner. Eng. 2013, 45, 67–72. [Google Scholar] [CrossRef]
  22. Zhu, Y.G.; Zhang, G.F.; Feng, Q.M.; Lu, Y.P.; Ou, L.M. Autogenous-carrier flotation of fine ilmenite. Chin. J. Nonferrous Met. 2009, 19, 554–560. [Google Scholar] [CrossRef]
  23. Yang, Y.; Xu, L.; Liu, S.; Deng, J. Influence of particle size on flotation separation of ilmenite, olivine, and pyroxene. Physicochem. Probl. Miner. Process. 2021, 57, 106–117. [Google Scholar] [CrossRef]
  24. Fan, G.; Cao, Y. The shear flocculation flotation behaviors of micro fine ilmenite and titanaugite. J. China Univ. Min. Technol. 2015, 44, 532–539. [Google Scholar] [CrossRef]
  25. Lange, A.G.; Skinner, W.M.; Smart, R.S.C. Fine: Coarse particle interactions and aggregation in sphalerite flotation. Miner. Eng. 1997, 10, 681–693. [Google Scholar] [CrossRef]
  26. Oats, W.J.; Ozdemir, O.; Nguyen, A.V. Effect of mechanical and chemical clay removals by hydrocyclone and dispersants on coal flotation. Miner. Eng. 2010, 23, 413–419. [Google Scholar] [CrossRef]
  27. Peng, Y.; Bradshaw, D. Mechanisms for the improved flotation of ultrafine pentlandite and its separation from lizardite in saline water. Miner. Eng. 2012, 36–38, 284–290. [Google Scholar] [CrossRef]
  28. Yan, X.; Wei, L.; Meng, Q.; Wang, J.; Yang, Q.; Zhai, S.; Lu, J. A study on the mechanism of calcium ion in promoting the sedimentation of illite particles. J. Water Process Eng. 2021, 42, 102153. [Google Scholar] [CrossRef]
  29. Chen, Y.; Hu, S.; Li, J.; Weng, L.; Wu, C.; Liu, K. Improvement on combustible matter recovery in coal slime flotation with the addition of sodium silicate. Colloids Surf. A 2020, 603, 125220. [Google Scholar] [CrossRef]
  30. Liao, Y.; Hao, X.; An, M.; Yang, Z.; Ma, L.; Ren, H. Enhancing low-rank coal flotation using mixed collector of dodecane and oleic acid: Effect of droplet dispersion and its interaction with coal particle. Fuel 2020, 280, 118634. [Google Scholar] [CrossRef]
  31. Hu, Y.; Dai, J. Hydrophobic aggregation of alumina in surfactant solution. Miner. Eng. 2003, 16, 1167–1172. [Google Scholar] [CrossRef]
  32. Hartmann, R.; Kinnunen, P.; Illikainen, M. Cellulose-mineral interactions based on the DLVO theory and their correlation with flotability. Miner. Eng. 2018, 122, 44–52. [Google Scholar] [CrossRef]
  33. Farahat, M.; Hirajima, T.; Sasaki, K.; Doi, K. Adhesion of Escherichia coli onto quartz, hematite and corundum: Extended DLVO theory and flotation behavior. Colloids Surf. B Biointerfaces 2009, 74, 140–149. [Google Scholar] [CrossRef] [PubMed]
  34. Ni, C.; Bu, X.; Xia, W.; Peng, Y.; Yu, H.; Xie, G. Observing slime-coating of fine minerals on the lump coal surface using particle vision and measurement. Powder Technol. 2018, 339, 434–439. [Google Scholar] [CrossRef]
  35. Wang, X.; Zhou, S.; Bu, X.; Ni, C.; Xie, G.; Peng, Y. Investigation on interaction behavior between coarse and fine particles in the coal flotation using focused beam reflectance measurement (FBRM) and particle video microscope (PVM). Sep. Sci. Technol. 2021, 56, 1418–1430. [Google Scholar] [CrossRef]
  36. Li, Q.; Zhang, W.; Tan, J.; Liang, L.; Xie, G. Mechanism of slime coating on coal surface with different metamorphic degrees. Colloids Surf. A 2023, 675, 132038. [Google Scholar] [CrossRef]
  37. Li, D.; Yin, W.; Liu, Q.; Cao, S.; Sun, Q.; Zhao, C.; Yao, J. Interactions between fine and coarse hematite particles in aqueous suspension and their implications for flotation. Miner. Eng. 2017, 114, 74–81. [Google Scholar] [CrossRef]
  38. Han, H.; Yin, W.; Wang, D.; Zhu, Z.; Yang, B.; Yao, J. New insights into the dispersion mechanism of citric acid for enhancing the flotation separation of fine siderite from hematite and quartz. Colloids Surf. A 2022, 641, 128459. [Google Scholar] [CrossRef]
  39. Zhang, H.; Zhou, F.; Yu, H.; Liu, M. Double roles of sodium hexametaphosphate in the flotation of dolomite from apatite. Colloids Surf. A 2021, 626, 127080. [Google Scholar] [CrossRef]
  40. Huang, X.; Zhang, Q. Interaction Behavior between Coarse and Fine Particles in the Reverse Flotation of Fluorapatite and Dolomite. Langmuir 2023, 39, 12931–12943. [Google Scholar] [CrossRef]
  41. Vincent, B. Early (pre-DLVO) studies of particle aggregation. Adv. Colloid Interface Sci. 2012, 170, 56–67. [Google Scholar] [CrossRef]
  42. Lyklema, J. Fundamentals of Interface and Colloid Science: Soft Colloids; Elsevier: Amsterdam, The Netherlands, 2005; Volume 5. [Google Scholar]
  43. Verwey, E.J.W.; Overbeek, J.T.G. Theory of the stability of lyophobic colloids. J. Colloid Sci. 1955, 10, 224–225. [Google Scholar] [CrossRef]
  44. Schenkel, J.H.; Kitchener, J.A. A test of the Derjaguin-Verwey-Overbeek theory with a colloidal suspension. Trans. Faraday Soc. 1960, 56, 161–173. [Google Scholar] [CrossRef]
  45. Hogg, R.; Healy, T.W.; Fuerstenau, D.W. Mutual coagulation of colloidal dispersions. Trans. Faraday Soc. 1966, 62, 1638–1651. [Google Scholar] [CrossRef]
  46. Pugh, R.J.; Kitchener, J.A. Theory of selective coagulation in mixed colloidal suspensions. J. Colloid Interface Sci. 1971, 35, 656–664. [Google Scholar] [CrossRef]
  47. Israelachvili, J.; Pashley, R. The hydrophobic interaction is long range, decaying exponentially with distance. Nature 1982, 300, 341–342. [Google Scholar] [CrossRef]
  48. Claesson, P.M.; Blom, C.E.; Herder, P.C.; Ninham, B.W. Interactions between water—Stable hydrophobic Langmuir—Blodgett monolayers on mica. J. Colloid Interface Sci. 1986, 114, 234–242. [Google Scholar] [CrossRef]
  49. Van Oss, C.J. Interfacial Forces in Aqueous Media; CRC Press: Boca Raton, FL, USA, 2006. [Google Scholar]
  50. van Oss, C.J.; Chaudhury, M.K.; Good, R.J. Monopolar surfaces. Adv. Colloid Interface Sci. 1987, 28, 35–64. [Google Scholar] [CrossRef]
  51. Van Oss, C.J.; Good, R.J.; Chaudhury, M.K. Determination of the Hydrophobia Interaction Energy-Application to Separation Processes. Sep. Sci. Technol. 1987, 22, 1–24. [Google Scholar] [CrossRef]
  52. van Oss, C.J.; Chaudhury, M.K.; Good, R.J. Interfacial Lifshitz—Van der Waals and Polar Interactions in Macroscopic Systems. Chem. Rev. 1988, 88, 927–941. [Google Scholar] [CrossRef]
  53. Van Oss, C.J.V.; Good, R.J. Surface Tension and the Solubility of Polymers and Biopolymers: The Role of Polar and Apolar Interfacial Free Energies. J. Macromol. Sci. Chem. 1989, 26, 1183–1203. [Google Scholar] [CrossRef]
  54. van Oss, C.J.; Giese, R.F.; Costanzo, P.M. DLVO and Non-DLVO Interactions in Hectorite. Clays Clay Miner. 1990, 38, 151–159. [Google Scholar] [CrossRef]
  55. Bergström, L. Hamaker constants of inorganic materials. Adv. Colloid Interface Sci. 1997, 70, 125–169. [Google Scholar] [CrossRef]
  56. Semsari Parapari, P.; Irannajad, M.; Mehdilo, A. Effect of acid surface dissolution pretreatment on the selective flotation of ilmenite from olivine and pyroxene. Int. J. Miner. Process. 2017, 167, 49–60. [Google Scholar] [CrossRef]
  57. Ma, Z.; Shi, X.; Xu, L.; Wang, D.; Xue, K.; Jing, L.; Meng, J. Selective flotation separation of spodumene from feldspar using a novel mixed anionic/cationic collector NaOL/ND13. Miner. Eng. 2023, 201, 108152. [Google Scholar] [CrossRef]
  58. Wang, Y.; Wang, D.; Xu, L.; Xue, K.; Zhang, X.; Shi, X.; Liu, C.; Meng, J. Synthesis and utilization of a novel oleate hydroxamic acid collector for the flotation separation of bastnaesite from barite. Miner. Eng. 2023, 204, 108405. [Google Scholar] [CrossRef]
  59. Liu, W.; Zhang, J.; Wang, W.; Deng, J.; Chen, B.; Yan, W.; Xiong, S.; Huang, Y.; Liu, J. Flotation behaviors of ilmenite, titanaugite, and forsterite using sodium oleate as the collector. Miner. Eng. 2015, 72, 1–9. [Google Scholar] [CrossRef]
  60. Wang, D.; Zhu, Z.; Yang, B.; Yin, W.; Drelich, J.W. Nano-scaled roughness effect on air bubble-hydrophilic surface adhesive strength. Colloids Surf. A 2020, 603, 125228. [Google Scholar] [CrossRef]
  61. Wang, D.; Jiang, Y.; Zhu, Z.; Yin, W.; Asawa, K.; Choi, C.H.; Drelich, J.W. Contact Line and Adhesion Force of Droplets on Concentric Ring-Textured Hydrophobic Surfaces. Langmuir 2020, 36, 2622–2628. [Google Scholar] [CrossRef]
  62. Yin, W.; Wang, D.; Drelich, J.W.; Yang, B.; Li, D.; Zhu, Z.; Yao, J. Reverse flotation separation of hematite from quartz assisted with magnetic seeding aggregation. Miner. Eng. 2019, 139, 105873. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of samples: (a) ilmenite; (b) forsterite.
Figure 1. XRD spectra of samples: (a) ilmenite; (b) forsterite.
Minerals 14 01041 g001
Figure 2. Effect of solution chemical environment on the floatability of ilmenite and forsterite: (a) NaOL concentration; (b) pH.
Figure 2. Effect of solution chemical environment on the floatability of ilmenite and forsterite: (a) NaOL concentration; (b) pH.
Minerals 14 01041 g002
Figure 3. Effect of different forsterite contents on ilmenite floatability.
Figure 3. Effect of different forsterite contents on ilmenite floatability.
Minerals 14 01041 g003
Figure 4. Effect of pH on zeta potential of ilmenite and forsterite in the presence and absence of NaOL.
Figure 4. Effect of pH on zeta potential of ilmenite and forsterite in the presence and absence of NaOL.
Minerals 14 01041 g004
Figure 5. Effect of fine forsterite on particle size distribution of mixed minerals.
Figure 5. Effect of fine forsterite on particle size distribution of mixed minerals.
Minerals 14 01041 g005
Figure 6. SEM images of flotation products: (a) −151 + 74 μm ilmenite and −45 + 25 μm forsterite; (b) −151 + 74 μm ilmenite and −25 + 19 μm forsterite; (c) −151 + 74 μm ilmenite and −19 μm forsterite; (d) partial enlargement of the Figure 6c.
Figure 6. SEM images of flotation products: (a) −151 + 74 μm ilmenite and −45 + 25 μm forsterite; (b) −151 + 74 μm ilmenite and −25 + 19 μm forsterite; (c) −151 + 74 μm ilmenite and −19 μm forsterite; (d) partial enlargement of the Figure 6c.
Minerals 14 01041 g006
Figure 7. DLVO and the E-DLVO interaction energy profiles for ilmenite and forsterite in the presence and absence of NaOL: (a) −151 + 74 μm ilmenite and −45 + 25 μm forsterite; (b) −151 + 74 μm ilmenite and −25 + 19 μm forsterite; (c) −151 + 74 μm ilmenite and −19 μm forsterite; (d) −19 μm forsterite and −19 μm forsterite.
Figure 7. DLVO and the E-DLVO interaction energy profiles for ilmenite and forsterite in the presence and absence of NaOL: (a) −151 + 74 μm ilmenite and −45 + 25 μm forsterite; (b) −151 + 74 μm ilmenite and −25 + 19 μm forsterite; (c) −151 + 74 μm ilmenite and −19 μm forsterite; (d) −19 μm forsterite and −19 μm forsterite.
Minerals 14 01041 g007
Table 1. Results of multi-element chemical analysis of ilmenite and forsterite samples (mass fraction, %).
Table 1. Results of multi-element chemical analysis of ilmenite and forsterite samples (mass fraction, %).
SampleTiO2Fe2O3CaOSiO2Al2O3MgO
Ilmenite50.6446.480.190.560.304.13
Forsterite0.019.890.1340.000.1449.30
Table 2. The average contact angles with and without NaOL in deionized water and glycerin (θ°).
Table 2. The average contact angles with and without NaOL in deionized water and glycerin (θ°).
MineralsIlmeniteForsterite
Without NaOLWith NaOLWithout NaOLWith NaOL
DI water30.2288.1534.7287.63
Glycerin39.7251.9327.7046.97
Table 3. Values of components of surface energies and Hamaker constants of ilmenite and forsterite.
Table 3. Values of components of surface energies and Hamaker constants of ilmenite and forsterite.
MineralsA (10−20 J) γ S L W (mJ/m2) γ S (mJ/m2)
Without NaOLWith NaOLWithout NaOLWith NaOL
Ilmenite13.5689.80162.07198.490.780
Forsterite18.59123.14180.85145.770.109
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, S.; Yang, Y.; Wang, D.; Yan, W.; Li, W. Influence of Particle Size on Flotation Separation of Ilmenite and Forsterite. Minerals 2024, 14, 1041. https://doi.org/10.3390/min14101041

AMA Style

Zhang S, Yang Y, Wang D, Yan W, Li W. Influence of Particle Size on Flotation Separation of Ilmenite and Forsterite. Minerals. 2024; 14(10):1041. https://doi.org/10.3390/min14101041

Chicago/Turabian Style

Zhang, Senpeng, Yaohui Yang, Donghui Wang, Weiping Yan, and Weishi Li. 2024. "Influence of Particle Size on Flotation Separation of Ilmenite and Forsterite" Minerals 14, no. 10: 1041. https://doi.org/10.3390/min14101041

APA Style

Zhang, S., Yang, Y., Wang, D., Yan, W., & Li, W. (2024). Influence of Particle Size on Flotation Separation of Ilmenite and Forsterite. Minerals, 14(10), 1041. https://doi.org/10.3390/min14101041

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