Next Article in Journal
Metallic Copper (Cu[0]) Obtained from Cu2+-Rich Acidic Mine Waters by Two Different Reduction Methods: Crystallographic and Geochemical Aspects
Next Article in Special Issue
Editorial for Special Issue “Water within Minerals Processing”
Previous Article in Journal
Distribution of Heavy Metals in the Commune of Coronel, Chile
Previous Article in Special Issue
Simulating the Effect of Water Recirculation on Flotation through Ion-Spiking: Effect of Ca2+ and Mg2+
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Considering Specific Ion Effects on Froth Stability in Sulfidic Cu-Ni-PGM Ore Flotation

by
Malibongwe S. Manono
* and
Kirsten C. Corin
Centre for Minerals Research, Department of Chemical Engineering, Faculty of Engineering & the Built Environment, University of Cape Town, Rondebosch, Cape Town 7700, South Africa
*
Author to whom correspondence should be addressed.
Submission received: 26 November 2021 / Revised: 17 February 2022 / Accepted: 24 February 2022 / Published: 4 March 2022
(This article belongs to the Special Issue Water within Minerals Processing)

Abstract

:
The mining and mineral processing of Cu-Ni-PGM sulfide ores in South Africa occurs in semi-arid regions. The scarcity of water resources in these regions has become one of the biggest challenges faced by mineral concentrators. As a result, concentrators are forced to find ways through which they can manage and control their water usage. The recycling and re-use of process water in mineral concentration plants has therefore become a common practice. This practice is beneficial in that it reduces reliance on municipal water and harnesses compliance to stringent environmental regulations on freshwater usage. This approach also offers a better response to the Sustainable Development Goals (SDGs) for the mining industry, as water and its preservation form part of the SDGs. This practice could, however, be somewhat concerning to a process operator because recirculated water often has higher concentrations of ions compared to fresh or potable water. This is because an unintended change in the process water quality may affect critical aspects of flotation such as the stability of the froth. This issue has led to the need for both the mining industry and researchers in the field to find the ions in process water that have the greatest impact on froth stability. Thus, the authors of this study investigated the effects of various ions common in the process water of a typical Cu-Ni-PGM ore on froth stability using a 3 L bench scale flotation cell. Solids and water recoveries were used as proxies for froth stability. These were further complemented by bubble size, water recoveries, foam height, and dynamic foam stability from two-phase flotation systems. A two-phase foam study resulted in observations that supported findings from a three-phase study. Generally, single salt solutions containing Ca2+ and Mg2+ ions resulted in higher water recoveries both in the two-phase foam and three-phase froth studies, increases in foam heights and dynamic foam stability, and a decrease in bubble size compared to the solutions that contained Na+. SO42− also resulted in increased foam stability compared to Cl and NO3. These results showed that the divalent inorganic electrolytes—Ca2+, Mg2+, and SO42−—were more froth- and foam-stabilizing than the monovalent inorganic electrolytes—Cl, NO3, and Na+. This finding was in agreement with previous research. The findings of this study are deemed crucial in the development of a process water management protocol in sulfidic Cu-Ni-PGM ore concentrators. However, more comparative three-phase froth stability tests are needed as subjects of future investigative work to further ascertain specific ion effects on froth stability in sulfide ores.

Graphical Abstract

1. Introduction

1.1. Measures of the Stability of the Froth

Flotation, as a physicochemical separation process, consists of two distinct zones: the pulp zone where mineral recovery takes place and the froth zone where concentrated mineral particles are separated from the bulk [1]. The pulp zone, which promotes bubble–particle collision, is crucial in flotation because within the pulp, valuable minerals of a certain hydrophobicity are attached to air bubbles and are ultimately transported to the surface while gangue minerals with a measurable level of hydrophilicity are inhibited from attaching to the bubbles. These gangue minerals are not recovered to the concentrate by “true flotation”; instead, they are removed into the tailings or recovered into the concentrate by non-selective transport mechanisms, namely entrainment and entrapment. It is therefore important to limit these two mechanisms [2,3].
The froth zone is crucial in ensuring maximum froth recovery, in that there needs to be a certain level of froth stability to ensure the maximum recovery or collection of the particles recovered from the pulp zone [1,3]. The dynamics of the froth zone, however, are controlled by the flow of air bubbles from the pulp through to the froth surface and by bubble coalescence. Froth stability is therefore determined by the ability of the air bubbles to resist coalescence and bursting, so a more stable froth is characterised by a decrease in bubble coalescences [1].
Several studies have shown that froth stability plays a major role in flotation performance, but there is no single specific criterion to quantify froth stability [4,5,6,7,8,9,10,11,12,13]. There are, however, many parameters available to indicate froth stability; Farrokhpay [8] listed the following parameters for use as indicators for froth stability: froth half-life time [14,15], maximum froth equilibrium height [4,16,17], bubble growth across the froth phase [18], air recovery, amount of solids loaded onto the bubbles [17], froth velocity, and froth rise velocity [4,19]. Most of these parameters are affected by a plant’s operating conditions such as ore mineralogy, reagents chemistry, and air velocity. Therefore, any working criterion for the best froth stability and therefore froth flotation performance would need to account for major operating fluctuations in order to develop froth control strategies. Water quality has become one such fluctuating variable in most sulfides concentrators, especially in water-scarce regions. Previous research has shown that the rupturing of rising bubbles results in the falling back of solid mineral particles and liquid film drainage, so the amount of water recovered in the flotation cell concentrate launder can be used to quantify froth stability. This research led to the proposal that the larger the amount of water recovered, the more stable the froth. Similar proposals have been made by several researchers, as mentioned in Farrokhpay [8].
Multiple methods have been developed to assess foam and or froth stability [8,16]. They are generalized into dynamic and static tests, but there are questions regarding their consistency since the presence of solids affects foam stability. Foam refers to a two-phase system comprising air bubbles and liquid, while froth refers to a three-phase system consisting of air bubbles, liquid particles, and solid particles [8]. Foam is the dispersion of gas bubbles stabilised by surfactants that adsorb onto the liquid–gas interphase and are separated by thin liquid film [8,20]. Foam stability is linked to liquid drainage in the foam and the rupture of liquid films [8].
Dynamic tests are conducted for a system that is in a dynamic equilibrium between the rates of formation and of decay of foam. Static tests are conducted by allowing the foam to drop from the maximum height without regeneration by agitation or air input. The dynamic test is the most used test in flotation since it is believed to closely approximate real flotation system dynamics [4,16,17]. The dynamics in flotation are such that air is continually introduced into the pulp, generating bubbles that move from the pulp zone to the interface between the pulp and froth zones while the liquid drains downwards; accordingly, the quality of the formed froth varies with height, and the top of the froth can reach a breaking condition that leads to the collapse of the top layer [12]. One other advantage of the dynamic test is that it is reproducible, accurate and convenient to apply [16].
Bikerman [21] originally proposed a dynamic froth stability column for non-overflowing froth columns, and this technique has been used, modified and applied in many areas. One such application was on mineral flotation on a laboratory scale by Barbian et al. [16] for measuring the froth performance. With this set-up, measurements are carried out using a vertical column that is inserted through the froth to far enough below the pulp–froth interface in the flotation cell to avoid froth leakage from the bottom. The froth rises inside the column, and froth height (H) is measured as a function of time, t, until the equilibrium maximum height (Hmax), which is measured using pulp–froth interface as a reference, is achieved. The equilibrium maximum height is dependent on the froth stability and operating conditions. It is worth noting that at the equilibrium height, the rates of growth and bursting are equal. This method was well-described by Barbian et al. [4] and allows for the determination of, e.g., the dynamic foam stability factor, the froth rise rate, the froth rise velocity, and the froth retention time.
In linking froth height to froth flotation performance, industries have experienced increases in grade and recovery with increases in froth height, though recovery may also decrease with increasing froth height [4,8,16,17]. Froth height is one of the factors that can be changed to maximise flotation performance. As previously discussed, the literature suggests that there are many other correlations and parameters that can be used to quantify froth stability, and these continue to be topics of future investigative research.

1.2. Current Understanding into Electrolytes on Froth Stability

It is known that pulp chemistry in mineral flotation has important roles in not only froth stability but also mineral recoveries and grades as a result of interactions taking place at the air–water, air–mineral, and water–mineral interfaces [3,22,23,24,25,26,27,28,29,30,31,32,33]. These interactions may affect critical aspects of flotation such as the hydrophobicity or hydrophilicity of mineral particles, bubble coalescence, froth stability, and entrainment. Understanding reagent or chemical effects on any of these critical aspects of flotation can be a complex challenge because these various interfacial interactions are interrelated (e.g., any effects on the hydrophobicity of mineral particles may have serious implications on froth stability, which would also need to be studied so as to have a holistic view of how certain parameters—whether chemical, physical or operational—affect flotation performance) [34]. It is thus argued that any change effected by flotation control variables on the pulp chemistry could bring about significant step-changes to flotation performance.
In general, until recent years, the effect of ions during flotation on mineral recovery, especially on complex and polymetallic sulphide ores, was not well-documented in the literature [35,36]. However, the effect of calcium and magnesium ions on a select few sulphide minerals such as pyrite and sphalerite was reported on as early as the 1980s [37,38]. The sources of ions in flotation cells are mainly salts dissolved in the flotation water and other flotation reagents such as pH modifiers. It was reported that an apparent consequence of a high salt content is that a flotation circuit can operate without the addition of a frother [39]. This observation was in agreement with previous evidence suggesting that flotation in saline water presented in lower reagent consumption [40]. It is speculated that salts could have some similar capabilities as frothers, such decreasing bubble size and consequently increasing froth stability due to a decrease in surface tension that prevents bubbles from coalescing [39]. There are, however, significant differences between inorganic salts and frothers. High salt concentrations are required for coalescence inhibition. Previously, it was reported that inorganic ions are usually not able to form froth (foam) in a two-phase system [41]. It was reported that only in the presence of hydrophobic particles could salt solutions form a stable froth since salts are categorised as surface-inactive agents whereas frothers are surface-active agents [41]. However, recent studies have shown that inorganic ions, despite being surface-inactive agents, are easily able to form a stable foam in a two-phase system [42]. These findings agree with those of Craig et al. [43], who showed that although salts are surface-inactive surfactants, their addition to solutions containing frothers brought about great reductions in surface tension and bubble coalescence than when the frothers were added to solutions of de-ionised water. Researchers have attempted to determine a transition concentration at which salts inhibit bubble coalescence [43,44]. An important similarity between frothers and salts is the ability to produce bubbles with small size. However, bubble size is not the primary factor causing changes in bubble coalescence, as bubble coalescence mainly occurs at the frit surface where the momentum of the gas stream is high and can overcome the resistance of liquid film between the gas–liquid interfaces [45,46]. Evidently, the gas velocity is the primary factor in determining and or inhibiting bubble coalescence. Zieminski and Whittemore [47] showed that many ions of high valence have a greater effect on hindering bubble coalescence than monovalent ions.
Recently, it has been acknowledged that water should be considered a reagent in flotation, and as such, its quality is an important factor to consider because it may alter the pulp chemistry and thereby affect the froth phenomena and overall flotation performance [38,48,49,50,51,52]. Although flotation research in the past has considered electrolyte–reagent–mineral interactions on mineral hydrophobicity and therefore pulp chemistry, there seems to have been very little or no consideration of the consequences and implications that such interactions would have on the froth phase despite existing evidence in literature on the impact of inorganic electrolytes on gas dispersion properties [43,52,53].
The research gap begs the question of how reagents’ chemistry and the tripartite electrolyte–reagent–mineral interactions could affect the bubble size and water recoveries within flotation and other above-mentioned froth-stability-determining factors. The authors of this paper argue that meaningful research into water quality in flotation should consider this identified gap. Thus, building on the work published by Manono et al. [42], in which the effects of increasing ionic strength of synthetic plant waters on gas dispersion properties were considered with the use of complex water matrices, the current study was aimed at decoupling specific ion effects in order to understand which ions are detrimental or beneficial for froth stability. It is known that certain electrolytes present in process water exhibit frother-like gas dispersion properties in that they inhibit bubble coalescence [10,42,54,55,56,57,58]. Castro et al. [56] reported that water with an increased concentration of electrolytes showed a significant degree of frothing ability, hence the need for more investigations into this topic.
A few investigative studies have been published in this regard. Biçak et al. [59] suggested that the sensitivity of flotation to water quality is such that an accumulation of both organic and inorganic ions in the process water may also affect froth stability. Corin and Wiese [49] also reported that process water quality, specifically that of increased ionic strength, affected froth stability in that the frother dosage and ionic strength of plant water could be interchangeably used to increase froth stability, proving that ions in process water could cause the required step changes in froth stability and flotation performance without further increases in the dosage of the frother. These effects were supported Quinn et al. [39], who showed that certain concentrators processing sulphides in Western Australia and some parts of Chile did not add frothers in their flotation circuits due to the high salinity (from ~40,000 to 60,000 mg/L) of their process water and yet achieved the desired froth stability.
An earlier study conducted at the University of Cape Town (UCT) by Barker [37] showed that divalent cations such as Ca2+ and Mg2+ in solution resulted in an increase in the kinematic viscosity compared to the monovalent Na+ and K+, and froth stability concurrently increased. As a mechanism, Barker [37] proposed that the increase in the kinematic viscosity caused an increase in the liquid film thickness of the bubble, causing a slower drainage and thus a more stable froth. Biçak et al. [59] corroborated Barker’s proposal by showing that at high electrolyte concentrations, bubbles became more stable and did not coalesce even in the absence of a frother. A similar investigation carried out by Kurniawan et al. [60] on froth properties in coal flotation, using MgCl2, NaCl and NaClO3 solutions in the absence and presence of Dowfroth 250, showed that in the presence of Dowfroth 250, MgCl2 led to the generation of smaller bubbles, more stable froths, and higher recoveries than NaCl. NaClO3, however, resulted in the lowest froth stability and mineral recoveries. Their findings could be linked to the differences in the cation–anion combinations for the three different single salts used.
On the mechanism of the impact of ions at the air–water interface and froth stability, Biçak et al. [59] concluded that the rate of coalescence of air bubbles decreased due to either the frother molecules absorbed at the air–water interface or the presence of a stable liquid film layer on the surface of the bubbles. Biçak et al.’s work was in agreement with that of Yousef et al. [61], who stated that flotation plants that use seawater and process water with high salinity are likely to experience voluminous and stable froths due to the frothing properties of dissolved ions. These findings corroborated the work by Quinn et al. [39]. On the impact of specific ions in process water, Biçak et al. [59] argued that dissolved metal ions and sulphide oxidation products, mainly in the form of SO42− and S2O32−, had the most influence on both froth stability and surface chemistry. It was then hypothesized that increased froth stability in concentrated electrolytes was caused by an increased stability of the water layer between air bubbles [59]. Corin et al. [48] reported higher solids and water recoveries in increasing ionic strengths of synthetic plant water and suggested that these were caused by increased froth stability. These results supported the findings of Schwarz and Grano [37] and Tao et al. [38], who linked water recovery with froth stability in flotation as more stable froths were accompanied by an increase in water recovery.
Bournival et al. [62] investigated NaCl and MIBC as bubble coalescence inhibitors in relation to froth stability. It was found that NaCl at higher concentrations was as effective as MIBC in preventing bubble coalescence in a dynamic environment. The unbundling of such an effect may not be easy, as Castro et al. [56] purported that “the coalescence of bubbles in flotation is a particularly complex phenomenon especially when it takes place in saline or seawater, since both a surface-active compound (frother) and a surface-inactive compound (inorganic salt) are able to stabilise bubbles against coalescence and therefore reduce bubble size giving rise to stable froths”. Castro et al. [56] were preceded by Grau et al. [54] and Laskowski et al. [63], who showed that bubble coalescence could be completely prevented at frother concentrations exceeding a certain concentration known as the critical coalescence concentration (CCC). Laskowski et al. [62] also showed that inorganic electrolytes inhibited bubble coalescence and that inorganic electrolytes had a critical concentration beyond which their ability to inhibit coalescence was exhausted. Such an idea was not entirely new, as Zieminski et al. [64] and Lessard and Zieminski [44] had previously proposed such a phenomenon with specific salts that they had examined. A mechanism to the interaction of frothers and inorganic electrolytes at the air–water interface had been proposed by Castro et al. [56] after it was seen that increasing frother concentration resulted in a decrease in the solution surface tension, as can be expected for a surface-inactive agent, but they more interestingly showed that the ability of a frother to decrease the surface tension of a solution was increased in the presence of a coalescence-inhibiting inorganic electrolyte.

1.3. Gap in the Literature and Basis for Future Studies

Based on the presented literature, several mechanisms have been proposed to better understand the action of inorganic electrolytes in flotation. These range from the action of inorganic ions in (i) disrupting the hydration layers surrounding the particles and enhancing bubble–particle attachment, (ii) reducing the electrostatic interactions between bubbles and particles, and (iii) increasing the charge on the surface of the bubbles to prevent primary bubble coalescence [37,65]. However, as previously mentioned by Manono et al. [45], there is little or no information in the literature on the action of electrolytes on bubble coalescence and froth stability in relation to the action of ions on interactions occurring in the pulp phase that affect valuable mineral hydrophobicity, the hydrophilicity of gangue, and the resulting impact on froth stability. Such research should bring about a better understanding of the effects that the tripartite electrolyte–reagent–mineral interactions occurring in the pulp phase have on the froth phase phenomena. Thus, research is required to consider the impact of common process water ions on the pulp chemistry and froth stability in Cu-Ni-PGM sulphide ores, since recent work showed that for such ores, many of the effects seen have more significant implications on gangue management rather than on the recovery of sulfides due to the ores’ sulfides content being relatively low compared to gangue [66]. The management of gangue against varying flotation process control measures requires an understanding of how process variables affect the stability of the froth, as the stability of the froth has grave implications regarding whether the operators limit or promote entrainment [67,68].
Current understanding suggests that investigative research into the influence of each specific process water constituent on flotation could result in both technical and financial benefits, i.e., costs could be saved in process water treatment if research proves that flotation circuits could still perform at their best in recycled process water containing constituents that would have previously been removed (as long as key aspects of flotation such as the froth behaviour are well-understood and controlled against varying process water quality). This would allow for an implementation of closed water circuits in flotation processes and a contribution to the sustainability of the flotation process if flotation process water is screened for inorganic electrolytes that would be detrimental to the process. It is proposed that if such implementation were to take place, there would be minimal process water discharge to the environment, minimal make-up water requirements, a reduction in process water treatment before reuse, savings in treatment costs, and a reduction in fresh reagent dosing. Thus, the aim of this study was to ascertain whether there are ions in flotation with dominant effects on froth stability using various single salts such as CaCl2, CaNO3, CaSO4, MgCl2, Mg(NO3)2, MgSO4, NaCl, NaNO3 and Na2SO4. Froth stability was examined using a three-phase batch flotation system with, solid and water recoveries being used as proxies for froth stability. To complement the three-phase study, gas dispersion properties were measured using two-phase batch flotation and column studies to generate experimental data on bubble size, water recoveries, foam height and foam collapse time.

2. Materials and Experimental Methods

2.1. Water Preparation

Synthetic plant water (hereafter referred to as SPW) and various single salt solutions of the ionic strength and compositions shown in Table 1 were prepared. The synthetic plant water and single salt solutions comprised de-ionised water and inorganic salts of analytical grade supplied by Merck to ensure that the water contained the required total dissolved solids (TDS) and ionic strength. The ionic strength (I) was calculated using Equation (1). The preparation of the single salts was such that these solutions had a fixed ionic strength equivalent to that of SPW (0.0213 M). It is worth mentioning that the composition of synthetic plant water was based on the onsite process water quality of a typical Cu-Ni-PGM ore concentrator [3].
I = 1 2 i = 1 n c i z i 2
where I is the total ionic strength, a is a measure of the electrical intensity of a solution, ci represents the concentration of the ith species for an ideal or dilute solution, and zi is the ionic valency of the ith species in the solution, the summation being taken over all positive and negative ions (n) present in solution [69].

2.2. Ore Preparation and Milling

A sulfidic Cu-Ni-PGM containing ore from the Merensky Reef found within the Bushveld Igneous Complex of South Africa, with the mineralogical composition shown in Figure 1, was used throughout the bench scale flotation tests. It can be seen from Figure 1 that only ~1.09% constituted sulphide minerals and the rest were gangue, with talc (which is a naturally floatable gangue and therefore problematic for flotation) notably comprising 3.51% of this ore. A bulk sample of the selected ore was crushed, riffled, and split into 1 kg samples using a rotary splitter. The 1 kg ore samples were milled in a stainless-steel tumbling mill at 66% solids in the presence of synthetic plant water or a single salt solution with Sodium Iso-Butyl Xanthate (SIBX) as the collector dosed into the mill at a dosage of 150 g/t. The addition of synthetic plant water or a single salt solution was such that after milling, a target grind of 60% passing 75 µm was attained. This grind size is typical of industrial concentrator rougher circuits for Merensky ores. The grinding media used inside the tumbling mill were 20 rods of stainless steel with varying diameters: 6 × 25 mm, 8 × 20 mm, and 6 × 16 mm.

2.3. Batch Flotation

For the three-phase batch flotation procedure, after milling, the milled Merensky ore slurry was transferred to a 3 L UCT Barker flotation cell, and the specific water or single salt solution under investigation was added to the cell to ensure a pulp density of 35%. Then, the impeller was set to an agitation speed of 1200 rpm, and a feed sample was taken. DOW 200 frother at 40 g/t was allowed to condition for 1 min, after which air was introduced to the cell at 7 L/min. A constant froth height of 2 cm was maintained. The pH of the flotation experiments was measured and found to be ~9. This was unadjusted and uncontrolled because it was found to represent the acceptable and natural operating flotation pH for ores typical of the selected ore. Four concentrates were collected for 2, 4, 6, and 10 min, and they were scraped at 15 s intervals. At the end of flotation, a tails sample was taken. All feed, tail, and concentrate samples were filtered, dried, weighed, and analysed for S in the UCT Chemical Engineering Analytical Services Laboratory using an in-house Leco. Total S assays were conducted to determine the sulphide gangue fraction of the solids reporting to the concentrate because previous investigations showed strong correlations between froth stability and gangue recoveries, which is important for the management of gangue and the concentrate grade [66]. All tests were performed in duplicate for reproducibility. The two-phase batch flotation procedure was described by Manono et al. [42].

2.4. Two-Phase Foam Column Procedure

Specific ion effects on foam stability were investigated using a two-phase foam column test rig. The test rig was discussed in detail by Manono et al. [42]. Two-phase foam column tests were conducted for various single salts in distilled water at pH 9, representing the pH at which the batch floats were conducted. These foam studies were conducted in the presence of a frother (DOW 200) at a dosage of 5 ppm. The pH was adjusted using a stock solution of NaOH to achieve the desired pH before the solution was added into the column, where the initial air–water interface height was measured. Air was pumped into the column at a flow rate of 2 L/min. This allowed for the development of the foam phase until an equilibrium height was reached. This equilibrium condition was reached after approximately 3 min for selected single salt solutions, namely CaCl2, CaSO4, Ca(NO3)2, NaCl, NaNO3 and Na2SO4. The equilibrium condition is defined as the state at which the bubble collapse rate at the top of the foam is equal to the air flow rate through the sparger at the bottom of the column. Once equilibrium was reached, air was stopped and the foam collapse time (i.e., the time taken for the foam to collapse to the initial air–water interface) was immediately measured. All experiments were performed in duplicate to minimise error.

2.5. Two-Phase UCT Bubble Size Analyser

The UCT Bubble Sizer (Randall et al. [70]), a schematic of which is shown in Figure 2 was used for all bubble sizing tests. This apparatus consists of a capillary tube that passes between two pairs of photo-cell-diode detectors. The detectors are mounted 5 mm apart at right angles to one another. Bubbles drawn up the capillary by vacuum are transformed into cylinders, so as they pass the detectors, a change in light intensity is monitored. This change occurs because of the different refractive indices and optical densities of the air–glass and liquid–glass boundaries. To obtain an accurate bubble size distribution, 3000 bubbles are sized in any one test. The bubbles are then drawn from the capillary into a burette so that the volume of bubbles can be determined [71].
The bubble sizer was used with a 1 L glass column fitted with a frit in the base. This column was attached to a gas sparger to allow for the generation of bubbles. To study the effect of different ions in water, SPW and selected single salt solutions were used. All solutions were set to the specified ionic strength in de-ionised water. For all tests, the addition of a frother (DOW 200) was kept constant at 5 ppm. Tests were conducted for 10–20 s each. A peristaltic pump was used to draw the samples through the capillary and into the measuring burette; this pump was set at 790 rpm. All bubble sizes measured were Sauter mean diameters. Bubble size measurements were repeated for each condition, and the results were shown to be reproducible. The reproducibility of this system was also demonstrated by Tucker et al. [72].

3. Results and Discussion

3.1. Specific Ion Effects on Water Recoveries

Figure 3 shows clearly discernible differences in the water and gangue recoveries in various single salts, though sulphide recoveries remained constant despite changes in single salts. Therefore, it could be said that changes in the stability of the froth did not affect the flotation recovery of sulphides, as the amount of sulphides reporting to the concentrate was well within an average of ~6.5 g. SPW, Na2SO4, NaCl, and NaNO3 resulted in significantly lower water recoveries compared to the Ca2+- and Mg2+-bearing single salt solutions. Conversely, CaCl2, CaSO4, Ca(NO3)2, MgCl2, MgSO4, and Mg(NO3)2 resulted in significantly higher water recoveries. It is worth mentioning that for Ca2+, Ca(NO3)2 yielded lower water recoveries than CaCl2 and CaSO4. A similar trend was observed with Mg2+, as Mg(NO3)2 yielded lower water recoveries than MgCl2 and MgSO4. SO42−, however, yielded larger water recoveries than Cl and NO3 in Ca2+ and Mg2+. Furthermore, for Ca2+ and Mg2+, gangue recoveries followed similar trends as water recoveries. However, Na+-containing salts resulted in higher gangue recoveries than Ca2+ and Mg2+. This behaviour was not in line with the lower water recovery trend with Na+ salts, as the lower water recoveries could be used to infer that the stability of the froth in Na+ was somewhat poorer. Therefore, higher gangue content in Na+ salts could be attributed to the poor passivation of gangue particles by monovalent cations, which was reported by Manono et al. [73], who showed that the zeta potential of talc and the coagulation of a Merensky ore were lower than when single salt solutions of Ca2+ and Mg2+ were used.
It is generally accepted that water recovery in flotation is a good indicator of the stability of the froth [11]. The results in Figure 3 show higher water recoveries in the presence of divalent cations, Ca2+ and Mg2+, compared to the monovalent Na+. For example, Ca2+ and Mg2+ resulted in water recoveries mostly within the range 770–880 g, while Na+ resulted in water recoveries within the range 600–750 g. Additionally, it is apparent that the presence of the divalent SO42− anion increased the water recoveries and froth stability compared to the monovalent Cl and NO3, except when combined with Na+. This was in line with the work of Farrokhpay and Zanin [7], who showed that froth stability was higher in the presence of multivalent metal ions than monovalent metal ions.
Figure 4 shows the amount of water (foam) reporting to the concentrate for various single salts including SPW in the absence of solids. As synthetic plant water and Na+ resulting in only 690 and ~730 g of recovered water, respectively, it can be seen that Na+ and SPW resulted in lower water recoveries compared to the Ca2+ and Mg2+ single salt solutions because these divalent cations, Ca2+ and Mg2+, resulted in ~740 g of water recovered. The anion effects on water recoveries were not as apparent in the three-phase study, the results of which are reported in Figure 3; here, it can be seen that there were no significant difference between the water recoveries obtained for NaCl, Na2SO4, and NaNO3. The same can be said for Ca2+, except for Mg2+ because Mg(NO3)2 resulted in lower water recoveries than MgCl2 = MgSO4. This difference between the results shown in Figure 3 and Figure 4 could be attributed to the role of the different species that form in solution with studied anions because these species may behave differently in solution in the presence of mineral particles, as was shown by the zeta potential results reported by Manono et al. [73].
In terms of the higher water recoveries in Ca2+ and Mg2+ than in Na+, Figure 4 shows a similar trend to that shown in Figure 3. The non-apparent effect on the anion effect on water recoveries per cation type seen in Figure 3 compared to the apparent anion effects shown in Figure 4 suggests that no direct correlation between the water recoveries in the three-phase studies and the water recoveries in two-phase studies can be made without considering pulp rheology and how specific anions combined with cations would form—possibly as different ionic complexes in solution that could act differently on the mineral surface and thus differently affect mineral hydrophobicity and froth stability. To further examine this, the results of foam studies regarding foam height, foam collapse time, and bubble size data are presented.

3.2. Specific Ion Effects on Foam Height and Foam Collapse Time

Figure 5 depicts the foam stability, measured as foam collapse time and the foam height for the selected single salts. The monovalent Na+-containing single salt solution resulted in lower foam stabilities compared to the divalent Ca2+-containing single salt solutions. For example, Na+ resulted in an average foam height and foam stability of ~108 mm and ~8.6 s, respectively, whereas these values for Ca+ were ~249 mm and ~28 s, respectively. Considering the impact of ion combinations, for Na+-containing single salt solutions, the stability of the foam was greater with SO42− ions in solution compared to NaNO3 and NaCl. NaNO3, however, resulted in the lowest foam stability of all Na+ solutions, and this trend was similar to that shown in Figure 3 concerning water recoveries. For Ca2+, SO42−-containing solutions resulted in higher foam stabilities compared to Cl and NO3, whereas NO3 resulted in the lowest foam stability. Once again, this trend is similar to that of water recoveries results shown in Figure 3. The foam height for Na+-containing salts was in the order of Na2SO4 > NaNO3 > NaCl, and the foam heigh for Ca2+ was in the order of CaSO4 = Ca(NO3)2 > CaCl2.
The results in Figure 3 on water recoveries for a three-phase system proved to be in line with the proposal that divalent ions would have a greater impact on the behaviour of the froth. Figure 4′s foam stability results also support this proposal because the two-phase foam stability test results shown in Figure 5 clearly show that Ca2+ resulted in higher foam stabilities than the monovalent Na+-containing single salt solutions across all anion types. The observations from Figure 5 also corroborate well with the specific anion effect trend observed in Figure 3, i.e., SO42− generally resulted in higher foam stabilities compared to Cl and NO3 and an even lower foam stability was observed for NO3. The findings of this work are in line with those of Kurniawan et al. [60], who showed that Mg2+ resulted in increased froth stability and smaller froth bubble size and thus higher mineral recovery compared to Na+ when single salt solutions of MgCl2, NaCl, and NaClO3 were investigated during coal flotation in the absence and presence of Dowfroth 250. Their findings were also in line with those of other authors who showed that water with an increased concentration of electrolytes resulted in an increase in froth stability [33,36,59,60]. Biçak et al. [59] attributed this effect to the presence of a stable layer on the surface of the bubbles, and they showed higher froth stabilities in the presence of SO42− and S2O32− than any other anions; the findings of this paper were similar to ours in that the SO42−-containing solutions had higher water recoveries and foam stability than Cl and NO3, a trend that was particularly noticeably in the presence of the divalent Ca2+ and Mg2+. It is important to mention that these previous studies do not seem to have been oriented towards understanding process water chemistry effects on froth stability in PGM ores. Thus, the findings of this paper should serve as a tool for two-phase foam stability studies as complements to three-phase studies considering the effects of specific ions on froth stability during the flotation of Cu-Ni-PGM ores.

3.3. Specific Ion Effects on Bubble Size

Figure 6 depicts the Sauter mean bubble diameter for SPW and various single salt solutions investigated in this study. The bubble diameter did not significantly change with ion type, although it was expected that the divalent Ca2+ and Mg2+ would result in smaller Sauter mean bubble diameters compared to the monovalent Na+. It is worth nothing the slight decreases in bubble diameter seen with Ca2+, Mg2+, and SPW-containing solutions compared to Na+ (as expected), which is demonstrated by the fact that Ca2+ and Mg2+ resulted in an average bubble diameter of ~1.31 mm while Na+ resulted in an average of 1.50 mm. The slight decreases in bubble diameter in Ca2+ and Mg2+ support the higher water recoveries observed in Figure 3 and Figure 4 and are consistent with previous studies [39,43,58,65,74].

3.4. Concluding Discussion

Lessard and Zieminski [44] investigated the effects of various inorganic electrolytes, shown in Table 2, on bubble coalescence and interfacial gas transfer in aqueous solutions. In their experiments, several pairs of bubbles were contacted, enabling the evaluation of the percentage degree of bubble coalescence as a function of the molar concentration of the inorganic electrolyte. Their findings showed a sharp transition concentration from higher degrees of coalescence to a 50% reduction on bubble coalescence. Thus, this concentration, which resulted in 50% bubble coalescence inhibition, was defined as the transition concentration at which coalescence was sharply reduced. This enabled a comparison of the effectiveness of the salts in retarding bubble coalescence. Table 2 shows that the monovalent cations, Na+ and K+, exhibited higher transition concentrations compared to the multivalent cations, Al3+, Ca2+ and Mg2+. This lower transition concentration achieved by the multivalent cations was indicative of their superior ability to retard bubble coalescence and thus their froth-stability-enhancing behaviour compared to monovalent cations. Additionally, considering the transition concentration of MgSO4 in comparison to MgCl2 and that of Na2SO4 in comparison with NaCl, it is evident that SO42− had a lower transition concentration of 0.032 mol.dm−3 and 0.061 mol.dm−3 in Mg2+ and Na+, respectively, compared to Cl. Lessard and Zieminski’s findings could explain why higher water recoveries and foam stabilities were obtained for SO42−-containing solutions than Cl and NO3 solutions in this study.
In this study, we aimed to answer the question of which ions have the greatest impact on froth stability through batch flotation and two-phase foam column studies with various single salts. The results showed that generally divalent cations of Ca2+ and Mg2+ led to greater water recoveries compared to the monovalent Na+. SO42−-containing salts resulted in higher water recoveries than NO3 and Cl. NO3 had a lesser froth-stabilising effect than SO42−, as indicated by the lower water recoveries and foam stabilities reported in Figure 3 and Figure 5, respectively. Foam column studies were performed using CaCl2, CaSO4, CaNO3, NaCl, Na2SO4, and NaNO3 at an ionic strength of 0.0213 mol.dm−3 equivalent of 1 SPW. The foam collapse time was longer for Ca2+-containing solutions than those that contained Na+ for each corresponding anion type. SO42− resulted in a longer foam collapse time than Cl and NO3. The NO3 ion resulted in the lowest foam stability. The foam stability trend reported in Figure 5 agreed with that of the water recoveries reported in Figure 3 and Figure 4, especially for the effect of the monovalent cation vs. divalent cations on foam stability and water recovery. In line with the work of by Lessard and Zieminski [44], this was attributed to the ability of the ions to retard bubble coalescence, although the results presented in Figure 6 do not demonstrate any strongly apparent effect on bubble diameter, possibly due to the low ion concentration used in this study compared to the bubble coalescence transition concentration proposed by Lessard and Zieminski [44]. The Sauter mean bubble diameters were, however, slightly smaller in divalent cations than in monovalent cations. Table 3 depicts the molar concentrations of the single salts studied in this work to further examine this behaviour.
From Table 3, it can be seen that although all the single salts were studied at a fixed ionic strength of 0.0213 mol.dm−3 equivalent to the ionic strength of typical of process water in Merensky concentrators [75], solution concentration or molarity differed with each salt, as did the specific ion molar concentrations. For the monovalent Na+, a higher amount of the salt needed to be dissolved in the solution compared to the divalent Ca2+ and Mg2+-containing single salts, as shown by the higher TDS and total molarity for Na+ salts compared to both Ca2+ and Mg2+. Table 3, with specific species concentrations or molarities, is presented to enable comparison with the results of Lessard and Zieminski [44], who reported bubble coalescence for single salts in the form of the 50% transition concentration presented in Table 2. Lessard and Zieminski [44] showed that for a fixed cation (e.g., MgSO4 vs. MgCl2 and NaCl vs. Na2SO4), the concentration required to retard 50% bubble coalescence was lower for SO42− than Cl, while for a fixed anion type, the required concentration to retard half of the coalescence occurring between bubbles was lower for Mg2+ and Ca2+ than Na+. These results agree with the findings of this study. From Table 3, it can also be seen that for all single salts, a lower specific anion concentration was required for the SO42− to reach a total ionic strength of 0.0213 mol.dm−3 compared to Cl and NO3. However, all the tested single salts were present at molarities lower than their concentrations required to retard 50% bubble coalescence. It is evident that specific ion concentration and specific ion valency are important aspects in considering froth stability, as shown by the fact that Ca2+ in combination with SO42− had a concentration of 0.0053 mol.dm−3, which is lower than that of Ca2+ (0.0071 mol.dm−3) in combination with Cl and NO3 but still resulted in the highest water recoveries and therefore highest froth stability. This is attributed to the divalent nature of the SO42− anion compared to the monovalent Cl and NO3 anions; hence, lower concentrations would be required for the SO42− anion compared to the Cl or NO3 anions to achieve the same effect on bubble coalescence [44]. This trend was sustained across cation type, as shown in Table 3. However, it is acknowledged that gas dispersion properties that enable the more robust characterisation of froth such, e.g., bubble size, froth collapse time, froth height, gas hold-up, and froth rise velocity, need to be determined in the presence of solids. This together with the findings of this study should form the basis for future investigative work.

4. Conclusions

Water recoveries and foam stability studies showed that SO42−, Ca2+ and Mg2+ had greater froth stabilising effects than Cl−, NO3 and Na+. It was also shown that NO3 had the least froth stabilising effect. This proves the importance of considering changes in froth stability with water quality variations, as this could have serious implications on the management of gangue and concentrate grades.
Thus, the following key conclusions can be drawn:
  • SO42−, Ca2+, and Mg2+ resulted in higher water recoveries in the three-phase systems compared to Cl, NO3, and Na+.
  • Ca2+ and Mg2+ resulted in higher water recoveries in the two-phase systems compared to Na+. No apparent anion effect was observed.
  • Ca2+ resulted in increased foam stability compared to Na+.
  • SO42 resulted in higher foam stability compared to Cl, and NO3 led to the lowest foam stability.
  • Foam heights were far more increased in single salt solutions consisting of Ca2+ and SO42− than those containing Cl, NO3, and Na+.
  • There was no apparent effect on bubble diameter. However, slight decreases in bubble diameter were observed in the presence of Ca2+ and Mg2+ compared to Na+.

Author Contributions

Conceptualization, M.S.M. and K.C.C.; methodology, M.S.M.; software, M.S.M.; validation, M.S.M. and K.C.C.; formal analysis, M.S.M.; investigation, M.S.M.; resources M.S.M. and K.C.C.; data curation, M.S.M. and K.C.C.; writing—original draft preparation, M.S.M.; writing—review and editing, M.S.M. and K.C.C.; visualization, M.S.M.; supervision, K.C.C.; project administration, M.S.M. and K.C.C.; funding acquisition, M.S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported in part by the National Research Foundation (NRF) of South Africa and the University of Cape Town (Grant number 118062—Water Quality in Flotation: Causes, Impacts, and Process Management Strategies, 2019–2021). Any opinions, findings, conclusions, or recommendations expressed in any publication generated by NRF supported research is that of the authors, and the NRF accepts no liability whatsoever in this regard. The APC was funded by the University of Cape Town.

Data Availability Statement

Not applicable.

Acknowledgments

Financial and technical contributions from the South African Minerals to Metals Research Institute (SAMMRI). Members of the Centre for Minerals Research’s Reagent Research Group are also acknowledged.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Wills, B.A.; Napier-Munn, T. Mineral Processing Technology. An Introduction to the Practical Aspect of Ore Treatment and Mineral Recovery; Elsevier Science & Technology Books: Amsterdam, The Netherlands, 2006; Volume 7. [Google Scholar]
  2. Martinovic, J.; Bradshaw, D.J.; Harris, P.J. Investigation of surface properties of gangue minerals in platinum bearing ores. J. S. Afr. Inst. Min. Metall. 2005, 105, 349–356. [Google Scholar]
  3. Wiese, J.; Harris, P.; Bradshaw, D. The influence of the reagent suite on the flotation of ores from the Merensky reef. Miner. Eng. 2005, 18, 189–198. [Google Scholar] [CrossRef]
  4. Barbian, N.; Hadler, K.; Ventura-Medina, E.; Cilliers, J.J. The froth stability column: Linking froth stability and flotation performance. Miner. Eng. 2005, 18, 317–324. [Google Scholar] [CrossRef]
  5. Johansson, G.; Pugh, R.J. The influence of particle size and hydrophobicity on the stability of mineralized froths. Int. J. Miner. Process. 1992, 34, 1–21. [Google Scholar] [CrossRef]
  6. Schwarz, S.; Grano, S. Effect of particle hydrophobicity on particle and water transport across a flotation froth. Colloids Surf. A Physicochem. Eng. Asp. 2005, 256, 157–164. [Google Scholar] [CrossRef]
  7. Farrokhpay, S.; Zanin, M. An investigation into the effect of water quality on froth stability. Adv. Powder Technol. 2012, 23, 493–497. [Google Scholar] [CrossRef]
  8. Farrokhpay, S. The significance of froth stability in mineral flotation—A review. Adv. Colloid Interface Sci. 2011, 166, 1–7. [Google Scholar] [CrossRef]
  9. Schreithofer, N.; Wiese, J.; McFadzean, B.; Harris, P.; Heiskanen, K.; O’Connor, C. Frother-depressant interactions in two and three phase systems. Int. J. Miner. Process. 2011, 100, 33–40. [Google Scholar] [CrossRef]
  10. Ramos, O.; Castro, S.; Laskowski, J.S. Copper-molybdenum ores flotation in sea water: Floatability and frothability. Miner. Eng. 2013, 53, 108–112. [Google Scholar] [CrossRef]
  11. Wiese, J.G.; Harris, P.J.; Bradshaw, D.J. The effect of increased frother dosage on froth stability at high depressant dosages. Miner. Eng. 2010, 23, 1010–1017. [Google Scholar] [CrossRef]
  12. Tao, D.; Luttrell, G.H.; Yoon, R.H. A parametric study of froth stability and its effect on column flotation of fine particles. Int. J. Miner. Process. 2000, 59, 25–43. [Google Scholar] [CrossRef]
  13. McFadzean, B.; Marozva, T.; Wiese, J. Flotation frother mixtures: Decoupling the sub-processes of froth stability, froth recovery and entrainment. Miner. Eng. 2016, 85, 72–79. [Google Scholar] [CrossRef]
  14. Zanin, M.; Wightman, E.; Grano, S.R.; Franzidis, J.P. Quantifying contributions to froth stability in porphyry copper plants. Int. J. Miner. Process. 2009, 91, 19–27. [Google Scholar] [CrossRef]
  15. Iglesias, E.; Forgiarini, A. A new method to estimate the stability of short-life foams. Colloids Surf. A Physicochem. Eng. Asp. 1995, 98, 167–174. [Google Scholar] [CrossRef]
  16. Barbian, N.; Ventura-Medina, E.; Cilliers, J.J. Dynamic froth stability in froth flotation. Miner. Eng. 2003, 16, 1111–1116. [Google Scholar] [CrossRef]
  17. Barbian, N.; Hadler, K.; Cilliers, J.J. The froth stability column: Measuring froth stability at an industrial scale. Miner. Eng. 2006, 19, 713–718. [Google Scholar] [CrossRef]
  18. Ata, S.; Ahmed, N.; Jameson, G.J. A study of bubble coalescence in flotation froths. Int. J. Miner. Process. 2003, 72, 255–266. [Google Scholar] [CrossRef]
  19. Ventura-Medina, E.; Cilliers, J.J. A model to describe flotation performance based on physics of foams and froth image analysis. Int. J. Miner. Process. 2002, 67, 79–99. [Google Scholar] [CrossRef]
  20. Harvey, P.A.; Nguyen, A.V.; Jameson, G.J.; Evans, G.M. Influence of sodium dodecyl sulphate and Dowfroth frothers on froth stability. Miner. Eng. 2005, 18, 311–315. [Google Scholar] [CrossRef]
  21. Bikerman, J. Foams ([Applied Physics and Engineering] 10); Springer: Berlin/Heidelberg, Germany, 1972. [Google Scholar]
  22. Sheni, N.; Corin, K.; Wiese, J. Considering the effect of pulp chemistry during flotation on froth stability. Miner. Eng. 2018, 116, 15–23. [Google Scholar] [CrossRef] [Green Version]
  23. Bournival, G.; Zhang, F.; Ata, S. Coal Flotation in Saline Water: Effects of Electrolytes on Interfaces and Industrial Practice. Miner. Process. Extr. Metall. Rev. 2021, 42, 53–73. [Google Scholar] [CrossRef]
  24. Sinche-Gonzalez, M.; Fornasiero, D. Understanding the effect of sulphate in mining-process water on sulphide flotation. Miner. Eng. 2021, 165, 106865. [Google Scholar] [CrossRef]
  25. October, L.L.; Corin, K.C.; Manono, M.S.; Schreithofer, N.; Wiese, J.G. A fundamental study considering specific ion effects on the attachment of sulfide minerals to air bubbles. Miner. Eng. 2020, 151, 106313. [Google Scholar] [CrossRef]
  26. Moimane, T.M.; Corin, K.C.; Wiese, J.G. The effect of varying pulp reagent chemistry on the flotation performance of a South African PGM ore. Miner. Eng. 2016, 95, 155–160. [Google Scholar] [CrossRef]
  27. Shortridge, P.G.; Harris, P.J.; Bradshaw, D.J. The influence of ions on the effectiveness of polysaccharide depressants in the flotation of talc. In Proceedings of the Polymers in Mineral Processing. 3rd UBC-McGill Bi-Annual International Symposium on Fundamentals of Mineral Processing; Laskowski, J.S., Ed.; CIM: Quebec City, QC, Canada, 1999; pp. 155–169. [Google Scholar]
  28. Cisternas, L.A.; Gálvez, E.D. The use of seawater in mining. Miner. Process. Extr. Metall. Rev. 2018, 39, 18–33. [Google Scholar] [CrossRef]
  29. Michaux, B.; Rudolph, M.; Reuter, M.A. Challenges in predicting the role of water chemistry in flotation through simulation with an emphasis on the influence of electrolytes. Miner. Eng. 2018, 125, 252–264. [Google Scholar] [CrossRef]
  30. Jeldres, R.I.; Forbes, L.; Cisternas, L.A. Effect of Seawater on Sulfide Ore Flotation: A Review. Miner. Process. Extr. Metall. Rev. 2016, 37, 369–384. [Google Scholar] [CrossRef]
  31. Wang, B.; Peng, Y. The effect of saline water on mineral flotation - A critical review. Miner. Eng. 2014, 66, 13–24. [Google Scholar] [CrossRef]
  32. Muzinda, I.; Schreithofer, N. Water quality effects on flotation: Impacts and control of residual xanthates. Miner. Eng. 2018, 125, 34–41. [Google Scholar] [CrossRef]
  33. Bicak, O.; Ozturk, Y.; Ozdemir, E.; Ekmekci, Z. Modelling effects of dissolved ions in process water on flotation performance. Miner. Eng. 2018, 128, 84–91. [Google Scholar] [CrossRef]
  34. Fuerstenau, M.C. Sulphide Mineral Flotation. In Principles of Flotation; King, R.P., Ed.; SAIMM: Johannesburg, South Africa, 1982; pp. 159–182. [Google Scholar]
  35. Slatter, K.A.; Plint, N.D.; Cole, M.; Dilsook, V.; De Vaux, D.; Palm, N.; Oostendorp, B. Water Management in Anglo Platinum Process Operations: Effects of Water Quality on Process Operations, Proceedings of the International Mine Water Conference, Pretoria, South Africa, 19–23 October 2009; Document Transformation Technologies: Irene, South Africa, 2009; pp. 46–55. ISBN 978-0-9802623-5-3. [Google Scholar]
  36. Muzenda, E. An Investigation into the Effect of Water Quality on Flotation Performance. World Acad. Sci. Eng. Technol. Int. J. Chem. Mol. Nucl. Mater. Metall. Eng. 2010, 4, 562–566. [Google Scholar]
  37. Barker, L.M. The Effect of Electrolytes on the Flotation of Pyrite. Master’s Thesis, University of Cape Town, Cape Town, South Africa, 1986. [Google Scholar]
  38. Rao, S.R.; Finch, J.A. A review of water re-use in flotation. Miner. Eng. 1989, 2, 65–85. [Google Scholar] [CrossRef]
  39. Quinn, J.J.; Kracht, W.; Gomez, C.O.; Gagnon, C.; Finch, J.A. Comparing the effect of salts and frother (MIBC) on gas dispersion and froth properties. Miner. Eng. 2007, 20, 1296–1302. [Google Scholar] [CrossRef]
  40. Yoon, R.H.; Sabey, J.B. Coal flotation in inorganic salt solutions. In Interfacial Phenomena in Coal Technology; Botsaris, G.D., Glazman, Y.M., Eds.; Surfactant Science Series; CRC Press: Boca Raton, FI, USA, 1989; Volume 32, pp. 87–114. [Google Scholar]
  41. Lekki, J.; Laskowki, J.S. A new concept of frothing in flotation systems and general classification of flotation frothers. In Proceedings of the The XI International Mineral Processing Congress(IMPC1975), The International Mineral Processing Congress, Calgari, AB, Canada, 20–26 April 1975; pp. 427–448. [Google Scholar]
  42. Manono, M.S.; Corin, K.C.; Wiese, J.G. The effect of ionic strength of plant water on foam stability: A 2-phase flotation study. Miner. Eng. 2013, 40, 42–47. [Google Scholar] [CrossRef]
  43. Craig, V.S.J.; Ninham, B.W.; Pashley, R.M. The effect of electrolytes on bubble coalescence in water. J. Phys. Chem. 1993, 97, 10192–10197. [Google Scholar] [CrossRef]
  44. Lessard, R.R.; Zieminski, A.S. Bubble coalescence and gas transfer in aqueous electrolytic solutions. Ind. Eng. Chem. Fundam. 1971, 10, 260–269. [Google Scholar] [CrossRef]
  45. Firouzi, M.; Nguyen, A. V On the effect of van der Waals attractions on the critical salt concentration for inhibiting bubble coalescence. Miner. Eng. 2014, 58, 108–112. [Google Scholar] [CrossRef]
  46. Firouzi, M.; Nguyen, A. V Effects of monovalent anions and cations on drainage and lifetime of foam films at different interface approach speeds. Adv. Powder Technol. 2014, 25, 1212–1219. [Google Scholar] [CrossRef] [Green Version]
  47. Zieminski, S.A.; Whittemore, R.C. Behaviour of gas bubbles in aqueous electrolyte solutions. Chem. Eng. Sci. 1971, 26, 509–520. [Google Scholar] [CrossRef]
  48. Corin, K.C.; Reddy, A.; Miyen, L.; Wiese, J.G.; Harris, P.J. The effect of ionic strength of plant water on valuable mineral and gangue recovery in a platinum bearing ore from the Merensky reef. Miner. Eng. 2011, 24, 131–137. [Google Scholar] [CrossRef]
  49. Corin, K.C.; Wiese, J.G. Investigating froth stability: A comparative study of ionic strength and frother dosage. Miner. Eng. 2014, 66, 130–134. [Google Scholar] [CrossRef]
  50. Wang, B.; Peng, Y. The behaviour of mineral matter in fine coal flotation using saline water. Fuel 2013, 109, 309–315. [Google Scholar] [CrossRef]
  51. Ikumapayi, F.; Makitalo, M.; Johansson, B.; Rao, K.H. Recycling of process water in sulphide flotation: Effect of calcium and sulphate ions on flotation of galena. Miner. Eng. 2012, 39, 77–88. [Google Scholar] [CrossRef] [Green Version]
  52. Manono, M.S.; Corin, K.C.; Wiese, J.G. An investigation into the effect of various ions and their ionic strength on the flotation performance of a platinum bearing ore from the Merensky reef. Miner. Eng. 2012, 36–38, 231–236. [Google Scholar] [CrossRef]
  53. Manono, M.; Corin, K.; Wiese, J. Perspectives from literature on the influence of inorganic electrolytes present in plant water on flotation performance. Physicochem. Probl. Miner. Process. 2018, 54, 1191–1214. [Google Scholar]
  54. Grau, R.A.; Laskowski, J.S.; Heiskanen, K. Effect of frothers on bubble size. Int. J. Miner. Process. 2005, 76, 225–233. [Google Scholar] [CrossRef]
  55. Laskowski, J.; Castro, S. Flotation in concentrated electrolyte solutions. Int. J. Miner. Process. 2015, 144, 50–55. [Google Scholar] [CrossRef]
  56. Castro, S.; Miranda, C.; Toledo, P.; Laskowski, J.S. Effect of frothers on bubble coalescence and foaming in electrolyte solutions and seawater. Int. J. Miner. Process. 2013, 124, 8–14. [Google Scholar] [CrossRef]
  57. Henry, C.L.; Dalton, C.N.; Scruton, L.; Craig, V.S.J. Ion-specific coalescence of bubbles in mixed electrolyte solutions. J. Phys. Chem. C 2007, 111, 1015–1023. [Google Scholar] [CrossRef]
  58. Craig, V.S.J. Bubble coalescence and specific-ion effects. Curr. Opin. Colloid Interface Sci. 2004, 9, 178–184. [Google Scholar] [CrossRef]
  59. Biçak, Ö.; Ekmekçi, Z.; Can, M.; Öztürk, Y. The effect of water chemistry on froth stability and surface chemistry of the flotation of a Cu-Zn sulfide ore. Int. J. Miner. Process. 2012, 102–103, 32–37. [Google Scholar] [CrossRef]
  60. Kurniawan, A.U.; Ozdemir, O.; Nguyen, A.V.; Ofori, P.; Firth, B. Flotation of coal particles in MgCl2, NaCl, and NaClO3 solutions in the absence and presence of Dowfroth 250. Int. J. Miner. Process. 2011, 98, 137–144. [Google Scholar] [CrossRef]
  61. Yousef, A.A.; Arafa, M.A.; Ibrahim, S.S.; Abdel Khalek, M.A. Seawater usage in flotation for minerals beneficiation in arid regions-Simulation and Application. In Proceedings of the 22nd International Mineral Processing Congress; Lorenzen, L., Bradshaw, D.J., Eds.; South African Institute of Mining and Metallurgy: Cape Town, South Africa, 2003; Volume 2, pp. 1023–1031. [Google Scholar]
  62. Bournival, G.; Pugh, R.J.; Ata, S. Examination of NaCl and MIBC as bubble coalescence inhibitor in relation to froth flotation. Miner. Eng. 2012, 25, 47–53. [Google Scholar] [CrossRef]
  63. Laskowski, J.S.; Cho, Y.S.; Ding, K. The effect of frothers on bubble size and foam stability in potash ore flotation systems. Can. J. Chem. Eng. 2003, 81, 63–69. [Google Scholar] [CrossRef]
  64. Zieminski, S.; Caron, M.; Blackmore, R. Behaviour of air bubbles in dilute aqueous solutions. Ind. Eng. Chem. Fundam. 1967, 6, 89–96. [Google Scholar] [CrossRef]
  65. Quinn, J.J.; Sovechles, J.M.; Finch, J.A.; Waters, K.E. Critical coalescence concentration of inorganic salt solutions. Miner. Eng. 2014, 58, 1–6. [Google Scholar] [CrossRef]
  66. Manono, M.; Corin, K.; Wiese, J. Water quality effects on a sulfidic PGM ore: Implications for froth stability and gangue management. Physicochem. Probl. Miner. Process. 2018, 54, 1253–1265. [Google Scholar] [CrossRef]
  67. O’Connor, C.; Wiese, J.; Corin, K.; McFadzean, B. On the Management of Gangue Minerals in the Flotation of Platinum Group Minerals. Min. Metall. Explor. 2019, 36, 55–62. [Google Scholar] [CrossRef]
  68. Connor, C.O.; Wiese, J.; Corin, K.; Mcfadzean, B. A review of investigations into the management of gangue in the flotation of platinum group minerals. Physicochem. Probl. Miner. Process. 2018, 54, 1107–1115. [Google Scholar]
  69. Garrels, R.M.; Christ, C.L. Solutions, Minerals, and Equilibria; Jones and Bartlett Publishers: London, UK, 1990. [Google Scholar]
  70. Randall, E.W.; Goodall, C.M.; Fairlamb, P.M.; Dold, P.L.; O’Connor, C.T. A method for measuring the sizes of bubbles in two- and three-phase systems. J. Phys. E 1989, 22, 827–833. [Google Scholar] [CrossRef]
  71. Goodall, C.M. The Effects of Flotation Variables on the Bubble Size, Mixing Characteristics and Froth Behaviour in Column Flotation Cells. Ph.D. Thesis, University of Cape Town, Cape Town, South Africa, 1992. [Google Scholar]
  72. Tucker, J.P.; Deglon, D.A.; Franzidis, J.P.; Harris, M.C.; O’Connor, C.T. An evaluation of a direct method of bubble size distribution measurement in a laboratory batch flotation cell. Miner. Eng. 1994, 7, 667–680. [Google Scholar] [CrossRef]
  73. Manono, M.S.; Corin, K.C.; Wiese, J.G. The Behavior of Gangue During the Flotation of a Sulfidic PGM-Bearing Ore in Response to Various Monovalent and Divalent Ions in Process Water. Front. Chem. 2020, 8, 79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kracht, W.; Rebolledo, H. Study of the local critical coalescence concentration (l-CCC) of alcohols and salts at bubble formation in two-phase systems. Miner. Eng. 2013, 50–51, 77–82. [Google Scholar] [CrossRef]
  75. Wiese, J.G. Investigating Depressant Behaviour in the Flotation of Selected Merensky Ores. Master’s Thesis, Built Faculty of Engineering & the Built Environment, University of Cape Town, Cape Town, South Africa, 2009. [Google Scholar]
Figure 1. QEMSCAN modal composition of sulfides and gangue minerals of the selected Merensky ore.
Figure 1. QEMSCAN modal composition of sulfides and gangue minerals of the selected Merensky ore.
Minerals 12 00321 g001
Figure 2. A schematic representation of the UCT bubble size analyser (adapted from the work of Randall et al. [70] and Goodall [71]).
Figure 2. A schematic representation of the UCT bubble size analyser (adapted from the work of Randall et al. [70] and Goodall [71]).
Minerals 12 00321 g002
Figure 3. Amounts of water, total sulphides, and total gangue reporting to the concentrate for all single salts and SPW in a three-phase system.
Figure 3. Amounts of water, total sulphides, and total gangue reporting to the concentrate for all single salts and SPW in a three-phase system.
Minerals 12 00321 g003
Figure 4. Amounts of water reporting to the concentrates for all single salts and SPW in a two-phase system.
Figure 4. Amounts of water reporting to the concentrates for all single salts and SPW in a two-phase system.
Minerals 12 00321 g004
Figure 5. Specific ion effects on foam stability and foam height.
Figure 5. Specific ion effects on foam stability and foam height.
Minerals 12 00321 g005
Figure 6. Specific ion effects on the mean bubble diameter.
Figure 6. Specific ion effects on the mean bubble diameter.
Minerals 12 00321 g006
Table 1. Chemical compositions of synthetic plant water and the various single salt solutions tested in this study.
Table 1. Chemical compositions of synthetic plant water and the various single salt solutions tested in this study.
Water TypeCa2+ (mg/L)Mg2+ (mg/L)Na+ (mg/L)Cl (mg/L)SO42− (mg/L)NO3 (mg/L)CO32− (mg/L)TDS (mg/L)Ionic Strength [M]
SPW80701532872401761710230.0213
NaCl--490755---12450.0213
Na2SO4--326-682--10090.0213
NaNO3--490--1321-18100.0213
CaCl2285--503---7880.0213
CaSO4213---512--7250.0213
Ca(NO3)2285----880-11650.0213
MgCl2-173-503---6760.0213
MgSO4-129--512--6410.0213
Mg(NO3)2-173---880-10530.0213
Table 2. Bubble coalescence transition concentration in different inorganic electrolytes [44].
Table 2. Bubble coalescence transition concentration in different inorganic electrolytes [44].
SaltTransition Concentration (mol·dm−3)
(50 % Coalescence)
MgSO40.032
AlCl30.035
MgCl20.055
CaCl20.055
Na2SO40.061
LiCl0.160
NaCl0.175
NaBr0.220
KCl0.230
Table 3. Molarity-based composition of the single salts used in this study.
Table 3. Molarity-based composition of the single salts used in this study.
Water TypeCa2+
[M]
Mg2+
[M]
Na+
[M]
Cl
[M]
SO42−
[M]
NO3
[M]
CO32−
[M]
Total
Molarity
[M]
TDS
(mg/L)
Ionic Strength [M]
NaCl--0.02130.0213---0.021312450.0213
Na2SO4--0.0142-0.0071--0.007110090.0213
NaNO3--0.0213--0.0213-0.021318100.0213
CaCl20.0071--0.0124---0.00717880.0213
CaSO40.0053---0.0053--0.00537250.0213
Ca(NO3)20.0071----0.0142-0.007111650.0213
MgCl2-0.0071-0.0142---0.00716760.0213
MgSO4-0.0053--0.0053--0.00536410.0213
Mg(NO3)2-0.0071---0.0142-0.007110530.0213
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manono, M.S.; Corin, K.C. Considering Specific Ion Effects on Froth Stability in Sulfidic Cu-Ni-PGM Ore Flotation. Minerals 2022, 12, 321. https://0-doi-org.brum.beds.ac.uk/10.3390/min12030321

AMA Style

Manono MS, Corin KC. Considering Specific Ion Effects on Froth Stability in Sulfidic Cu-Ni-PGM Ore Flotation. Minerals. 2022; 12(3):321. https://0-doi-org.brum.beds.ac.uk/10.3390/min12030321

Chicago/Turabian Style

Manono, Malibongwe S., and Kirsten C. Corin. 2022. "Considering Specific Ion Effects on Froth Stability in Sulfidic Cu-Ni-PGM Ore Flotation" Minerals 12, no. 3: 321. https://0-doi-org.brum.beds.ac.uk/10.3390/min12030321

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