University
stringclasses
19 values
Text
stringlengths
458
20.7k
Virginia Tech
minutes. The receding contact angle of the hydrophobic gold surface in water is 53.5 ± 2.0o. A close fit was obtained between the experimental data and numerical results, in which the hydrophobic interaction was included in the extended DLVO theory with K = 12 x 10−18 J for 132 the estimation of the surface forces. A comparison of the results obtained at 10 minutes and at 30 minutes hydrophobization time showed that the hydrophobic force became more attractive when the gold surfaces were immersed in a 10-5 M KEX solution for a longer immersion time. It was shown in the green line that the overall total force increased slowly and became attractive when the film ruptured. Note that, the deviation was found at t > 2 s between the experimental data and the theoretical prediction. The result showed that the theoretical prediction with an inclusion of the hydrophobic interaction of a power law overestimated the overall interaction force. At t > 2 s, the film thickness was below 100 nm. It appears that the hydrophobic force of a power law might overestimate the force at a short-range distance. Figure 10.9 shows the interaction force in wetting films formed on the gold surfaces hydrophobized in the 10-5 M KEX solution for 120 minutes. At 120 minutes, the contact angle of water on the gold surface decreased slightly, and θ = 51.0 ± 1.2o. We have shown that the force r constant (K ) for the hydrophobic force decreased to 10.4 x 10−18 J at a hydrophobizaiton time 132 of 120 minutes. It is clearly shown that the use of a power law for the hydrophobic interaction might overestimate the hydrophobic interaction at short-range separation distance. However, it quantitatively predicted the surface force at a long-range distance. Figure 10.10 shows the disjoining pressure in a wetting film between an air bubble and a hydrophobic gold surface treated in a 10-5 M KEX solution for different immersion time. The black curve shows the van der Waals dispersion force (Π ) in a wetting film formed on a gold d surface with the Hamaker constant (A ) of −14.8 x 10−20 J. The curve labeled “0 min” 132 represents the disjoining pressure between an air bubble and a bare gold surface. A repulsion was found at a long-range distance, while an attraction was found at a short-range distance as predicted by the HHF theory. When the gold surfaces were rendered hydrophobic in the 10-5 M KEX solutions, the disjoining pressure became increasingly attractive. It was shown that the hydrophobic force constant (K ) in wetting films increased when the gold surfaces were 132 rendered hydrophobic in KEX solutions from 10 minutes to 30 minutes, and decreased at 120 minutes. An increase of the hydrophobic force at short hydrophobizaiton time might be attributed to a formation of the hydrophobic monolayer on gold surfaces, while a decrease of the 227
Virginia Tech
Figure 10.10 Disjoining pressure isotherms in TLFs between air bubbles and gold surfaces treated in KEX solutions for the varying hydrophobization time. The inclusion of hydrophobic force in the extended DLVO theory is used to explain the negative disjoining pressure in a wetting film between an air bubble and a hydrophobic gold surface. hydrophobic force at longer time might be attributed to the formation of a multilayers above the monolayers. The adsorption of the multilayers might render the hydrophilic head group of xanthate towards the water phase, resulting in a decrease of solid hydrophobicity. As shown from the contact angle measurement, the contact angle decreased slightly when leaving the gold surfaces in a KEX solution for a longer hydrophobization time. 10.5 Summary The interaction forces between air bubbles and gold surfaces were studied in water using the force apparatus for deformable surfaces (FADS). The new apparatus is capable of directly measuring the interaction force between a solid surface and an air bubble across a thin liquid film, while monitoring the bubble deformation during the course of the interaction. In the present work, we have studied the effect of the solid hydrophobicity on the forces acting between an air 228
Virginia Tech
Chapter 11. Conclusions and Recommendations for Future Research 11.1 Conclusions The major findings and contributions from this work may be summarized as follows. 1. When an air bubble is pressed against a flat substrate immersed in water, the bubble flattens and creates a thin liquid film (TLF) of water between the bubble and the substrate. The curvature change associated with the formation of the wetting film creates an excess pressure (p) in the film, which in turn causes the film of water to drain. The film drainage continues until the excess pressure becomes equal to the disjoining pressure () in the film. The TLF is stable when the disjoining pressure is positive (repulsive) and is unstable when the disjoining pressure is negative (attractive). One can readily determine the positive disjoining pressures when the wetting films are stable. However, no one has ever been able to measure the negative disjoining pressures as the films drain too fast to do meaningful measurements. 2. In the present work, the thin film pressure balance (TFPB) technique, originally developed for the study of foam films, has been modified to measure the negative disjoining pressures in wetting films. It is equipped with a high-speed video camera to record the interference patterns (Newton rings) of the fast-evolving wetting films. The interference patterns were then analyzed offline to reconstruct the spatial and temporal (spatiotemporal) profiles of the wetting films with a nano-scale resolution. The experimental data obtained in this manner were analyzed using the Reynolds lubrication theory to determine the changes in disjoining pressure () with time and film thickness (h). The results showed that long-range hydrophobic forces were present in the wetting 233
Virginia Tech
films formed on the gold surfaces hydrophobized by short-chain alkylxanthates. According to the thermodynamic analysis based on the Frumkin-Derjaguin isotherm, the kinetics of film thinning is expedited by the long-range hydrophobic forces, while the short-range hydrophobic forces are responsible for the rupture of the wetting film formed on a hydrophobic surface. 3. The role of collector on flotation has been studied by measuring the disjoining pressures of the wetting films with and without hydrophobization of the substrate. The results showed that collector coating increases the hydrophobicity of minerals and thereby creates a negative disjoining pressure, so that wetting films thin faster and ruptures. 4. The kinetics of film thinning has been studied by using the different sizes of the bubbles to form wetting films. The results showed that the wetting films formed with smaller air bubbles thin faster due to the larger curvature pressures. 5. The negative disjoining pressures observed in the wetting films formed on hydrophobic surfaces are the consequence of asymmetric hydrophobic interactions between the air/water and solid/water interfaces in wetting films. It has been found that the hydrophobic force constant (K ) for the asymmetric hydrophobic interactions can be 132 predicted from the hydrophobic force constant (K ) for the symmetric hydrophobic 131 interactions between hydrophobic surfaces and the symmetric hydrophobic force constant (K ) between air bubbles using the geometric mean combining rule. In view of the fact 232 that the geometric mean combining rule is used for the van der Waals force, the hydrophobic force may be considered a molecular force. 6. It has been shown in the present work that wetting films can be destabilized by the attractive double-later interactions in wetting films. In the presence of 3x10-5 M Al3+ ions, the silica/water interface is positively charged while the air/water interface is negatively charged. The double-layer interaction between the two oppositely charged surfaces created a negative disjoining pressure. As a consequence, the wetting film thinned fast 234
Virginia Tech
and ruptured with a finite contact angle. The measured contact angle is in agreement with the prediction from the Frumkin-Derjaguin isotherm. 7. It has been a challenge to directly measure the interaction forces involved in bubble- particle interactions. The main reason was that air bubbles deform during the interaction, which makes it difficult to determine the separation distances between the two macroscopic surfaces. As a result, much of the data reported in the literature are inconsistent with flotation practices. The force apparatus for deformable surfaces (FADS) developed in the present work is capable of measuring both hydrodynamic and surface forces involved in bubble-particle interactions. The system has two optical systems, one to directly measure the deflection of the cantilever spring using a fiber optic system, and the other to record the interference patterns (Newton rings) of the fast-evolving wetting films using a high-speed video camera. The former is used for direct force measurement, and the latter is used to reconstruct the spatial and temporal film profiles of the wetting films offline. 8. Analysis of the FADS data show that an air bubble approaching a solid surface deforms in response to both the hydrodynamic and surface forces in the system. In the presence of a strong repulsive disjoining pressure, the wetting film becomes flat. In the presence of a strong attractive force, a concave (pimpled) wetting film is formed when the approach speed is slow, while a convex (dimpled) wetting film is formed when the approach speed is high. An unstable, either pimpled or dimpled, film was developed under an attractive force. It has been shown in the present work that bubble-particle interactions are controlled initially by hydrodynamics, followed by surface forces. Unlike the interactions between rigid particles, the energy associated with bubbles (or other soft matter) is conserved by shape changes when subjected to an external force. 9. The direct force measurements conducted with the FADS showed that hydrophobic forces are present in wetting films. The measured hydrophobic forces increase with increasing surface hydrophobicity. This finding is consistent with the thermodynamic 235
Virginia Tech
prediction from the Frumkin-Derjaguin isotherm that a wetting film ruptures only when its disjoining pressure is negative. 11.2 Recommendation for Future Research On the basis of the present work, future research directions under the topic of the wetting film are recommended as follows. 1. FADS developed in the present work is a scientific breakthrough for the study of the TLF formed on a solid surface. However, FADS requires the upgrades for better performances. First of all, the current design is capable of measuring the force in a thin liquid film formed on a solid surface with a resolution of 5 nN. A better mechanical design with low drift and background noise is essential. Secondly, the spatiotemporal thickness profiles of the liquid films were determined from the interference fringes of the fast-evolving wetting films by means of a high-speed camera. The reconstruction of the spatiotemporal profiles was done using the monochromatic interferometry technique. However, such technique naturally lacks of the capability of determining the order of the fringes. Additionally, the use of the monochromatic interferometry is particularly limited in determining the film thickness below 20 nm. Future research will focused on developing a multi-wavelength interferometry technique to monitor the separation distance with a sub-angstrom resolution. Thirdly, FADS was specifically designed for the force measurement between an air bubble and a solid surface, with a lack of the accessories to study other soft matters, such as oil droplets, supercritical CO . A multifunctional 2 platform will be constructed with an environment chamber for dust control and pressure regulation. 2. The results in Chapter 10 showed that xanthate adsorption on gold surfaces in an open circuit created a strong hydrophobic attraction in a wetting film between an air bubble and a hydrophobic gold surface. The attraction might be the hydrophobic interaction due to a rise of the surface hydrophobicity. It has been well documented that xanthate 236
Virginia Tech
Figure A.2 Spatial profiles of the wetting films obtained at t = 1340 ms. Solid line represents a fitting curve using a six-order polynomial fitting method. where p is the curvature pressure due to the surface tension, p the hydrodynamic pressure and cur Π the disjoining pressure. By definition, the interaction force, F(t), exerting on the cantilever surface can be obtained from the integral of the hydrodynamic pressure and the disjoining pressure,    F(t) 2 p(r,t)(r,t) rdr (A.2) r0 in which r = 0 represents the symmetric axis in a cylindrical coordinate, i.e., the center of the film. By substituting eq. (B.1) to eq. (B.2), one can obtain an alternative expression for the interaction force between an air bubble and a solid surface,  F(t) 2 p (r,t)rdr (A.3) cur r0 in which the interaction force, F(t) is expressed in term of p . In a thin film, the curvature cur pressure, p , can be estimated from the profiles of the thin liquid film using the following cur equation. 2    h p   r  (A.4) cur R r r  r  239
Virginia Tech
Figure A.3 Curvature pressure vs. radial position of the film at t = 1340 ms. The profiles of the p was obtained from the numerical analysis of the thickness profiles shown cur in Fig. B.2 using the eq. (B.4). where R is the radius of the bubble in the far field, and γ is the interfacial tension of air/water interface. In eq. (B.4), the second term of p is the local Laplace pressure evaluated at radial cur position of the film. At flat film, ∂h/∂r is equivalent to zero at flat film, and thus p = 72 N/m2 cur for 2 mm radius of an air bubble. Thus, the interaction force can be determined when the spatial thickness profiles of the TLF is known. The calibration is carried out by pressing an air bubble towards a hydrophilic cantilever surface using a piezo actuator. Initially, a manual micrometer was used to lower the position of the cantilever until an equilibrium film is formed. Afterwards, the position of a bubble is controlled by means of a piezo actuator. When the piezo actuator extends by applying the voltage, the thin film expands at equilibrium film thickness, resulting an increase of the total force. Thus, the spring constant can be calibrated from a linear fit between the interaction force and the deflection of the cantilever by increasing the size of the flat film. 240
Virginia Tech
Figure A.4 A linear fit between the interaction force and the deflection. It shows that spring constant k = 3.24 N/m. Fig. (B.1) shows the deflection vs. time by manually elevating an air bubble towards a cantilever in steps. It was shown that the interaction force increased when the piezo extended, while remained constant when the piezo stopped. The film profiles were tracked simultaneously by capturing the interference fringes by means of a high-speed camera. Fig. (B.2) shows the spatial thickness profiles of the wetting film at t = 1340 ms. The profiles were fitted using a six- order polynomial fitting method. Using the eq. (B.4), the curvature pressure can be obtained, as shown in Fig. (B.3). From the integral of curvature pressure using eq. (B.3), we obtained that F = 525.5 nN. In a same manner, we determined the force at other time. Table B.1 shows the force and deflection at different time. A plot of the force vs. deflection is shown in Fig. (B.4). From a best linear fit, we obtain that the spring constant k = 3.24 N/m. 241
Virginia Tech
where the hydrodynamic pressure was obtained by integrating the velocity from the infinity to the local radial position. By integrating the hydrodynamic pressure over the entire thin film, one can obtain an expression for the hydrodynamic force, R F 2 prdr (B.5) r0 Fig. (B.1) shows the effect of approaching speed on the hydrodynamic force in a thin liquid film of water between two solid surfaces. The results were shown between a 2 mm sized particle and a flat solid surface. It was found that the numerical results covering 300 and 500 µm radii of the film area were in a good agreement with the lubrication theory, while those covering 100 µm radii of area underestimated the hydrodynamic force. Thus, all the calculations for the hydrodynamic force in this work was done by the integral of the hydrodynamic pressure over the film areas at r = 0 - 300 µm. Figure B.1 Effect of approaching velocity on the hydrodynamic force exerting on the solid surface when a 2 mm radii particle is used. The solid lines and the circle points represent the hydrodynamic force predicted using the eq. (B.1) and the numerical analysis using the eqs. (B.2)-(B.5), respectively. The hydrodynamic forces obtained from the integral of the hydrodynamic pressure over the radial distance of (a) 100 µm, (b) 300 µm and (c) 500 µm were compared. 244
Virginia Tech
Modeling Bubble Coarsening in Froth Phase from First Principles Seungwoo Park ABSTRACT Between two neighboring air bubbles in a froth (or foam), a thin liquid film (TLF) is formed. As the bubbles rise upwards, the TLFs thin initially due to the capillary pressure (p ) created by c curvature changes. As the film thicknesses (H) reach approximately 200 nm, the disjoining pressure (П) created by surface forces in the films also begins to control the film drainage rate and affect the waves motions at the air/water interfaces. If П < 0, both the film drainage and the capillary wave motion accelerate. When the TLF thins to a critical film thickness (H ), the cr amplitude of the wave motion grows suddenly and the two air/water interfaces touch each other, causing the TLF to rupture and bubbles to coalesce. In the present work, a new model that can predict H has been developed by considering cr the film drainage due to both viscous film thinning and capillary wave motion. Based on the H cr model, bubble-coarsening in a dynamic foam has been predicted by deriving the geometric relation between the thickness of the lamella film, which controls bubble-coalescence rate, and the Plateau border area, which controls liquid drainage rate. Furthermore, a model for predicting bubble-coarsening in froth (3-phase foam) has been developed by developing a film drainage model quantifying the effect of particles on p . The c parameter p is affected by the number of particles and the local capillary pressure (p ) around c c,local particles, which in turn vary with the hydrophobicity and size of the particles in the film. Assuming that films rupture at free films, the p corrected for the particles in lamella films has been used to c determine the critical rupture time (t ), at which the film thickness reaches H , using the Reynolds cr cr equation. Assuming that the number of bubbles decrease exponentially with froth height, and knowing that bubbles coalesce when film drains to a thickness H , a bubble coarsening model has cr been developed. The model predictions are in agreement with the experimental data obtained using particle of varying hydrophobicity and size.
Virginia Tech
Acknowledgement Foremost, I would like to express my sincerest gratitude to my research advisor, Dr. Roe- Hoan Yoon, for providing me an opportunity to challenge such an attractive research area and teaching me how to generate new ideas from fundamental studies. I am also grateful to Dr. Gerald Luttrell for introducing me hydrodynamic theories, Dr. Gregory Adel for teaching me how to make a professional presentation through his seminar courses, Dr. Saad Ragab for his critical comments on this dissertation, and Dr. Sunghwan Jung for his continued academic advice and valuable classes. The completion of this work would not have been possible without the support and encouragement of all of the staff members and students at Center for Advanced Separation Technologies. Particular appreciations are expressed to Gaurav Soni and Kaiwu Huang for incorporating the theoretical models developed in this study into a simulator, Dr. Lei Pan for useful discussions and innovative suggestions, and Whiusu Shim for helping me prepare samples and assisting me with image-processing. I also have been blessed with Dr. Jai-Koo Park and Dr. Sungsoo Cha, who helped me achieve career planning during their visit. I am also grateful to FLSmidth for funding continuously and providing me with necessary facilities for this project. I wish to express my sincere thanks to Prof. Siyoung Jeong for propelling me forward since I was an undergraduate. Finally, I want to acknowledge my family for their love, sacrifice, and patience. iii
Virginia Tech
Chapter 1. Introduction 1.1 Preface Froth flotation is a method for selectively separating a particulate material from one another by means of the air bubbles dispersed in liquid medium. The flotation technology has been widely used for upgrading pulverized ores in the mining industry, removing contaminants from waste water in the water treatment industry, capturing printing ink from paper fibers in the paper recycling industry, and extracting bitumen form oil sands. In the minerals separation process, the surface of desired mineral particles is rendered hydrophobic by the absorption of collector (hydrophobizing reagent), whereas that of undesirous minerals remain hydrophilic. The Gibbs energy change (∆G) associated with the bubble-particle attachment is given by the following form [1], G(cos1) (1.1) where γ is the interfacial tension between liquid and air phases and θ is the water contact angle. Eq. (1.1) indicates that only the attachment between a bubble and a particle with θ > 0° is thermodynamically favorable. Hence, only the hydrophobized particles selectively can attach to air bubbles, form particle-bubble aggregates, which rise upward due to buoyancy, and form a froth layer at the surface of the pulp phase. True recovery in the froth phase occurs when the particles attached to bubbles reaches the top of the froth and finally flow into a launder without detachment. As the particle-coated bubbles rise upward along the froth height, they coalesce with each other and become larger. As the bubbles become coarser, the bubble surface area decreases, thereby detaching desired hydrophobic particles from bubbles and causing the true recovery to decrease. However, less hydrophobic particles tend to be detached preferentially, so the froth phase contributes to enhance the grade of a froth product [2]. 1
Virginia Tech
On the other hand, particles suspended in the pulp can also be recovered without adhesion to bubbles through the entrainment mechanism. The froth recovery due to entrainment occurs when interstitial liquid between bubbles drag particles from the pulp and allow them to reach the launder. The dragging effect may stem from the agitation in the pulp (mechanical entrainment), the wake behind a rising bubble, or bubble swarm crowding effect (hydraulic entrainment). Since entrainment is non-selective, not only hydrophobic but also hydrophilic particles can be recovered, lowing the grade product [3]. As stated, flotation is a multi-phase process involving a variety of physicochemical phenomena and those are interacting each other. As a result, quantitative flotation modelling and simulations are highly required to explore the flotation science and optimize the operation of overall process. In the following section, literatures associated with the two distinct phases, pulp and froth phases, respectively, are briefly reviewed, particularly in modelling perspective. Next, the objective of the present doctoral research is presented. 1.2 Literature Review 1.2.1 Pulp A. Bubble Generation Model In a flotation cell, bubbles are generated by splitting injected air into tiny air bubbles and the splitting pressure is originated from rotational motion of a rotor. The splitting pressure is opposed by the surface tension force tending to resist deformation and minimize the surface area of the air/water interface. The ratio of splitting pressure to capillary pressure is defined as Weber number (We). Hinze [4] suggested that if We reaches a critical value, bubble breaks. The critical Weber number (We ) is obtained from the following expressions, cr Splitting pressure u2d We   2,max (1.2) cr Capillary pressure  where ρ is the liquid density, u2 is the mean square velocity difference between two points, d 2,max is maximum stable bubble size, and γ is the surface tension. The value of u2 is given by [5], 2
Virginia Tech
B. Drift-Flux Model The bubbles generated in the pulp phase of a flotation cell rise upwards due to buoyancy, and the bubble velocity relative to water, or the bubble slip velocity (U ), is expressed as [10], slip V V U  g  l (1.8) slip 1  where V and V are the superficial gas and liquid (slurry) velocities, respectively, and ε is the liquid g l volume fraction. The +/- signs indicates counter-current and-co-current flows of gas and liquid, respectively. Empirically, it was found that U is a function of the terminal rise velocity (U) of a single slip t bubble in an infinite pool and ε [11], U Um1 (1.9) slip t where m is the empirical parameter that depends on flow patterns. In using Eq. (1.9), the value of U is given by [12], t gd 2 U  2 (d 1.5 mm) (1.10) t 18(10.15Re0.687) 2 where g is the gravitational , d is the bubble diameter, μ is the liquid viscosity, and Re is the 2 Reynolds number, given by d ρU / μ . In the case of a coarse bubble1(1.5 mm ≤ d ≤ 10 mm), U 2 t 2 t is independent of d and reaches approximately 21 cm/s [13]. 2 By combining Eqs. (1.8) ~ (1.10), one can predict one of unknown parameters if the others are measurable. Banisi and Finch [14] and Ityokumbul et al. [15] used the drift-flux analysis to predict the average bubble size dispersed in the pulp in a flotation column. Stevenson et al. [16] used the theory to predict liquid overflow rate from a foam column. 1.2.2 Froth In the Section A below, the bulk motion of a froth (or a foam) in a flotation cell is described. The froth is composed of a cloud of air bubbles. Between three bubbles, a triangular channel called 4
Virginia Tech
a Plateau border (PB) is formed, through which liquid drains, as will be described in Section B below. Although liquid drainage models neglect the water presence in the thin film (lamella) formed between two bubbles, the lamella films play an importance role forth and foam stability since bubbles coalesce when the thickness of a lamella film reaches a critical value (h ). In Section c C, the thinning kinetics of lamella films is discussed. In section D, a number of methodology to quantify froth stability is reviewed. A. Bulk Motion of a Froth Based on continuum-flow approach, the bulk motion of a froth has been described by the Laplace equation [17-19], 20 (1.11) where Ψ is the stream function. As is well known, Eq. (1.11) requires incompressible and irrotational flow conditions. The incompressible condition has been assumed due to the fact that the water content in flotation froth may not be enough to compress bubbles and change the volume of them. On the other hand, the irrotational flow assumption has been made based on the premise that the geometry of a flotation cell is generally simple and thus swirling flow cannot be created. B. Foam Drainage Verbist et al. [20] calculated the liquid drainage velocity through a Plateau border (PB) based on the assumption of Poiseuille-type flow and no-slip boundary condition on air/water interface. Consider a single vertical PB with a cross-sectional area A. From the Young-Laplace equation, one can obtain the pressure difference across the PB,   pp  p  (1.12) g R g A/C PB where p is the pressure in the PB, p is the gas pressure in a bubble, and R is the radius of the g PB PB, given by A/C . Here C is the dimensionless geometric parameter. The pressure gradient along the vertical direction x is written as, 5
Virginia Tech
p     A    0.5CA3/2 (1.13)   y y R x   PB The pressure gradient should be balanced by gravity, ρg, and dissipative force, A fu 0.5CA3/2 g (1.14) x A where u is the liquid drainage velocity averaged over A and f is the PB drag coefficient, can vary with geometry of PB and boundary condition of PB wall. Hence, the drainage rate u is gA C A u   (1.15) fu 2fA1/2 x The first and the second term on the right side represent the contributions from gravity and capillary suction, respectively. Eq. (1.15) have some limitations for describing the liquid drainage in a flotation froth. In a flotation froth, air bubbles are not stationary and moves relatively to the liquid. The rising bubble, therefore, could may affect the liquid flow by dragging, but the dragging effect is neglected. Inertial effect is also neglected in Eq. (1.15), but in a froth consisting of large bubbles, inertia may become significant. Moreover, it assumes that viscous loss occurs only in the PBs (channel- dominated hypothesis). However, several researchers have recently reported that most viscous loss occurs at the node, where four PBs meet [21-24]. In a dry foam, the liquid content in the node is quite negligible compared to the same in the PB, thus the channel-dominated hypothesis may be applicable. C. Film Drainage Reynolds [25] used simplified Navier–Stokes equations with lubrication approximation to describe the squeezing liquid flow between two flat immobile surfaces. Similarly, the drainage of a lamella film can be described by the Reynolds lubrication equation. The pressure in the lamella film can be calculated by the following equation, 6
Virginia Tech
p 2u  (1.16) r z2 where p is the excess pressure in the lamella, u is the axial liquid velocity along the radial direction r, and z is the vertical coordinate. By substituting Eq. (1.16) into continuity equation, one can deduce the film thinning rate, dH 1   p  rH3   (1.17) dt 12r r r  where H is the film thickness and t is the drainage time. By considering the averaged p within the entire film, one can obtain the Reynolds lubrication equation, dH 2H3p  (1.18) dt 3R2 f where R is the film radius. Here the average pressure p is given by, f p p  (1.19) c where p is the capillary pressure and П is the disjoining pressure incorporating intermolecular c forces. The capillary pressure p is given by, c 2 p  (1.20) c R where R is the bubble radius. In calculating П, in the original model, only the contributions from the electrostatic and van der Waals forces, according to the classical Derjaguin-Landau-Verwey- Overbeek (DLVO) theory. However, in the present study, according the extended DLVO theory [26], the contribution from the hydrophobic force is considered additionally, e  A K     64C RTtanh2 s exp(H) 232  232 (1.21) el vw hp el 2 4kT  6H3 6H3 where П , П , and П represent the contribution from the electrostatic, van der Waals, el vw hp hydrophobic forces, respectively, C the electrolyte concentration, R the gas constant, T the el 2 7
Virginia Tech
absolute temperature, e the electronic charge, ψ the surface potential at the air/water interfaces, k s the Boltzmann’s constant, and κ the reciprocal Debye length. D. Froth Stability The froth (or foam) stability, by definition, is the ability of air bubbles in froth phase to resist coarsening and bursting. As is well known, froth stability has a significant effect on determining product grade and recovery in flotation. Therefore, various criteria have been proposed to quantify the froth stability. In general, they can be categorized into two subgroups, i.e. static tests and dynamic tests. In static tests, froth is freely allowed to coalesce without creating additional bubbles. Iglesias [27] used a froth decay rate as an indicator of froth stability. When froth height (h) reaches f an equilibrium, the air supply to the cell was discontinued and immediately started to measure the h as a function of time t, as follows, f dh(t) Froth decay rate = f (1.22) dt Instead of monitoring a decaying foam continuously, Zanin [28] simply measured the time need for a foam height to drop to one half of its initial height and the half-life time was used as a measure of froth stability. On the other hand, in dynamic tests, the froth stability is measured while bubbles are being generated continuously. Therefore, during the tests, the bubbles at the base of the froth are allowed to experience coalescences while moving upwards along the froth height. Due to these similarities between the dynamic tests and real flotation tests, as compared to static tests, dynamic tests may provide more reliable froth stability criterion in flotation applications [27]. Bikerman [29] was the first to propose a methodology for a dynamic test. He introduced the concept of the dynamic stability factor (Σ), which is the ratio of the maximum volume of froth to the air flowrate, h A  f,max (1.23) Q 8
Virginia Tech
where A is the cross-sectional area of a cell and Q is the volumetric air flowrate. Barbian [30] suggested the maximum froth depth h can be a froth stability indicator at equilibrium status. f,max In the case of growing froth, he monitored h as a function of time t and the froth rising velocity, f u (t) = dh(t)/dt, was used as an indicator. The model suggests that as froth becomes stable, u (t) f approaches the superficial gas velocity (V ). The air recovery has been used by several researchers g and it is defined as the fraction of air injected into a flotation cell that overflows the cell lip as unburst bubbles [31-33]. The primary limitation of the air recovery concept is that it is volume- based approach though the flotation performance relies mainly on the bubble surface area. Ata [34] measured the bubble size distribution along froth height and the froth stability was quantified in terms of the bubble growth rate. Laplante [35] presented the froth retention time (FRT), which is given by the following equation, h FRT  f (1.24) V g Eq. (1.24) can also be regarded as the measure of average life time of bubble. More recently, Hu [36] monitored the change in electrical impedance by means of a pair of electrodes immersed in froth phase. It was found that bubble coalescence or froth structure variation can vary the value of electrical impedance sensitively. Based on this finding, the degree of impedance variation was used as a froth stability criteria in his study. 1.2.3 Flotation Model The flotation as a first-order rate process can be represented as follows [37, 38], dN 1  kN  Z P (1.25) dt 1 12 where N is the number of particles in the cell at time t, k is the flotation rate constant, P is the 1 probability of flotation, and Z is the collision frequency between particle 1 and bubble 2, given 12 by the following relation [7], Z 23/21/2N N d2 u2 u2 (1.26) 12 1 2 12 1 2 9
Virginia Tech
where N is the number of bubbles on the cell, d is the collision diameter (sum of radii of bubbles 2 12 and particles), and u2 and u 2 are root mean square velocities of the particles and bubbles, 1 2 respectively. Here, the values of u2 and u 2 can be determined as presented previously by Schubert 1 2 [7] and Lee et al. [39], respectively. In using Eq. (1.25), P may be expressed in the form, PPP(1P)P (1.27) c a d f where P is the probability of collision, P the probability of attachment, P the probability of c a d detachment, and P the probability of particles surviving froth phase. f In using Eq. (1.27), P can be determined from stream functions for water around air c bubbles [40],  3 3  Re d  P c tanh2   2 1 16 10.249Re0.56    d1 2    (1.28) where Re is the Reynolds number, d and d the diameters of a particle and a particle, respectively. 1 2 In using Eq. (1.27), P can be readily calculated using the following equation [40, 41], a  E  P exp 1  (1.29) a  E    k where E is the energy barrier and E is the kinetic energy of a particle. E is equal to the maximum 1 k 1 potential energy between a bubble and a particle, which can be obtained from the extended DLVO theory and E is calculated using the following relation, k E 0.5m(U )2 (1.30) k 1 Hcr where m is the mass of a particle, and U is the velocity of a particle approaching a bubble at 1 Hcr the critical rupture thickness (H ) of the wetting film in between. The value of U is inversely cr Hcr proportional to the hydraulic resistance force against the thinning. 10
Virginia Tech
In using Eq. (1.27), P is given as follows [26], d  W E  P exp a 1 (1.31) d  E    k where W is the work of adhesion and E ’is the kinetic energy that detaches a particle from a a k bubble surface. Here W is obtained as follows, a W r2(1cos)2 (1.32) a 1 where γis the surface tension of water, r is the particle radius, and θ is the water contact angle. In 1 using Eq. (1.31), E’ may be found by the following relation [26], k   2 E' 0.5m (d d ) /v (1.33) K 1 1 2 where  is the energy dissipation rate and  is the kinematic viscosity. Once particles enter the froth phase, only a part of them survives the froth phase and reaches the launder. In determining P using Eq. (1.27), the probability of particles surviving froth phase (P) is given as [42, 43], f P P P R exp()R exp(0.03250.063d) (1.34) f fa fe max f FW,max 1 in which P and P are the froth recovery due to attachment and entrainment, respectively, R fa fa max the maximum recovery factor, α fitting parameter, τ the froth retention time, R themaximum f FW,max feedwater recovery to froth, and ∆ρ the specific gravity difference between particle and water. Here R is given by, max S     d R  t  6V /d / 6V /d  2,b (1.35) max S g 2,t g 2,b d b 2,t where S and S are the surface area rates of the bubbles at the top and bottom of the froth phase; t b d and d are bubble diameters at the top and bottom of the froth phase, respectively; and V is 2,t 2,b g the superficial gas velocity. 11
Virginia Tech
1.3 Research Objective As shown in Eqs. (1.34) and (1.35), the froth recovery depends strongly on the bubble- coarsening factor (d /d ), which is the bubble size ratio between the bottom and the top of a froth 2,b 2,t phase. As a result, it is of critical importance to understand the mechanisms of bubble-coarsening phenomena in a flotation froth. In determining the bubble-coarsening factor, the bubble size at the bottom of a froth (d ) may be considered to be the same as the bubble size in the pulp phase, 2,b which can be readily predicted from bubble generation model. However, at present, there are no models that can predict the bubble size at the top of a froth (d ) from d . The primary objective 2,t 2,b of the present study is, therefore, to develop a model for predicting the bubble-coarsening in a froth phase from first principles. For this to be possible, it is essential to know the critical thickness (H ) of lamella films at cr which two neighboring bubbles in a froth coalesce and bubble-coarsening occurs. Therefore, the present work is also aimed to develop a predictive model for the value of H . cr 1.4 Organization The body of this dissertation consists of six chapters. Chapter 1 provides a comprehensive review of flotation models. Some of them are used to develop theoretical models presented in the following chapters. This chapter also introduces the goal of the present work. Chapter 2 introduces a new predictive model for the critical rupture thickness (H ) of a cr foam film. The model has been developed on the base of the capillary wave model and the extended Derjaguin-Landau-Verwey-Overbeek (DLVO) theory. By considering the contribution from hydrophobic disjoining pressure, this model can predict H reasonable well. cr Chapter 3 describes a new model for predicting the bubble-coarsening in a dynamic foam as a function of aeration rate, foam height, and surface tension (frother dosage). This foam model is on the base of the H model developed in Chapter 2. The model has been validated using a foam cr column equipped with a high-speed camera. 12
Virginia Tech
Chapter 4 presents a new first principle froth model for predicting the bubble-coarsening in a froth (3-phase foam) as a function of aeration rate, froth height, surface tension (frother dosage), particle hydrophobicity (collector dosage), and particle size. This chapter is focused on the effect of particle hydrophobicity on bubble-coarsening. This model has been validated in a series of flotation tests using monosized spheres of varying hydrophobicity. Chapter 5 is the same as Chapter 4 except that it is mainly focused on the effect of particle size on bubble-coarsening. The froth model developed in Chapter 4 is capable of predicting the bubble-coarsening as a function of particle size. The model predictions are in good agreement with the changes in bubble sizes measured experimentally in the presence of different sizes of particles. Chapter 6 summarizes the key findings and accomplishments presented in the foregoing chapters and suggests future research topics. References [1] Laskowski, J., The relationship between floatability and hydrophobicity, in Advances in mineral processing: A half century of progress in application of theory and practive. 1986. [2] Seaman, D.R., E.V. Manlapig, and J.P. Franzidis, Selective transport of attached particles across the pulp–froth interface. Minerals Engineering, 2006. 19(6–8): p. 841-851. [3] George, P., A. Nguyen, and G. Jameson, Assessment of true flotation and entrainment in the flotation of submicron particles by fine bubbles. Minerals Engineering, 2004. 17(7): p. 847-853. [4] Hinze, J.O., Fundamentals of the hydrodynamic mechanism of splitting in dispersion processes. AIChE Journal, 1955. 1(3): p. 289-295. [5] Batchelor, G. Pressure fluctuations in isotropic turbulence. in Mathematical Proceedings of the Cambridge Philosophical Society. 1951. Cambridge Univ Press. [6] Schulze, H.-J., Physico-Chemical Elementary Processes in Flotation: Analysis from the Point of View of Colloid Science Including Process Engineering Considerations 1984: Elsevier Science Ltd [7] Schubert, H., On the turbulence-controlled microprocesses in flotation machines. International Journal of Mineral Processing, 1999. 56(1–4): p. 257-276. 13
Virginia Tech
Chapter 2. Prediction of the Critical Rupture Thickness of Foam Films ABSTRACT In flotation, a froth recovery depends critically on bubble-coarsening. The bubble- coarsening occurs when a thin liquid film (TLF) confined between two bubbles breaks. As a precursor to developing a model that can predict the bubble-coarsening in a froth (or foam), we have developed a model for predicting the critical film thickness (H ) at which a foam film cr ruptures. The model has been derived by considering the film drainage due to viscous film thinning and the capillary wave motion at air/water interfaces. This approach is based on the premise that if the disjoining pressure (П) created by surface forces in the TLF is negative (attractive), both the film drainage and the capillary wave motion accelerate and consequently the TLF ruptures. The model predictions are consistent with the H values measured experimentally. It has been found cr in the present study that at a relatively low frother (or surfactant) concentration, corrugation of air/water interfaces grow faster in amplitude and thereby the TLF ruptures at a larger film thickness. At a high frother concentration, hydrophobic force is dampened, causing the grown of fluctuation to decrease and hence causing the H to decrease. cr 17
Virginia Tech
2.1 Introduction In froth (or foam) phase, two air bubbles are in close contact with each other and a thin liquid film (TLF) is formed in between. As the bubbles rise along the height of a froth phase, the water in the TLF drains initially due to the capillary pressure (p ) created from the changes in c curvature. As the film thickness (H) reaches approximately 150-200 nm, the process begins to slow down due to the repulsive electrical double-layer forces between the two air/water interfaces facing each other. When the p becomes equal in magnitude to the disjoining pressure ( ) due to c e double-layer repulsion, the film thinning stops at an equilibrium film thickness (H ). If there exists e an attractive force in the TLF, the film thinning continues and the film ruptures and the two bubbles become one at a critical film rupture thickness (H ). cr Scheludko [1, 2] was the first to present a theoretical model to predict H . The model was cr derived based on the premise that the air/water interfaces of TLFs are thermally fluctuated and that the TLF becomes unstable and ruptures if the negative disjoining pressure gradient along film thickness H exceeds the capillary pressure gradient along H, A 2 H  232 (2.1) cr 128 where A is the Hamaker constant of water between two air bubbles and  is the wavelength of 232 the surface fluctuations. The primary limitation of this model is that the value of  is undefined. Based on Scheludko’s theory [1, 2], Vrij and his co-workers [3-5] developed a more advanced H model in that not only the model defined the values of  by calculating the Gibbs cr energy change associated with fluctuation growth, but also it included the role of film drainage in determining film stability. The following is the H model developed by the authors, cr A 2R2 H 0.207 232 f (2.2) 7 cr p c where R is the radius of contact area and γ is the surface tension. Unfortunately, Eq. (2.2) fails to f predict H accurately, particularly at low surfactant (frother) concentrations, where air bubbles cr become more hydrophobic [6]. The failure of Eq. (2.2) may arise from the fact that at the time the 18
Virginia Tech
classical Derjaguin-Landau-Verwey-Overbeek (DLVO) theory was developed and no one knew that air bubbles in water are hydrophobic. Further, no one thought about the possibility of hydrophobic force playing a role in foam stability. Recently, it has shown that the hydrophobic force is also present in foam films in addition to the van der Waals force, and that the former is larger than the latter at low frother concentrations [7, 8]. More recently, Do [9] modified Eq. (2.2) by considering the possibility that the TLF between air bubbles drains faster due to slip on the hydrophobic air/water interfaces. In addition, he introduced the variation of film thinning velocity from the difference between the wavy film radius and the flat film radius by considering the geometric relation, 2 H  18 4K 232 1 3  1 42 /2 1 1 42 /2 2 sin2d3 cr   3 A    0  42 /2 1   232   (2.3)   A 2R2 1/7  0.207  232 f      p    c where K is the hydrophobic force constant of water between two air bubbles, η is the amplitude 232 of the wave at the air/water interfaces in a foam film, and λ is the wavelength. However, in this model, the possibility that the amplitude of capillary waves grows faster in the presence of stronger attractive hydrophobic force was not considered. Moreover, this model describes the film thinning process using the classical DLVO theory, which neglects the contributions from the hydrophobic force. Also, the fact that value of / in Eq. (2.3) needs to be empirically determined is problematic. In the present work, we have derived a model for predicting H by incorporating the cr hydrophobic force in the capillary wave model developed by Vrij. This new model has been developed by considering the effects of hydrophobic force on the film drainage and the wave motion. 2.2 Model Incorporating Hydrophobic Force The capillary wave model developed by Vrij et al. [3-5] described the film thinning process and the capillary wave motion to predict the critical rupture thickness (H ) of a foam film. In cr 19
Virginia Tech
Section 2.2, we incorporate the hydrophobic force into the Vrij el al.’s model to predict H at low cr frother concentrations. 2.2.1 Film Thinning The kinetics of thinning of a horizontal thin liquid film (TLF) between two air bubbles can be described in view of the Reynolds lubrication equation, dH 2H3p  (2.4) dt 3R2 drain f where H is the TLF thickness, t the drainage time, μ the dynamic viscosity, R the film radius, drain f and p the driving force for TLF thinning. In Eq. (2.4), the driving force p is given by the expression, p p  (2.5) c where p is the capillary pressure and П is the disjoining pressure. The capillary pressure p is c c given by, 2 p  (2.6) c R where R is the bubble radius and γ is the surface tension of water. In calculating П, in the original Vrij’s model, only the contributions from the electrostatic and van der Waals forces are included, according to the classical DLVO theory. However, in the present study, according the extended DLVO theory, the contribution from the hydrophobic force is considered additionally, e  A K     64C RTtanh2 s exp(H) 232  232 (2.7) el vw hp el 2 4kT  6H3 6H3 where П , П , and П represent the contribution from the electrostatic, van der Waals, el vw hp hydrophobic forces, respectively, C the electrolyte concentration, R the gas constant, T the el 2 absolute temperature, e the electronic charge, ψ the surface potential at the air/water interfaces, k s the Boltzmann’s constant, and κ the reciprocal Debye length. 20
Virginia Tech
2.2.2 Growth of Surface Fluctuation Figure 2.1 shows one of corrugating components of air/water interfaces of the TLF due to thermal motion. Corrugation due to thermal motion increases the surface area of the air/water interface and hence the free energy. However, the corrugation causes the film thickness at the valleys to decrease as the inverse curve of the film thickness due to the van der Waals attraction, which in turn causes the free energy to decrease and the film thinning kinetics to accelerate. By considering film thinning due to the growth of fluctuation, Vrij estimated the time it takes for a film to thin due to fluctuation (t ) as follows, fluct 2 d  t 24fH3  (2.8)   fluct dH  HH where f is the adjustable parameter. f = 6 was assumed in the original model. We also use the same value in the present study. In using Eq. (2.9), we recognize the contributions from П in addition hp to П and П , as presented in Eq. (2.7). el vw Figure 2.1 One of the capillary wavy patterns at the facing air/water interfaces confining the foam film with the average film thickness H and the local thickness h.  represents the wavelength and  represents the amplitude of the wave motion. 21
Virginia Tech
2.2.3 Calculation of H cr Vrij assumed that a TLF ruptures when two surfaces touch each other. This condition can be represented by the following relation, dt dt drain  fluct 0 (2.9) dH dH HHm HHm where H is the mean film thickness at which satisfies the above requirement. m Substituting Eqs. (2.4) and (2.8) into Eq. (2.9), one can obtain H as follows, m   2  3 2  3H 3R2 1443H 4  2H 3   m f 0 (2.10)       m H HHm m H HHm H2 HHm  2(p c ) Vrij recognized the fact that the critical film rupture thickness (H ) may be slightly smaller than cr H due to the fact that the viscous film thinning continues while the wave is growing. Accordingly, m he suggested the following relation, H 0.845H (2.11) cr m 2.3 Model Validation To validate the model developed in the present work, the model predictions were compared with experimental H values reported by Wang [10]. Wang measured H values of the foam films cr cr stabilized with methyl isobutyl carbinol (MIBC) in the presence of 0.1 M NaCl, where П ≈ 0 due el to double-layer compression. In his study, foam films of small film radius (R < 50 μm) were tested f to avoid dimple formation. In the present work, H values have been predicted from Eqs. (2.10) and (2.11). In using cr them, the values of K were obtained from Wang and Yoon’s results [11]. They determined the 232 K values at different concentrations of MIBC by fitting experimental film thinning data to the 232 Reynolds equation coupled with the extended DLVO theory. The values of A γ, and ψ were 232, s used as reported by Wang [10], Comley et al. [12], and Srinivas et al. [13], respectively. Only the 22
Virginia Tech
As shown in Figure 2.2 (a), there is an excellent agreement between the experimental (filled circles) and the model predictions (line). The model include the contributions from the hydrophobic disjoining pressure, which is the reason for the excellent fit between theory and experiment. If the hydrophobic force is not considered, the model predicts substantially smaller H values as shown by the bottom most (green) line in Figure 2.2 (a). cr Figure 2.2 (b) shows the H values (the topmost blue line) predicted using Eqs. (2.10) and cr (2.11) at different concentrations of MIBC in the absence of NaCl. As shown, the model gives substantially higher H values than those obtained at 0.1 M NaCl, which can be attributed to the cr fact that the hydrophobic force (or K ) increases with decreasing NaCl concentration. It is 232 believed that the hydrophobic force originating from the cohesive energy of water is compromised in the presence of electrolytes. 2.4 Summary and Conclusions In this study, we have developed a predictive model for the critical rupture thickness (H ) cr of foam films. This new model has been derived on the basis of the capillary wave model and the extended DLVO theory, which includes the contributions from the attractive hydrophobic force. It has been found in the present work that at lower frother concentration, where the hydrophobic force is stronger, a foam film ruptures at higher film thickness due to faster growth rate of surface wave. It has also been found that H decreases with increasing frother concentration, which may cr be attributed to a decrease in the hydrophobic force, resulting in retarded wave motion. As is well known, bubble-coarsening in a froth (or foam) occurs when thin liquid film confined between two bubbles ruptures. This H model will be useful for developing a model describing bubble- cr coarsening phenomena in a froth. References [1] Scheludko, A., Sur certaines particularités des lames mousseuses. Proc. Konikl. Ned. Akad. Wet. B, 1962. 65: p. 86-99. [2] Sheludko, A., Thin liquid films. Advances in Colloid and Interface Science, 1967. 1(4): p. 391-464. 24
Virginia Tech
Chapter 3. Development of a Bubble-Coarsening Model in a Dynamic Foam ABSTRACT In a flotation froth (and foam), air bubbles become larger due to coalescence, causing bubble size to increase, bubble surface area to decrease, and hence causing less hydrophobic particles to drop off to the pulp phase below. Thus, bubble coarsening provides an important mechanism by which product grades are increased. On the other hand, excessive bubble coarsening results in low recoveries. In the present work, a model describing the process of bubbles becoming coarser in a foam as they rise to the top by deriving a mathematical relation between the Plateau border area, which controls liquid drainage rate, and the lamella film thickness, which controls bubble-coalescence rate. The model developed in the present work can predict the bubble size ratio between the top and bottom of a foam as functions of aeration rate, foam height, and surface tension (frother dosage). The model predictions are in good agreement with the changes in bubble sizes measured using a high-speed camera. 26
Virginia Tech
3.1 Introduction Froth flotation is the most widely used method of upgrading pulverized ores in the minerals industry. In this process, hydrophobic particles selectively attach to air bubbles in a pulp phase, forming particle-bubble aggregates, which rise upward due to reduced buoyancy and form a froth phase. In the froth phase, the bubbles coalesce with each other and become larger as they rise. As the bubbles become larger, the bubble surface area decreases, thereby restricting the number of hydrophobic particles that can be carried upward and subsequently flow into a launder. Therefore, in flotation modelling, froth recovery due to attachment (Ra) is given as a function of surface area f as follows, R a  R exp( ) (3.1) f max f,p where R is the maximum froth recovery factor representing the carrying capacity limit, α is a max fitting parameter, τ is the retention time of particles in the froth phase. In using Eq. (3.1), R f,p max can be expressed as follows, S 6V /d d R  t  g 2,t  2,b (3.2) max S 6V /d d b g 2,b 2,t where S and S is the bubble surface area flux at the top and base of a froth, respectively, d and t b 2,t d are the corresponding bubble sizes, and V is the superficial gas velocity. 2,b g On the other hand, the bubble coarsening helps increase the grade of a froth product, as less hydrophobic particles selectively drop off the coarsening bubbles [1]. Thus, it is of critical importance to understand and control bubble coarsening in the froth phase. Despite of its significance, the bubble coarsening in the froth phase has not been studied widely until now, due to the structural complexity and the opaqueness of a mineral-coated froth. Recently, Barbian et al. [2, 3] conducted dynamic froth stability tests in a flotation column and observed experimentally that increasing gas flow rate benefits the froth stability. Tao et al. [4] found that lowering froth height help increase froth stability by measuring the water recovery in a flotation column, as a measure of froth stability. Wang and Yoon [5] reported that the disjoining pressure (П) created by surface forces in the lamella films controls the foam stability. More recently, the first theoretical 27
Virginia Tech
model that can predict bubble-coarsening or bubble size distribution in foam has been developed by coupling a liquid drainage model with a population balance model [6]. The authors employed a single fitting parameter associated with the frequency of film rupture and assumed that it is a function of fluid viscosity and surface chemistry. The model was validated by a single experiment in a surfactant solution of a fixed concentration. In the present work, we have developed a bubble-coarsening model that can predict the bubble coarsening (d /d ) by deriving a mathematical relation between the Plateau border area, 2,t 2,b which controls liquid drainage rate, and the lamella film thickness, which controls bubble- coalescence rate. The model is based on the capillary wave model [7, 8], in which air/water interfaces in form films oscillate due to thermal motions such that the instantaneous thickness of a foam film becomes smaller than the average thickness. Because the van der Waals force varies as H-3, where H is film thickness, a small change in film thickness can greatly decrease the free energy of film rupture. The bubble-coarsening model developed in the present work is also based on the recognition that the air bubbles dispersed in water are hydrophobic [9], which will in turn increase drainage rate and hence facilitate bubble coarsening. The objective of the present work is to develop a model for predicting bubble-coarsening in foams and verify it in experiment. We have measured the bubble size ratio (d /d ) of a foam 2,t 2,b stabilized with Methyl isobutyl carbinol (MIBC) at varying frother concentrations, foam heights, and gas rates. 3.2 Model Development As is well known, bubbles become larger as they rise along the height of a foam mainly due to bubble coalescence. The bubble coarsening can be controlled by frother addition. Thermodynamically, frother decrease the free energy of bubble coalescence, i.e., G = -2, by c decreasing the surface tension, , at the air/water interface. Kinetically, frother decreases the rate of film thinning by decreasing capillary pressure, P = 2/R, where R is the Plateau boarder radius. It has been shown also that frother dampens the hydrophobic force in the lamellar (or foam) films, which has been shown to accelerate the rate of film thinning at film thicknesses (H) below approximately 250 nm [10]. As film thinning continues, H reaches a critical thickness (H ), where cr 28
Virginia Tech
the film ruptures catastrophically. Thus, it is necessary to model the kinetics of film thinning at the Plateau border and predict H for developing a mathematical for bubble coarsening. cr In a dry foam, when two bubbles face each other, they form a lamella, a thin liquid film (TLF). When three TLFs meet together, they form a Plateau border (PB). A PB is a channel in which liquid flows though. Four PB forms a vertex, where liquid meet together. In the absence of external water supply to a foam (e.g. wash water in column flotation), the base of the foam is relatively wet, whereas the top of the foam is dry due to the downward liquid drainage in the foam phase. In general, water amount in the TLF is small compared to the same in the PB, especially in the case of dry area. Therefore, for the modeling purpose of a foam or liquid drainage in the foam, most studies have been neglected water in the TLF. 3.2.1 Foam Drainage Now consider liquid drainage in a PB in the vertical direction. On the base of consideration of the force balance and the mass conservation, the drainage of liquid through a PB can be derived as follows [11, 12], 1 1 A U  gA  (3.3)  A x where U is the drainage rate; g is the gravitational acceleration; µ, ρ, and γ are the kinematic viscosity, density, and surface tension of water, respectively; A is the cross-sectional area of the Plateau border (PB), and x is the distance from the top of a foam. For the force balance, the balance between gravitational force (forcing downward), capillary force (forcing upward), and viscous force (forcing opposing direction of the liquid flow) is considered. In terms of viscous loss, there are two controversial hypotheses. 3.2.2 Steady State At a steady state, the downward liquid drainage velocity (U) is equal to the upward superficial gas velocity (-V ), g U V (3.4) g 29
Virginia Tech
Combining Eqs. (3.3) and (3.4) and then integrating from the top to the bottom of a foam, one obtains, V    gA  h gV A  g tantan1 b  f g (3.5) t g   V    2    g  where A is the PB area at the top of a foam and A is the same at the bottom, and h is the foam t b f height. 3.2.3 Bubble Coarsening Model In a dry foam, each PB is formed by the intersection of three lamellar films, and each corner of the surrounding polyhedron is the intersection of four PBs. Thus, it would be reasonable to assume that the number of PBs (N ) is proportional to the number of bubbles, which can be pb calculated by dividing the cross-sectional area (S) of a foam (or froth) column by bubble diameter, i.e., 4S/πd2. One can then write the following relation, N 4S/d 2 d 2 pb,t  2,t  2,b  (3.6) N 4S/d 2  d  pb,b 2,b  2,t  where N and N are the PB numbers at the top and bottom of a foam, respectively, and d and pb,t pb,b 2,t d are the corresponding bubble sizes. 2,b As a foam drains, A becomes smaller, or the foam becomes drier. As a foam becomes drier, the thickness (H) of the lamella films will also become thinner. As H becomes thinner, there will be a point where, a lamellar film will rupture catastrophically, which is referred to as critical rupture thickness (H ). Likewise, it may be reasonable to assume that bubbles begin to coalesce cr when A reaches A , provided that there is a mathematical relation between A and H . As bubbles cr cr cr coalesce, N will decrease. In the present work, N may be related to A as follows [11], pb pb cr  A  N  N expC cr  (3.7) pb 0  A    30
Virginia Tech
where N is the number of PBs at the base of a foam, and C is an adjustable parameter. Eq. (3.7) 0 suggests that N increases exponentially with the square root of A , which can be related to H pb cr cr analytically as will be shown later. Substituting Eq. (3.7) into Eq. (3.6), one obtains the following relation, d  1 1  2,b  expC A    (3.8) d cr  A A  2,t  b t  which will allow prediction of bubble size ratio (or bubble coarsening) if the values of A , A and b t A are known. cr According to Eq. (3.5), one can determine the value of A if A is known. One can determine t b A by considering the geometrical relationships between bubble size, lamellar film thickness, and b Plateau border radius [13, 14], A  0.124(d  )2 1.52d  H (3.9) b 2,b b 2,b b b where d is the bubble size at the base of a foam (or froth), which may be considered the same as 2,b the bubble size in the pulp phase; ε is the liquid fraction at the base of the foam under b consideration; and H is the lamella film thickness at the base. b One can determine d using a simple bubble generation model [15], 2,b 0.6 2.11 d   (3.10) 2,b  0.66   bg  where ε is the energy dissipation rate in the bubble generation zone of a flotation cell. bg In using Eq. (3.9), one can obtain ε using a drift-flux analysis [16], b 31
Virginia Tech
Figure 3.1 A model for the unit cell of the bubbles packed at the base of a foam (or froth). V V gd  2 U  g  l  2 b (3.11) slip 1  18(10.15Re0.687) b b where U is the slip velocity of bubbles, V and V are the superficial velocities of gas and liquid, slip g l respectively, and Re is the Reynolds number. By the geometrical considerations depicted in Figure 3.1, the H of Eq. (3.9) was b determined combining Eqs. (3.12) and (3.13), 3 4 d   2,b  N 3   2   2,b  1 (3.12) b S(d H ) 2,b b where N is the number of bubbles of size d in a foam column of cross-sectional area of S. 2,b 2,b Assuming that the bubbles at the base are spherical and form a two-dimensional monolayer as depicted in Figure 3.1, one calculate N as follows, 2,b  S  N 2N 2  (3.13) 2,b t,b  0.5(d H )2cos30  2,b b  32
Virginia Tech
Figure 3.2 A geometrical relationship between bubble size (R ), Plateau boarder radius 2 (R ), the critical Plateau boarder area (A ), and the critical lamella film pb cr thickness (H ) in a foam. cr Since each triangular unit contains one half of a bubble, N is twice the number of the triangles 2,b at the base (N ), which can be determined simply by dividing S with the area of a triangle t,b In calculating the bubble size ratio using Eq. (3.8), it is also necessary to know the value of A . In the present work, the values of A were calculated from those of H based on the cr cr cr geometric relation between A and H [14], as shown in Figure 3.2.  A ( 3 )R 2  3R H (3.14) cr 2 pb pb cr where R is the radius of curvature of PB. In Eq. (3.14), H can be determined from the pb cr methodology presented in Chapter 2. 3.3 Experimental 3.3.1 Materials In this study, the model presented in the foregoing section was validated using a specially designed foam column shown in Figure 3.3. The column (15cm wide × 15cm deep × 47 cm high) 33
Virginia Tech
All experiments was conducted using a Millipore pure water with a resistivity of 18.2MΩ·cm at 25°C. Methyl isobutyl carbinol (MIBC, 98% purity, Aldrich) was chosen as a frother since it is the most widely used in industry [17]. Prior to each test, the column was placed in a base solution (distilled water, hydrogen peroxide and, ammonium hydroxide at a ratio of 4:1:1) for 10 min. to clean the surface of the glass wall. The column was then thoroughly rinsed with distilled water and allowed to be dried. 3.3.2 Experimental Procedure The cell was first filled with aqueous solutions of methyl isobutyl carbinol (MIBC). Bubbles were generated by pumping the nitrogen gas through an air sparger located at the bottom of the foam column. The flow rate of gas was accurately controlled by adjusting the valve connected to a flow meter (Aalborg). The foam generated in this manner was allowed to freely overflow, mimicking flotation. The overflowed solution was recycled continually. The recirculation help keep the amount of MIBC in the solution, thus the value of surface tension may be kept. The foam height was monitored using a graduated scale embedded on the front wall of the column and was manually controlled by adjusting the level of pulp/froth interface. Behind the column, a 250 W light source was placed to illuminate the observation region with plastic board to obtain clear images. Once a steady state condition was reached, the images of the bubbles in a foam column were recorded by means of a high-speed camera (Fastec imaging). 3.3.3 Bubble Size Measurement The average bubble diameters at the base of the froth (d ) and the same at the top (d ) 2,b 2,b were then calculated by analyzing the images offline using the BubbleSEdit, image-analysis software. The average bubble diameters were given by Sauter mean diameter (d ), 32 nd3 d  i i (3.15) 32 nd2 i i where n is the number of bubbles with diameter d. In measuring the average bubble size in a foam i i or a froth, d is widely used since it can represent the average suitably [18] and also the flotation 32 35
Virginia Tech
Figure 3.4 Bubble size ratio (d /d )between the top and bottom of a foam as a function 2,t 2,b MIBC concentration at different gas rates at a foam height of 4 cm. The lines drawn through the experimental data points represent the model predictions obtained using Eq. (3.7) on the basis of the critical film rupture thicknesses (H ) cr calculated from Eqs. (2.11) and (2.12). rate is critically related with surface area of bubbles [19]. In the study, the values of bubble size ratio (d /d ) were used as a measure of froth stability. 2,t 2,b 3.4 Results and Discussion 3.4.1 Effect of Frother Concentrations Figure 3.4 shows the effect of frother (MIBC) concentration on bubble size ratio (d /d ), 2,t 2,b a measure of bubble coarsening. While the foam height was kept constant at 4 cm, the experiments were conducted at 51, 71, and 102 ppm MIBC, which are much higher than industrial flotation dosage ranging between 1 and 10 ppm [17]. Such a high dosage was employed because at lower 36
Virginia Tech
concentrations, stable foam phase was not generated, mostly due to the absence of particles. It is generally known that particles are a solid-state surfactant that can enhance the form stability [20, 21]. The values of the C parameter were obtained by fitting the experimental data to the model and summarized in Table 3.1. The fit between the experimental data and simulation results (curves) were excellent. Both the model prediction and experimental data showed that the bubble coarsening decreased with increasing MIBC concentration, indicating that the frother stabilized the foam and retarded bubble coarsening. In effect, MIBC caused a decrease in A or H . cr cr As suggested previously, the hydrophobic force present in foam films may be the major driving force for foam film rupture and that the hydrophobic force decreases with increasing frother concentration [9, 22]. Therefore, the decrease in hydrophobic force with increasing MIBC concentration can provide an explanation for the decrease in both d /d and H . 2,t 2,b cr 3.4.2 Effect of Foam Height Figure 3.5 compares the values of d /d at the foam heights of 2 and 4 cm. As expected, 2,t 2,b it was observed that the bubble size ratio increased with increasing foam height at a given superficial gas rate (V ). The reason is simply that bubbles will have a longer time to coalesce in a g taller froth height. The experiments were conducted at three different gas rates, i.e., V = 2, 3, and g 4 cm/s. At a higher gas rates, the froth/pulp interface was lost due to flooding [23]. As shown, the bubble size ratio decreased most probably due to shorter residence times of bubbles [4]. Table 3.1 Values of fitting parameter C of Eqs. (3.7) and (3.8) used to predict bubble coarsening. Superficial Gas MIBC (ppm) Rate (cm/s) 51 71 102 2 16.60 10.01 3.33 3 8.44 5.75 1.65 4 5.30 2.29 0.36 37
Virginia Tech
3.4.3 Effect of Gas Flow Rate As shown in Figure 3.5, the experiments were conducted at three different gas flow rates, i.e., V = 2, 3, and 4 cm/s. A lower gas rate was not enough to create a foam phase. On the other g hand, at a higher gas rates, the froth/pulp interface was lost due to flooding [23]. As shown, the bubble size ratio decreased with increasing gas rate. It may be due to the less chances for the bubble to be coalesce, most probably due to shorter residence times of bubbles in the foam phase [4]. 3.5 Summary and Conclusions A bubble-coarsening foam model has been developed. It is capable of predicting the bubble size at the top of a foam from the bubble size at the bottom. The input parameters of the bubble- coarsening model include froth height, frother dosage, gas flow rate, and the bubble size in the pulp phase. The model predictions show that bubble coarsening increases with decreasing frother dosage, which may be attributed to the increased hydrophobic force at lower frother dosages. It is also shown that bubble coarsening increases with increasing foam height and decreasing aeration rate, which can be attributed to longer residence times of the bubbles in the foam (or froth) phase. The model predictions are consistent with the experimental results. At this point, however, we are not considering the effects of particle size and particle hydrophobicity, which is known to affect froth stability [20, 21]. Thus, the model presented in this communication is for foam rather than for froth. References [1] Schwarz, S., et al., Water behaviour within froths during flotation. 2001. [2] Barbian, N., K. Hadler, and J.J. Cilliers, The froth stability column: Measuring froth stability at an industrial scale. Minerals Engineering, 2006. 19(6–8): p. 713-718. [3] Barbian, N., et al., The froth stability column: linking froth stability and flotation performance. Minerals Engineering, 2005. 18(3): p. 317-324. [4] Tao, D., G.H. Luttrell, and R.H. Yoon, A parametric study of froth stability and its effect on column flotation of fine particles. International Journal of Mineral Processing, 2000. 59(1): p. 25-43. 39
Virginia Tech
Chapter 4. Modeling Froth Stability: Effect of Particle Hydrophobicity ABSTRACT As bubbles rise in a froth phase, a thin liquid film (TLF) confined between bubbles thins by the capillary pressure (p ) and the disjoining pressure (П). Once the intervening TLF ruptures, c bubble-coarsening occurs. In the present work, we have studied the effect of particle hydrophobicity (or water contact angle θ) on the bubble-coarsening (or froth stability). The study was conducted by measuring the bubble size ratio between the top and bottom of a forth in the presence of monosized particles of varying hydrophobicity. The experimental results showed that the froth stability increases with θ up to θ = 70° and decreases with further increase in θ. We have also developed a model for predicting the bubble-coarsening in a froth by deriving a film drainage model quantifying the effect of θ on p . The model shows that with increasing θ up to θ = 70°, p c c may decrease and thereby the film thinning is decelerated and the froth becomes more stable. At θ > 70°, however, the froth becomes less stable due to increased p and, hence, increased drainage c rate. 42
Virginia Tech
4.1. Introduction Froth flotation is the most widely used method of upgrading pulverized ores in the minerals industry [1]. In this process, hydrophobic particles selectively attach to air bubbles in a pulp phase, forming particle-bubble aggregates, which subsequently rise upward due to reduced buoyancy and form a froth phase. In the froth phase, the bubbles coalesce with each other and become larger as they rise. As the bubbles become larger, the bubble surface area decreases, thereby restricting the number of hydrophobic particles that can be carried upward and subsequently flow into a launder. On the other hand, the bubble coarsening helps increase the grade of a froth product, as less hydrophobic particles selectively drop off the coarsening bubbles [2]. Thus, it is of critical importance to study and control bubble coarsening in the froth phase. It is known that bubble coarsening occurs when a thin liquid film (TLF) formed between two bubbles in a froth phase breaks as a result of thinning. The kinetics of thinning of a TLF can be described by the Reynolds lubrication equation [3], dH 2H3p  (4.1) dt 3R 2 f where H is the TLF thickness, t the drainage time, μ the dynamic viscosity, R the film radius, and f p the driving force for TLF thinning. The equation is derived based on the Navier-Stokes equations under the assumptions of plane-parallel films and no-slip boundary conditions at the air/water interfaces. In using Eq. (4.1), the driving force is given by the following relation, p p  (4.2) c which shows that the driving force is the sum of capillary (or curvature) pressure (p ) and c disjoining pressure (П). During the initial film thinning, the drainage of a TLF is governed by the p whose c magnitude is given by the Laplace equation [4], 43
Virginia Tech
2 p  (4.3) c R where R is the bubble radius and γ is the surface tension of water. When H reaches 200 nm, surface forces or disjoining pressure (П) between two air/water interfaces begin to act and control the thinning of the TLF. In Eq. (4.2), П is determined according to the extended DLVO theory [5], e  A K     64C RTtanh2 s exp(H) 232  232 (4.4) el vw hp el 2 4kT  6H3 6H3 where П is the disjoining pressure due to electrostatic interaction, П is the disjoining pressure el vw due to the van der Waals dispersion force, П is the disjoining pressure due to hydrophobic force, hp C is the electrolyte concentration, R the gas constant, T the absolute temperature, e the electronic el 2 charge (e = 1.6 ×10-19 C), ψ the surface potential at the air/water interfaces, k the Boltzmann’s s constant, κ the reciprocal Debye length, A the Hamaker constant, and K the hydrophobic 232 232 constant. When the H finally reaches a critical thickness (H ), the TLF ruptures and thereby bubble cr coalesces. The value of H was theoretically predicted by Vrij and his co-workers based on cr capillary wave theory [6-8],  2 3       2  3H 3R 2 1443H 4  2H 3   m f 0 (4.5) m  H  m  H  H2 2(p )    HH m   HH m  HH m  c where H 0.845H . The model assumes that the air/water interfaces of a foam film fluctuate cr m due to thermal or mechanical motion, creating corrugation. The amplitude of the wave motion grows if the disjoining pressure of the film () is negative, and the film ruptures when the two interfaces touch each other. In their original model, only the van der Waal disjoining pressure ( ) vw and the electrostatic disjoining pressure ( ) were considered, which makes it difficult to predict el H at low frother concentration, at which air bubbles become strongly hydrophobic [9]. Park et al. cr [10] predicted the values of H of foam films containing methyl-isobutyl carbinol (MIBC) using cr 44
Virginia Tech
the extended DLVO theory, which considers the contributions from the hydrophobic force ( ) hp [11]. The hydrophobic force constant (K ) determined by Wang [12] were used for this approach. 232 The model predictions are in agreement with the H values measured by Wang [12], and the results cr show that foam film stability increases with increasing frother dosage due to dampening of the hydrophobic force in the presence of a frother. Froth is a 3-phase foam, and its behavior is different from foam due to the presence of particles. As is well known, froth becomes more stable in the presence of particles [13-16]. Ata et al. [17] experimentally observed that the froth stability depends critically on particle hydrophobicity (or water contact angle θ). They measured the bubble size distribution within froth phase along the height in the presence of glass particles of different contact angles (θ = 50°, 66°, and 82°). The bubble size increased with froth height owing to increased bubble coalescence. They found also that the bubble growth rate was sensitive to θ. The bubble growth rate was lowest in the presence of the intermediate hydrophobic particles (66°) and highest in the presence of the weekly hydrophobic particles (50°). Experimentally, their results indicated that there is an optimum particle hydrophobicity (~66°) for the maximum froth stability. Binks [18] derived a model that can calculate the energy required to detach a particle from an air/water interface, G r2(1cos)2 (4.6) d 1 where r is the particle radius. For the TLF of a froth to rupture and disappear, the particle located 1 at the interface should be removed. Eq. (4.6) suggests that less hydrophobic particles can be readily washed off the TLF, whereas more hydrophobic particles will be more difficult to be washed off and thereby help stabilize the froth. If this was the case, the maximum froth stability should be achieved at θ = 90°. However, the Binks’ model alone cannot successfully explain the Ata et al.’s observations [17]. The discrepancy may be originated from the limitation that the Binks’ model is derived on the basis of only thermodynamic aspect, though the kinetics of the TLF is a more dynamic process including the hydrodynamic effect. Denkov et al. [19] proposed a stabilization mechanism of emulsions by absorbed particles. The mechanism was based on the recognition that the particles create changes in the curvature of 45
Virginia Tech
Figure 4.1 Sketch of an absorbed particle in the thin liquid film between two air bubbles. The curvature of the air/water interface around the particle creates the difference between the pressure inside of the bubble (p ) and the same in the air film (p ). 1 the air/water interface and hence the local capillary pressure (p ), which in turn affect thin film c,local rupture. Due to the similarity between a foam and an emulsion, their model may possibly be used to explain the froth stability. As shown in Figure 4.1, the p (≡ p - p ) arises from the local c,local air water curvature changes around the particle and it can be given by the Young-Laplace equation,  d p  (rsin) (4.7) c,local r dr where r is the radial position, and ϕ is the running slope angle. In their paper, the p was c,local calculated numerically by combining the following geometrical relationships, sinr p  2 c c (4.8) c,local b2 r2 c b2 r2 z  b dr (4.9) rc  2 2   r2 (b2 r2)2   p   c,local 46
Virginia Tech
  H 2 acosz (4.10) r asin() (4.11) c c where ϕ is the slope of the curvature at contact line, r the radial position of contact line, α the c c angle between the vertical line and contact line, ∆z the depth between the contact line and the interface, θ the contact angle, b the radius of the cell, and H the TLF thickness at the boundary of the cell. Figure 4.2 shows the p vs. H plots predicted from Denkov’s model. It indicates that c,local p increases slightly with decreasing H and reaches a maximum at H = 0. The model assumed c,local that the TLF ruptures at H = 0 and the maximum values of p were used as a criteria for c,local evaluating the emulsion stability. The higher the maximum values of p , the more stable c,local emulsions. Thus, the calculation results shown in Figure 4.2 suggest that the froth stability Figure 4.2 Local capillary pressure (p ) arising from a particle with contact angle (θ) c,local of 40°, 55°, 70° and 85°, respectively, as a function of the film thickness (H). The plots are obtained using Eqs. (4.8) ~ (4.11) with a = 70 μm, b = 210 μm, γ = 0.0724 N/m. 47
Virginia Tech
decreases with increasing θ. However, this trend is not compatible with previous experimental observations mentioned above [17]. The discrepancy may be due to several limitations of Denkov’s model. First, the model did not consider the whole area of a film. Only the force balance around particles was considered. Second, surface forces were neglected although recently it has shown that they are important in determining the stability of foam films [5, 20]. Finally, the model assumes that films rupture when film thickness is reduced to zero. Many investigators showed, however, that films rupture at critical rupture thicknesses, which are non-zero and vary significantly with surfactant type and dosage [5, 7, 8]. More recently, Morris et al. [21-23] conducted computational simulations to estimate the TLF stability by calculating the maximum values of p in the TLF containing multiple spherical c,local particles. Their approach was a step forward from Denkov et al.’s work in that the whole area of a film was modelled. However, their simulation result showing that the maximum values of p c,local and the TLF stability decrease with increasing θ still do not agree well with the Ata et al.’s observations [17]. The discrepancy may be attributed to the still unresolved limitations that surface forces and H were not considered and that the possible variation of the number of particles in a cr TLF with changes in θ. In the present work, we have conducted laboratory-scale flotation testes to measure froth stability in the presence of silica particles of varying surface hydrophobicity (θ = 40°, 55°, 70°, 85°). In each experiment, using a high-speed camera and image-analysis software, we have measured the bubble size ratio (d /d ), the ratio of average bubble diameter at the top of the froth 2,t 2,b (d ) and to the same at the base of the froth (d ), as a measure of froth stability. We have also 2,t 2,b developed a theoretical model that can explain the effects of θ on bubble coarsening (or froth stability). The model was derived by modelling the whole area of a TLF by considering both the capillary force and the surface forces. Furthermore, the model developed here enables us to predict d /d as a function of frother dosage, froth height, gas rate, specific power, particle size, particle 2,t 2,b hydrophobicity (collector dosage), particle concentration, etc. 48
Virginia Tech
4.2. Experiment 4.2.1 Materials and Hydrophobization of Silica Surfaces In the present study, silica spheres (Potters industries) of 35 μm diameter (d ) were used 1 for flotation test. The spheres along with a referential silica plate were firstly cleaned in Piranha solution (H O /H SO 3: 7 by volume) for 1 h at 120 °C, rinsed with ultrapure water for 10 min, 2 2 2 4 and then dried at 160 °C in an oven for 24 h. The silica surface is naturally hydrophilic; therefore, its surface was rendered hydrophobic using octadecyltrichlorosilane (OTS, 95% purity, Alfa Aesar). The spheres were immersed in a 10-4 M OTS-in-toluene (99.9% purity, Spectrum Chemical) solution along with a referential silica plate, so that both could have identical hydrophobicity. The hydrophobized surfaces were then rinsed with chloroform (99.9% purity, Fisher Chemical), acetone (99.9 % purity, Aldrich), and ultrapure water sequentially. The hydrophobicity of the particle surfaces, as measured by water contact angles (θ), was controlled by varying the Figure 4.3 Effects of immersion time on the contact angles of silica in 10-4 M OTS- in-toluene solutions. 49
Virginia Tech
a fixed gas rate of 6 L/min to obtain a superficial gas velocity of 1cm/s. The slurry was kept in suspension by agitating it at an impeller speed of 900 rpm. The froth was allowed to freely overflow, let to destabilize, and then recycled back to the cell by means of a peristaltic pump. The froth height was monitored using a graduated scale embedded on the front wall of the cell. Once a steady state condition was reached, the images of the bubbles in the froth were recorded by means of a high-speed camera. The average bubble diameters at the base of the froth (d ) and at the top (d ) were then obtained by analyzing the 2,b 2,b images offline using the BubbleSEdit, image-analysis software. The average bubble diameters were given by calculating the Sauter mean diameter (d ), 32 nd3 d  i i (4.12) 32 nd2 i i Figure 4.5 The bubble size ratio (d /d ) measured in the presence of particles with 2,t 2,b contact angle() of 40°, 55°, 70°, and 85°, respectively. The dotted line represents d /d in the absence of particles. 2,t 2,b 51
Virginia Tech
where n is the number of bubbles with diameter d. In measuring the average bubble size in a foam i i or a froth, d is most widely used in flotation literature. The number-mean diameter gives 32 excessive weights on fine bubbles, while the volume-mean diameters does son on coarse bubbles. On the other hand, flotation rate is shown is critically related to surface area of bubbles [17]. Therefore, the Sauter mean diameter may be considered most suitable for flotation studies [24]. 4.3 Experimental Results Laboratory flotation tests were conducted both in the absence and presence of monosized (35μm) silica particles. To investigate the effect of particle hydrophobicity (θ) on the bubble size ratios (d /d ), which is considered as a measure of froth stability in the present work, particles 2,t 2,b of different contact angles (θ = 40°, 55°, 70°, and 85°, respectively) were tested. In all tests, the frother dosage (10-5 M MIBC), gas flow rate (V = 1 cm/s), froth height (h = 4 cm), agitation speed g f (900 rpm), and the particle concentrations (5% w/w) were carefully kept constant. Figure 4.5 shows the experimentally measured values of d /d . As shown, the d /d 2,t 2,b 2,t 2,b ratios becomes smaller in the presence of particles. As shown, air bubbles becomes stable in the presence of particles of θ < 90°. Note also that d /d decreased with increasing θ, reaching a 2,t 2,b minimum at θ = 70°, and the increased with further increase in θ. The observed trends are consistent with Ata et al.’s results [17]. It was shown that bubbles in the froth phase immediately collapsed when very hydrophobic particles (θ > 90°) were added to a forth phase. 4.4 Model Development In section 4.4, we present a new methodology that can predict the bubble size ratios (d /d ) 2,t 2,b in a froth phase. In Subsection 4.4.1, we first calculate the driving force (p) for the thinning of a froth film containing particles with different contact angles (θ). To calculate p, we need to know N (the number of particles in a single film). Subsection 4.4.2 introduces a new model that can 1,film predict N as a function of θ. In Subsection 4.4.3, substituting the p values into the Reynolds 1,film equation, the thinning rates of the froth films are determined as a function of θ. After that, the critical rupture thickness (H ) model developed in Chapter 2 is used to predict the critical rupture cr time (t ) of the froth film as a function of θ. In Subsection 4.4.4, a model that can predict d /d cr 2,t 2,b from t is developed. cr 52
Virginia Tech
4.4.1 Driving Force for a Froth Film Thinning a. Foam Film Prior to modelling a froth film, consider a foam film stabilized in the presence of 10-5 M MIBC. As shown in Eq. (4.2), the driving force (p) for film thinning is the sum of the capillary pressure (p ) and the disjoining pressure (П). These parameters are functions of film thickness (H) c as shown in Eqs. (4.2), (4.3), and (4.4). The model parameters were calculated using the values of R = 0.44 mm and T = 298 K in accordance to the actual bubble sizes measured at the base of a foam and the temperature at which the experiments were conducted. The values of γ = 0.0724 N/m and ψ = -30 mV were also used as reported by Comley et al. [25] and Srinivas et al. [26]. Also s the values of K = 2.3×10-19 J and A = 4×10-21 J were used as reported by Wang [12]. Only the 232 232 value of κ-1= 30 nm was assumed. The calculation results are shown in Figure 4.6. As shown, p c is constant during film thinning due to the assumption of flat lamella film. On the other hand, П becomes more negative as the film becomes thinner, which can be attributed to the presence of Figure 4.6 The changes in driving pressure (p), capillary pressure (p ), and disjoining c pressure (П) as a function of a MIBC foam film thickness (H). The plots are drawn from Eqs. (4.2) ~ (4.4) with K = 2.3×10-19 J, A = 4×10-21 J, R = 0.44 232 232 mm, γ = 0.0724 N/m, T = 298 K, e = 1.6 ×10-19 C, ψ = -30 mV, and κ-1= 30 s nm. 53
Virginia Tech
attractive force in the film that varies as H-3. Note here that as the film drains, p increases mainly due to the increase in the negative disjoining pressure ( < 0). The critical rupture thickness (H ) of the foam film was determined using Eq. (4.5). In cr using Eq. (4.5), we needed to know the film size (R). Assuming that a bubble has a dodecahedron f structure, consisting of 12 films, the surface area of a single film (πR2) should be the same as f 4πR2/12. Thus, R can be determined from R as follows, f R  R/ 3 (4.13) f in which R = 0.44 mm as measured at the base of the froth (or foam) in this study. Eq. (4.13) gives the value of R to be 250 μm. In the present study, R was assumed to be independent of H in f f accordance to the previous experimental observations [20]. Figure 4.7 shows the model prediction of the values of H using Eq. (4.5). As shown, H decreases with MIBC concentration, which can cr cr be attributed to the effect of dampening of hydrophobic force in the presence of a surfactant [27]. At 10-5 M MIBC, where tests were conducted, H was found to be 171 nm. cr Figure 4.7 Plots of the critical film rupture thicknesses (H ) of a foam film predicted using cr the H model developed in Chapter 2 vs. MIBC concentration. cr 54
Virginia Tech
Figure 4.8 Sketch of a layer of particles located in the thin liquid film between two air bubbles. Near the particle a curvature change occurs (Section I), while the free film is formed away from the particle (Section II). Section III denotes a Plateau boarder area. b. Froth Film The presence of particles in froth films should cause the local curvatures at the air/water interface to change, which should in turn cause the capillary pressure (p ) to change from that of c free films. The presence of particles may also change the disjoining pressure (П) and hence the hydrodynamic pressure (p). Figure 4.8 represents a lamella film, in which three spherical particles are embedded. The film around each particle, may be subdivided into Sections I, II and III, representing the area in the vicinity of a particle, the area away from the particle, and the outside the film (Plateau border), respectively. The pressure (or force) balance in each section may be give as follows, p  p p (4.14) air I c,local 55
Virginia Tech
p  p 0 (4.15) air II 2 p  p  (4.16) air III R where p is the pressure in the air bubble, p and p are the those in Sections I and II, respectively, air I II p is the local capillary pressure that may be calculated using Eqs. (4.7)-(4.11), and R is the c,local bubble radius. Assuming that p ≈ 0, one obtains the capillary pressures in Sections I and II as III follows, 2 p  p (4.17) I R c,local 2 p  (4.18) II R The capillary pressure (p ) of the film as a whole may then be calculated as follows, c Overallcapillaryforce p AN  p A p   I I 1,film II II (4.19) c Filmarea AN  A I 1,film II where A and A is the film areas occupied by Section I and II, respectively, and N is the I II 1,film number of particles present in the single film. The values of N may vary with operating 1,film conditions. A model for predicting N is presented in Section 4.4.2. 1,film Likewise, one can calculate the overall disjoining pressure (П) using the following relation, Overall surfaceforce  AN  A   I I 1,film II II (4.20) Filmarea A II in which  and  are the disjoining pressures in Sections I and II, respectively. Section II is a I II free film; therefore, its disjoining pressure can be given as follow, e  A K     64C RT tanh2 s exp(H) 232  232 (4.21) II el vw hp el 4kT  6H3 6H3 56
Virginia Tech
On the other hand, the disjoining pressure in Section I (П) may be obtained by integrating the I local disjoining pressure along the radial direction, b  (r)(2r)dr   r c (4.22) I A I In the vicinity of a particle, (r) is small because the film thickness (H) increases sharply with decreasing r, approaching the length scale of the particle. At such large film thicknesses, disjoining pressures should be close to zero. Thus,   0. Substituting this into Eq. (4.20),    , which I II means that particles should not seriously affect П. 4.4.2 Number of Particles in a Froth Film In using Eqs. (4.19) and (4.20), we need to know the value of N (the number of particles 1,film in a froth film). This section shows a new model that can predict N theoretically. N may 1,film 1,film change continually along the froth height due to the continuing detachment of particles resulting Figure 4.9 Probability of particles surviving in a pulp phase P P (1-P ) vs. particle contact c a c angles (). The inset shows the probability of collision (P ), the probability of c attachment (P ), and the probability of detachment (P ), respectively. a d 57
Virginia Tech
from bubble coarsening. This model is aimed to model the base of a froth. Thus, N at the base 1,film of a froth is need to be predicted. The model for predicting N at the base of a froth was derived 1,film based on the assumption that the value of N may be proportional to the probability of particles 1,film reaching a froth phase. The probability of particles reaching a froth phase (P) is given by, PPP(1P) (4.23) c a d where P is the probability of collision, P the probability of attachment, and P the probability of c a d detachment. The probability functions are given as follows, respectively,  3 3  Re d  P c tanh2   2 1 16 10.249Re0.56    d1 2    (4.24)  E  P exp 1  (4.25) a   E   k Table 4.1 Input parameters for the simulation shown in Figure 4.9. Frother (MIBC) concentration (M) 10-5 Solids concentration (% w/w ) 5 Specific power (kW/m3) 22.8 Superficial gas rate (cm/s) 1.0 Froth height (cm) 4 Particle size (μm) 35 Particle Zeta potential (mV) -0.08 Bubble Zeta potential (mV) -0.03 58
Virginia Tech
 W E  P exp a 1 (4.26) d   E   k where Re is the Reynolds number, E is the energy barrier, E is the kinetic energy available during 1 k attachment process, W is the work of adhesion, and E’ is the kinetic energy of detachment. The a k plots of probability functions (P , P , and P ) are shown in Figure 4.9 and input parameters for this c a d simulation is summarized in Table 4.1. In the simulation, the bubble zeta potential was obtained from Comley et al.’s study [25], the particle zeta potential was assumed, and the other parameters were measured from experiments. Note in Figure 4.9 that the value of 1-P increases with d increasing θ, which is mainly due to the increase in adhesion force (W ) between a bubble and a a particle. It is also noteworthy that P is close to 1 even for hydrophilic particles. This is mainly due a to the relatively higher power dissipation rate (22.8 kW/m3) of laboratory flotation cells as compared to industrial flotation machines, resulting in higher kinetic energy. In the present study, N may be related to P (=P P (1 - P )) as follows, 1,film c a d N  N PP(1P) (4.27) 1,film 1,seg c a d Figure 4.10 The number of particles residing in a froth film (N ) predicted as a function 1,film of particle contact angles (). 59
Virginia Tech
where N is the number of the particles having a chance to collide with a bubble surface segment 1,seg in pulp phase ending up a single lamellar film in the froth phase. The value of N may be 1,seg proportional to the number of particles suspending in the pulp. In each experiment, we added same amount of particles, so the constant N value (N = 2667) was assumed. After that, as shown 1,seg 1,seg in Figure 4.10, we finally predicted the values of N as a function of θ. It was found that N 1,film 1,film critically depends on θ. Note here that more particles can locate in a film at higher θ, most probably due to lower P values. d Figure 4.11 The changes in driving pressure (p), capillary pressure (p ), and disjoining c pressure (П) as a function of the froth film thickness (H) containing particles with contact angle () of (a) 40°, (b) 55°, (c) 70°, and (d) 85°. The values of K = 2.3×10-19 J, A = 4×10-21 J, R = 0.44 mm, γ = 0.0724 N/m, T = 298 K, e 232 232 = 1.6 ×10-19 C, ψ = -30 mV, and κ-1= 30 nm were used. s 60
Virginia Tech
4.4.3 The Critical Rupture Time (t ) of a Froth Film cr Substituting the N values predicted in Section 4.4.2 into Eq. (4.19), one can determine particle the values of p . On the other hand, П can be determined using Eq. (4.21) using the values of K c 232 = 2.3×10-19 J, A = 4×10-21 J, R = 250 μm, T = 298 K, γ = 0.0724 N/m, ψ = -30 mV, and κ-1= 30 232 f s nm were used. Substituting the values of p and П obtained in this manner into Eq. (4.2), one can c obtain p as a function of particle hydrophobicity (θ). Figure 4.11 shows the calculated values of p , П, and p. It is noteworthy that as compared to the foam film (shown in Figure 4.6), the absorbed c particles decreased p and p during the process of film drainage, which can be attributed to the c local capillary pressure (p ) created by the particles. However, П was not changed, as c,local mentioned above. More importantly, Figure 4.11 shows that the values of p and p change with . c It was found also that with increasing  up to  = 70°, p decreased and, therefore, p decreases. On c the other hand, at  > 70° p and p increased with further increase in . c Figure 4.12 Effect of particle hydrophobicity on film thinning rate. The red line represents the critical rupture thickness (H ) of a 10-5M MIBC foam film predicted from cr the H model developed in Chapter 2. cr 61
Virginia Tech
Figure 4.13 Rupture mechanism of a thin liquid film in the presence of a particle. The film rupture may occur in the free film at H . cr Figure 4.12 shows the film thinning kinetics of a foam film and froth films containing particles with θ = 40°, 55°, 70°, and 85°, respectively. We obtained the film thinning rates by using the calculated p values shown in Figures 4.6 and 4.11 into the Reynolds equation shown in Eq. (4.1). Note in Figure 4.12 that the foam film thins fastest when p is large. In the presence of particles, the thinning rate is retarded due to relatively smaller p. In the case of the foam film, it is obvious that it ruptures at its critical rupture thickness (H ), which was predicted to 171 nm in cr Section 4.4.1. The foam film, therefore, ruptures in 2.8 s. A question that may be raised here is how to predict the H value for froth films and how to relate H of a foam film to that a froth film. cr cr Until now, there has been no model for predicting H of a froth film. cr Figure 4.13 shows a proposed rupture mechanism of a froth film. It may be reasonable to assume that the film rupture occurs in the free film section of a froth film. Since the free film section is exactly same as a foam film, we assume in the present work that H of a foam film and cr that of a froth film are identical. Accordingly, we can determine the critical rupture time (t ) of a cr Table 4.2 The critical rupture time (t ) of a froth film predicted as a function of . cr  (°) t (s) cr 40 3.1 55 3.3 70 3.7 85 3.2 62
Virginia Tech
froth film at different , as shown in Table 4.2. It has been found that t increases with  , reaching cr a maximum at θ = 70°, and then decreases with further increase in θ. 4.4.4 Prediction of Bubble Size Ratio (d /d ) from t 2,t 2,b cr Figure 4.14 shows the schematic representation of a model for predicting the bubble size ratio (d /d ) from t . As a bubble rises along the y direction in a froth phase, the number of 2,t 2,b cr bubbles decreases due to bubble coarsening. The number of bubbles can be calculated by dividing the cross-sectional area (S) of a flotation cell by bubble diameter, i.e., 4S/πd2. One can then write the following relation, 2 N S/(d /2)2 d  bubble,t  2,t  2,b  (4.28) N S/(d /2)2  d  bubble,b 2,b  2,t  Figure 4.14 A model for predicting d (bubble size at the top of the froth) from d (the 2,t 2,b same at the base). A bubble created in the pulp arrives at the base at the terminal velocity of U. Then it starts to rise along the y direction at the velocity of U t froth and thin liquid films between bubbles starts to drain. At a critical rupture time (t ), the film ruptures and the rising bubbles coalesce. cr 63
Virginia Tech
where N and N are the bubble numbers at the top and bottom of a froth, respectively. bubble,t bubble,b Eq. (4.28) indicates that the bubble size ratio can be obtained if the bubble number ratio is predicted. We assume that the number of bubbles decreases exponentially with froth height, y N (y) N exp(C ) (4.29) bubble bubble,b h f where C is the decay constant, h is the froth height, and y is the distance from the base of a froth, f which can be given by, yV t (4.30) g where V is the superficial gas velocity and t is the time. Knowing that bubbles coalesce at the g critical rupture time (t ), at which the film thickness reaches its H , Eq. (4.29) can be rewritten as cr cr follows, y N (y ) N exp(C ttcr ) (4.31) bubble ttcr bubble,b h f In the process of bubble coalescence, the total volume of bubbles should be conserved. If a single thin liquid film between two bubbles of initial diameter d ruptures and thereby one bubble- i coarsening event occurs, the diameter of the final bubble (d) should be 21/3 d. Therefore, one can f i obtain a more generalized relationship, d Nrupture f 2 3 (4.32) d i in which d is the initial bubble diameter, d is the final bubble diameter, and N is the number i f rupture of film rupture for coarsening the bubble diameter from d to d. i f Combining the relation shown in Eqs. (4.28) and (4.32), one can obtain the following equation, 64
Virginia Tech
Figure 4.15 Bubble size ratio (d /d )as a function of particle contact angle (). The lines 2,t 2,b drawn through the experimental data points represent the model predictions. N (y )  d 2  Nrupture2 bubble ttcr  2,b  2 3  (4.33) N d (y )   bubble,b  2 ttcr    where N is the number of the films that rupture at t = t among 12 films consisting a rupture cr dodecahedron-shaped bubble. Then, by combining Eqs. (4.28) ~ (4.33), finally one can deduce the following model predicting the bubble size ratio, 0.5 d   0.46h N  2,t  exp f rupture  (4.34)   d 2,b   t crV g  In using Eq. (4.34) note that the values of h, t , and V are known, while one needs to determine f cr g N . The values of N were back-calculated by fitting the model predictions using Eq. (4.34) rupture rupture to experimental results shown in Figure 4.5. In the present study, N = 5 was used at θ < 70°, rupture 65
Virginia Tech
Figure 4.16 Effect of particle contact angle () on bubble size ratio (d /d ), the number of 2,t 2,b particles (N ) in a froth film, and the local capillary pressure (p ) around 1,film c,local a particle. At  <70° the increase in froth stability with increasing  may be mainly due to the increase in N whereas at  >70° the decrease in froth 1,film, stability with increasing  may be mainly due to the decrease in p . c,local while N = 7 was assumed at θ = 85°. It indicates that more films ruptured at θ = 85°. This may rupture be probably due to the possible existence of very hydrophobic particles (θ > 90°), which act as a foam breaker. In Figure 4.15, the dots represent the experiential results shown in Figure 4.5 and the line represents the model predictions from Eq. (4.34). As shown, the model predictions are in good agreement with the experimentally measured d /d values. 2,t 2,b The model presented in the present work suggests that two independent parameters govern bubble–coarsening in a froth. One parameter is the number of particles (N ) in a film, and the 1,film 66
Virginia Tech
Figure 4.17 Effect of particle contact angle () on bubble size ratio (d /d ) and the 2,t 2,b capillary pressure (p ) at the critical film thickness (H ). c cr other is the local capillary pressure (p ) created by a single particle. The model shows that the c,local higher the N or the p , the lower the overall capillary pressure (p ), resulting in a slower 1,film c,local c drainage rate and hence a higher froth stability. Figure 4.16 shows the changes in N and p 1,film c,local with θ. Note here that at θ < 70°, the decrease in p with increasing θ can partially cause p to c,local c increase, but the effect of increasing N may overcome the p and hence decreases p . On 1,film c,local c the other hand, at θ > 70°, with increasing θ, the decrease in p , countering the N effect, c,local 1,film causes p to decrease c As a result, as shown in Figure 4.17, it was found that the overall capillary pressure p c significantly decreases with θ and begins to increase at θ = 70°. Figure 4.18 shows the effect of θ on driving force p for film drainage. This trend can give the explanation for the effect θ of on forth stability in terms of the drainage rate according to the Reynolds equation. 67
Virginia Tech
Figure 4.18 Effect of particle contact angle () on bubble size ratio (d /d ) and the 2,t 2,b driving pressure (p) for the film thinning at the critical film thickness (H ). cr 4.5 Summary and Conclusions The effect of particle hydrophobicity (or water contact angle θ) on the bubble-coarsening (or froth stability) has been studied by measuring the bubble size ratio between the top and bottom of a forth in the presence of monosized (35μm) silica particles of varying hydrophobicity (θ = 40°, 55°, 70°, and 85°, respectively). It has been found that the froth stability increased with increasing θ, reached a maximum at θ = 70°, and decreased with further increase in θ. In addition, we have developed a model for predicting the bubble-coarsening in a froth by deriving a film drainage model quantifying the effect of θ on the capillary pressure (p ), which c drives the drainage process. The parameter p is affected by the number of particles (N ) in a c 1,film film and the local capillary pressure (p ) around particles, which in turn vary with θ. The model c,local shows that as θ increases, p decreases but N increases sharply, causing p to decrease. As c,local 1,film c 68
Virginia Tech
Chapter 5. Modeling Froth Stability: Effect of Particle Size ABSTRACT A lamella film formed between two bubbles in a flotation froth drains due to the capillary pressure (p ) and the disjoining pressure (П). When the film breaks, the two bubbles become one c and the bubble size becomes coarser. In the present work, the effect of particle size (d ) on the 1 bubble-coarsening (or froth stability) was investigated. The study was conducted by measuring the bubble size ratio between the top and bottom of a forth in the presence of different sizes of particles (d = 11, 35, 71, and 119 μm). It was found that the froth stability decreases considerably as particle 1 size becomes coarser from 11 to 71 μm. However, as particle size increases further to 119 μm, the froth stability changes little. In the present work, a model for predicting the bubble-coarsening in a froth has also been developed by deriving a film drainage model quantifying the effect of d on p . The model indicates 1 c that as d increases, p increases and thereby the film thins faster and the froth becomes unstable. 1 c 72
Virginia Tech
5.1 Introduction As bubbles migrate upward in a froth phase, they coalesce with each other and become larger. As bubbles become larger, bubble surface area becomes smaller, restricting the number of hydrophobic particles that can be carried upward and flow into a launder. On the other hand, in the process of the bubble coarsening, less hydrophobic particles tend to be removed more easily, it can contribute to increase the grade of a froth product [1]. Thus, the throughput of a flotation cell depends on bubble coarsening. Therefore, it is important to understand the basic mechanisms of bubble coarsening and froth (or foam) stability. A foam is thermodynamically unstable due to the large surface area, which means high surface energy. Therefore, in pure water air bubbles collapse immediately to reduce the free energy. It is generally known that surfactants enhance the foam stability. When a surfactant adsorbs at air/water interface, it can reduce the thermodynamic instability by lowering the surface energy. In the case of an ionic surfactant, it can also enhance the foam stability by increasing the electrostatic repulsion force acting between two air/water interfaces and thereby regarding the film thinning. It is known that a solid particle can also act as a surfactant. A froth is a three-phase foam, where particles are present and a froth is generally more stable than a foam due to the presence of particles [2-5]. It has been reported that the froth stability depends on particle properties, including surface hydrophobicity [2, 3, 5-7], shape [8], concentration [2], and size [2-4]. With regard to particle size (d ) effect, several experiential studies have been reported. 1 Tang et al. [4] have found that smaller particles can increase the froth stability. They used silica particles of d < 770 nm, which is relatively fine as compared to flotation practice. Johansson and 1 Pugh [5] conducted static and dynamic froth stability tests in the presence of fine (26 ~ 44 μm) and coarse (74 ~ 106 μm) quartz particles with alcohol type frothers. They also found that generally fine particles can cause higher stabilizing effect, but how the change in d affect the froth stability 1 was explained. Ip et al. [2] have measured the froth life time, as indicator of the froth stability, in a flotation column in the presence of silica particles (40 μm < d < 150 μm) while the volume 1 faction of particles in the slurry was kept constant. They also observed that the froth stability increases as d decreases. It was suggested that at a constant volume faction, as d decreases, the 1 1 number of particles in the slurry and the collection efficiency can increases, which will result in 73
Virginia Tech
the high surface coverage of a bubble by particles. They assumed that the higher surface coverage by finer particles can enhance the froth stability. More recently, Aktas and his co-workers [3] also have shown that fine particles benefit the froth stability through dynamic stability tests. However, up to now, there is no theoretical model that can explain the influence of particle sizeon the froth stability. Fortunately, we have recently developed a froth model for predicting bubble coarsening in a froth phase as a function of particle size. The model is based on the recognition that particles can reduce the capillary pressure (p ), which contribute to thin a liquid film formed between bubbles c in a froth. As a result, particles can decrease the film drainage rate and thereby enhance the froth stability. The model is also based on the premise that parameter p is affected by the number of c particles and the local capillary pressure (p ) around particles, which vary with particle size in c,local the film. In the present work, the model was verified by comparison with our experiments testing the influence of particle size on froth stability. We have conducted a series of laboratory-scale flotation experiments with 10-5 M methyl-isobutyl carbinol (MIBC) solutions in the presence of silica particle of different sizes (d = 11, 35, 71, and 119 μm). In each experiment, the bubble size 1 ratio (d /d ), the ratio of average bubble diameter at the top of the froth (d ) and to the same at 2,t 2,b 2,t the base of the froth (d ), as a measure of the froth stability, was calculated. 2,b 5.2 Experiment 5.2.1 Materials and Hydrophobization of Silica Surfaces In the present study, to investigate the influence of particle size on froth stability, four different diameters (d = 11, 35, 71, and 119 μm) of silica spheres (Potters industries) were treated, 1 respectively. First, silica particles of identical size and a reference silica plate were cleaned by immersing them in boiling Piranha solution (H O /H SO 3: 7 by volume) for 1 h. Then they were 2 2 2 4 rinsed in ultrapure water for 10 min and subsequently allowed to dry thoroughly in an oven at 160 °C overnight. After that, the cleaned surfaces of both the particles and the plate were simultaneously hydrophobized by soaking them together in 10-4 M octadecyltrichlorosilane (OTS, 95% purity, Alfa Aesar)-in-toluene (99.9% purity, Spectrum Chemical) solution. After 74
Virginia Tech
hydrophobization, they were washed with chloroform (99.9% purity, Fisher Chemical), acetone (99.9 % purity, Aldrich), and ultrapure water, sequentially. Due to the same conditioning time of the particles and the plate, it may reasonable to assume that they exhibit the identical water contact angles (θ). Hence, the θ value of the particles were simply estimated by measuring that of the plates using a goniometer (Rame-hart instrument co.). By varying the immersion time in 10-4 M OTS-in- toluene, the θ value was controlled. Until θ reaches 40°, by repeating the procedure stated above, four different diameters (d = 11, 35, 71, and 119 μm) of silica spheres with θ = 40° were prepared 1 prior to experiments. 5.2.2 Experimental Procedure In the present study, we have carried out flotation tests by means of a Denver laboratory flotation machine. A 1.5 L of transparent glass cell was specially made to observe bubble- coarsening phenomena in a froth phase clearly. The cell was built with flat plates to reduce optical distortion. Prior to each experiment the cell was cleaned thoroughly with distilled water. The cell was, then, filled with 10-5 M methyl-isobutyl carbinol (MIBC, 98% purity, Aldrich) aqueous solution. Before generating air bubbles, the 60 g silica particles (solids concentration of 5% w/w) of an identical size were added to the MIBC solution and stirred at the rotation speed of 900 rpm for 5 min for wetting. After that, a froth phase was created by injecting air to the cell at the superficial gas velocity of 1 cm/s. The formed froth was allowed to overflow and then recirculated to the cell using a peristaltic pump. The froth height was adjusted to be 4 cm by controlling the pulp-froth level. During the experiment, the bubble images in a froth were recorded using a high- speed camera (Fastec imaging). We repeated the experiments with changing particle sizes. Additionally, an experiment was conducted in the absence of particles for comparison with foam stability. After the flotation tests, the Sauter mean bubble diameter at the base of the froth (d ) and 2,b the same at the top (d ) were calculated by analyzing the acquired images by means of 2,b BubbleSEdit, image-analysis software. Finally, the bubble size ratio (d /d ) of each particle size 2,t 2,b was obtained. 75
Virginia Tech
Figure 5.1 The bubble size ratio (d /d ) measured in the presence of different sizes of 2,t 2,b particles at θ = 40°. The dotted line represents d /d in the absence of 2,t 2,b particles. 5.3 Experimental Result Fig ure 5.1 shows the values of experimentally measured bubble size ratio (d /d ) in the 2,t 2,b absence of particles and in the presence of particles of different sizes at θ = 40°. It should be reasonable to assume that the higher d /d indicates the lower froth stability. As shown, it was 2,t 2,b observed that froth stability decreases considerably as particle size becomes coarser from 11 to 71 μm. As particle size increases further to 119 μm, the froth stability changes little. It is also noticeable that in the present experimental range the d /d values of the froth were lower as 2,t 2,b compared to the foam phase. It indicates that particles at θ = 40° can stabilize bubbles or thin liquid films between the bubbles over wide range of particle sizes. 76
Virginia Tech
Table 5.1 Input parameters for the simulation shown in Figure 5.3. Frother (MIBC) concentration (M) 10-5 Solids concentration (% w/w ) 5 Specific power (kW/m3) 22.8 Superficial gas rate (cm/s) 1.0 Froth height (cm) 4 Particle contact angle (°) 40 Particle Zeta potential (mV) -0.08 Bubble Zeta potential (mV) -0.03 where ρ is the density of the particle. In the present experiment, m was set to 60 g and ρ of silica 1 1 1 was 2.65 g/cm3. As shown in Figure 5.2, with increasing d , N sharply decreases as d -3. 1 1,pulp 1 The plots of probability functions (P , P , and P ) were then drawn as a function of d in c a d 1 Figure 5.3 and input parameters used for the calculations were listed in Table 5.1. While only the values of particle zeta potential was assumed, the bubble zeta potential was obtained from Comley et al.’s study [9] and the others were measured in experiments. Figure 5.4 shows the values of P P (1-P ) obtained from Figure 5.3. It was found that fine c a d particles are less unlikely to survive in the pulp due to lower P and P , while coarse particles are c a less unlikely to survive in the pulp due to higher P . d Finally, as shown in Figure 5.5, the values of N were obtained by substituting Eqs. (5.2) 1,film and (5.3) into (5.1). In using (5.2), the value of c was arbitrary chosen to 384244. It was found that N decreases as d increases. The N values shown in Figure 5.5 were used to calculate the 1,film 1 1,film driving force for froth films in the following section. 79
Virginia Tech
where П and П are the disjoining pressure arising in section I and II, respectively. The values of I II П and П were expressed as, respectively, I II b  (r)(2r)dr   rc  0 (5.10) I A I e  A K     64C RT tanh2 s exp(H) 232  232 (5.11) II el vw hp el 4kT  6H3 6H3 where П is the disjoining pressure due to electrostatic interaction, П is the disjoining pressure el vw due to van der Waals dispersion force, П is the disjoining pressure due to hydrophobic force, C hp el is the electrolyte concentration, R the gas constant, T the absolute temperature, e the electronic 2 charge (e = 1.6 ×10-19 C), ψ the surface potential at the air/water interfaces, k the Boltzmann’s s Figure 5.7 Effect of particle size on film thinning rate. The red line represents the critical rupture thickness (H ) of a 10-5M MIBC foam film predicted from the H model cr cr developed in Chapter 2. 83
Virginia Tech
constant, κ the reciprocal Debye length, A the Hamaker constant, and K the hydrophobic 232 232 constant. Figure 5.6 shows the effect of d on p, p , and П. In the calculations, T = 298 K was used 1 c from the measurement in the experiment. The values of γ = 0.0724 N/m and ψ = -30 mV were s used as reported by Comley et al. [9] and Srinivas et al. [12], respectively. Also the values of K 232 = 2.3×10-19 J and A = 4×10-21 J were used as reported by Wang [13]. Only the value of κ-1= 30 232 nm was assumed. As shown, p increases as d becomes larger from 11 to 71μm. As d increases c 1 1 further to 119 μm, p changes little. Since the value of П is independent of d , the change in p c 1 c dominate the variation of p. Note in Figure 5.5 that the variation trend of p with d is similar with 1 that of d /d . 2,t 2,b 5.4.3 The Critical Rupture Time (t ) of a Froth Film cr As shown in Figure 5.7, substituting the p values obtained in Figure 5.6 into Eq. (5.4), we calculated the film thinning rates of films containing particles of d = 11, 35, 71, and 119 μm, 1 respectively. In the calculation, assuming that a bubble has a dodecahedron structure, R was f calculated to 250 μm from the following geometrical considerations, R  R/ 3 (5.12) f in which R was 0.44 mm as measured at the base of the froth (or foam) in this study. It is noticeable that the thinning velocity of the film containing particles of d =11 μm is the slowest, which is 1 attributed to the smallest p values. As d gets coarser to 35 μm, the thinning rate becomes 1 significantly slower, mostly due to the significant decrease in p . As d becomes coarser to 71 μm, c 1 the thinning rate becomes little slower, mostly due to small decrease in p . It was also found that c the thinning rates of d = 71 μm and 119 μm are almost same due to similar p values. 1 As presented in Chapter 2, in the present work, the critical rupture thickness (H ) of the cr foam film were predicted by incorporating the hydrophobic force into the capillary wave theory, 84
Virginia Tech
  2  3 2  3H 3R 2 1443H 4  2H 3   m f 0       m H HHm  m H HHm  H2 HHm  2(p c ) (5.13) H 0.845H cr m The model prediction from Eq. (5.13) showed that H of the 10-5 M MIBC foam film of R = 250 cr f μm is 171 nm. In the present study, it is assumed that the film in the presence of particles has the identical H values as compared to the foam film, due to the hypothesis that the film rupture of cr the froth film may happen in the free film section, which is similar with a foam film. Consequently, the values of the critical rupture time (t ) of the froth film was obtained as a function of d , as cr 1 shown in Figure 5.8. Figure 5.8 The critical rupture time (t ) of a froth film predicted as a function of particle cr size. 85
Virginia Tech
5.4.4 Prediction of Bubble Size Ratio (d /d ) from t 2,t 2,b cr In Chapter 4, a theoretical model for predicting bubble size ratio (d /d ) from t was 2,t 2,b cr derived, as follows, 0.5 d   0.46hN  2,t exp f rupture (5.14)   d t V    2,b cr g where h is the froth height, V is the superficial gas rate, and N is the number of the films f g rupture that rupture at t among 12 films forming a dodecahedron-shaped bubble. In using Eq. (5.14), the cr values of h, t , and V are known, N needs to be determined. N was assumed to be 5. at f cr g rupture rupture d = 11 and 35 μm. At d =71 and 119 μm, however, based on the recognition that possibly more 1 1 films could rupture at t due to the increase of the area occupied by free film with small particle cr Figure 5.9 Bubble size ratio (d /d )as a function of particle size (d ). The lines drawn 2,t 2,b 1 through the experimental data points represent the model predictions. 86
Virginia Tech
Figure 5.10 Effect of particle size (d ) on bubble size ratio (d /d ) and the number of 1 2,t 2,b particles in a froth film (N ). The increase in d /d with increasing d may 1,film 2,t 2,b 1 be partially attributed to the decrease in N . 1,film number, slightly higher value (N = 5.5) was used. As shown in Figure 5.9, the model rupture predictions agree with the experimental results. In Chapter 4, it was found that two independent parameters determine the value of d /d . 2,t 2,b One parameter is the particle number in a film (N ) and the other is the local capillary pressure 1,film (p ) around a particle. The increase in both parameters causes the overall capillary pressure p c,local c to decrease, resulting in decrease in driving force p and drainage rate. It was found in the present work that the two parameters vary with particle size as shown in Figure 5.10 and 5.11, respectively. Note in Figure 5.10 that N decreases as a particle become coarser, which may partially attribute 1,film to the increase in overall capillary pressure p . Note in Figure 5.11 that p decreases as a particle c c,local become coarser, which may also partially attribute to the increase in p . c 87
Virginia Tech
Figure 5.11 Effect of particle size (d ) on bubble size ratio (d /d ) and the local capillary 1 2,t 2,b pressure (p ) when the film ruptures. The increase in d /d with increasing c,local 2,t 2,b d may be partially attributed to the decrease in p . 1 c,local As shown in Figure 5. 12, since both N and p cause p to increase with increasing 1,film c,local c particle size, it was found that p increases with particle size c Figure 5.13 shows the effect of particle size on driving force p for filming drainage. This trend can give the explanation for the effect of on particle size forth stability in terms of the drainage rate according to the Reynolds equation. 5.5 Summary and Conclusions In the present study, the effect of particle size (d ) on the bubble-coarsening (or froth 1 stability) has been studied by measuring the bubble size ratio between the top and bottom of a forth in the presence of different sizes of particle (d = 11, 35, 71, and 119 μm). We found that the froth 1 88
Virginia Tech
Chapter 6. Conclusions and Recommendations for Future Research 6.1 Conclusions The primary findings and contributions presented in the dissertation are summarized as follows. 1. A thin liquid film (TLF) confined between two bubbles in a froth phase (or in a foam) drains by the capillary pressure (p ) created from the changes in curvature and the disjoining c pressure (П) created by surface forces in the films. If П is negative (attractive), the film drainage rate and the wave motions at the air/water interfaces accelerate. When the TLFs thin to a critical film thickness (H ), the TLF ruptures and the two bubbles become one. The capillary wave model cr describes the film thinning process and the wave motions using the classical DLVO theory, which considers the repulsive double-layer force and the attractive van der Waals forces only. It has been found in the present work that the H values predicted from the capillary wave model are cr substantially smaller as compared to the experientially measured values in the case of the foam films stabilized by with weak surfactants, e.g., MIBC. 2. Based on the recognition that attractive hydrophobic force is also present in foam films in addition to the double-layer force and the van der Waals force, by incorporating the hydrophobic force in the capillary wave model, the author has developed a model that can predict H more cr accurately. The model shows that the hydrophobic force contributes to accelerate both film thinning rate and surface corrugation growth rate and thereby may cause the film to rupture at higher film thickness. 3. Based on the new H model, a model for predicting bubble-coarsening in a dynamic cr foam has been developed in the present work. The model was developed by deriving a 92
Virginia Tech
mathematical relation between the Plateau border area, which controls film drainage rate, and the lamella film thickness, which controls bubble-coalescence rate. The model is able to predict the bubble size ratio between the top and bottom of a foam as a function of surface tension (frother dosage), aeration rate, and foam height. The model was validated using a specially designed foam column equipped with a high-speed camera. It has been found that bubble-coarsening increases with decreasing frother dosage and aeration rate and increasing foam height. 4. In the present study, a model for predicting bubble-coarsening in a froth (3-phase foam) has been developed for the first time. The model was developed by deriving a film drainage model quantifying the effect of particles on p . The parameter p is affected by the number of particles c c and the local capillary pressure (p ) around particles, which in turn vary with the c,local hydrophobicity and size of the particles in the film. Assuming that films rupture at free films, the p corrected for the particles in lamella films has been used to determine the critical rupture time c (t ), at which the film thickness reaches H , using the Reynolds equation. Assuming that the cr cr number of bubbles decrease exponentially with froth height, and knowing that bubbles coalesce when film drains to a thickness H , a bubble coarsening model has been developed. This first cr principle model is capable of predicting the bubble size ratio between the top and bottom of a froth phase as a function of frother dosage, collector dosage (contact angle), particle size, aeration rate, and froth height. 5. The bubble-coarsening froth model developed in the present study has been verified using a Denver laboratory flotation cell using spherical particles of varying hydrophobicity and size. It has been found in the present study that bubble-coarsening decreases with particle contact angle (θ) up to θ = 70o due to a decrease in p , resulting in retarded film drainage rate. At θ > 70o, c on the other hand, the bubble-coarsening increases due to increased p and, hence, increased c drainage rate. In addition, it has been shown that bubble-coarsening decreases with particle size, which may be attributed to decreased p and drainage rate. c 6.2 Recommendations for Future Research Finally, the author of this dissertation recommends the following works for future research. 1. In a flotation model, a froth recovery is given by a function of the bubble-coarsening factor, which is the bubble size ratio between the top and bottom of a froth phase. By incorporating 93
Virginia Tech
VALIDATION AND APPLICATION OF A FIRST PRINCIPLE FLOTATION MODEL KAIWU HUANG ABSTRACT A first principle flotation model has been derived from the basic mechanisms involved in the bubble-particle and bubble-bubble interactions occurring in flotation. It is a kinetic model based on the premise that the energy barrier (E ) for bubble-particle interaction can be reduced 1 by increasing the kinetic energy (E ) for bubble-particle interaction and by increasing the k hydrophobic force in wetting films. The former is controlled by energy dissipation rate (), while the latter is controlled by collector additions. The model consists of a series of analytical equations to describe bubble generation, bubble-particle collision, attachment and detachment, froth recovery, and bubble coalescence in froth phase. Unlike other flotation models that do not consider role of hydrophobic force in flotation, the first principle model developed at Virginia Tech can predict flotation recoveries and grades from the chemistry parameters such as ζ- potentials, surface tension (), and contact angles () that may represent the most critical parameters to control to achieve high degrees of separation efficiencies. The objectives of the present work are to i) validate the flotation model using the experimental data published in the literature, ii) incorporate a froth model that can predict bubble coarsening due to coalescence in the absence of particles, iii) develop a computer simulator for a froth model that can predict bubble coarsening in the presence of particles, and iv) study the effects of incorporating a regrinding mill and using a stronger collector in a large copper flotation circuit. The model validation has been made using the size-by-class flotation rate constants (k ) ij obtained from laboratory and pilot-scale flotation tests. Model predictions are in good agreement with the experimental data. It has been found that the flotation rate constants obtained for composite particles can be normalized by those for fully liberated particles (k ), which opens max the door for minimizing the number of flotation products that need to be analyzed using a costly and time-consuming liberation analyzer. A bubble coarsening froth model has been incorporated into the flotation model to predict flotation more accurately. The model has a limitation, however, in that it cannot predict bubble- coarsening in the presence of particles. Therefore, a new computer simulator has been developed to predict the effects of particle size and particle hydrophobicity on bubble coarsening in froth phase. In addition, the first principle flotation model has been used to simulate flotation circuits that are similar to the Escondida copper flotation plant to study the effects of incorporating a re- grinding mill and using a more powerful collector to improve copper recovery. The flotation model developed from first principles is useful for predicting and diagnosing the performance of flotation plants under different circuit arrangements and chemical conditions.
Virginia Tech
ACKNOWLEDGEMENT My most sincere thanks go to my advisor, Dr. Roe-Hoan Yoon, who introduced me to the wonders and frustrations of scientific research. I thank him for his guidance, encouragement and support throughout this project. I would also like to thank Dr. Gerald Luttrell for his great help in dealing with problems in Excel-VBA. Finally, I would like to thank Dr. Greg Adel for serving on my committee. I would like to express my deepest appreciation to Professor Jean-Paul Franzidis, SA Research Chair, Department of Chemical Engineering, University of Cape Town, South Africa, for his permission to use the size-by-class flotation rate constants and liberation data presented in the Ph.D. thesis authored by his former student Dr. Simon Welsby at the University of Queensland, Australia. I would also like to express my appreciation to Dr. Jaakko Leppinen, Technology Director – Mineral Processing, Outotec, for providing experimentally determined flotation rate constants along with relevant liberation data. I am also grateful to FLSmidth for funding continuously for this project. I also want to express my sincere gratitude to Dr. Lei Pan, Dr. Seungwoo Park, Dr. Juan Ma, Dr. Aaron Noble, Gaurav Soni, Zhenbo Xia, and Biao Li for helping me from time to time by giving me valuable advice. Finally, I would like to express my eternal gratitude to my parents for their everlasting support and love. iii
Virginia Tech
Nomenclature and Symbols1 DLVO- Derjaguin and Landau, Verwey and Overbeek MIBC- Methyl Isobutyl Carbinol a & b- Fitting parameters in the normalized rate constant (k/k ) model max A- Cross-sectional area of the plateau border A - Plateau border area at the bottom of a foam b A - Critical rupture PB area cr A- Plateau border area at the top of a foam t A - Hamaker constant for van der Waals interaction between two air/water interfaces 232 b- Fitting parameters in the contact angle calculation i C- Fitting parameter in the bubble coarsening foam model C - Electrolyte concentration el d - Particle diameter 1 d - Bubble diameter 2 d - Collision diameter 12 d - Diameter of bubbles entering the froth phase 2,b d - Diameter of bubble at the top of froth phase 2,t e- Electronic charge E - Energy barrier 1 E - Kinetic energy of attachment k E’ - Kinetic energy of detachment k g- Gravitational acceleration constant ΔG- Change of Gibbs free energy h- Froth height f H- Thin liquid film thickness H - Critical rupture thickness cr H - Medium thickness of a thin liquid film m H - The closest separation distance between two bubble surfaces 0 k- Overall flotation rate constant k - Overall rate constant for the fully-liberated particles max k - Flotation rate constant in the pulp phase p k’- Boltzmann’s constant K - Hydrophobic force constant between bubble and particle 132 K - Hydrophobic force constant between two particles 131 K - Hydrophobic force constant between two bubbles 232 L- Rate constant ratio m - Mass of the particle 1 m - Mass of the bubble 2 n- Number of flotation cells N - Number of plateau borders at the base of a foam 0 1 Symbols in red are the fitting parameters in the model. x
Virginia Tech
N - Number of the particles 1 N - Number of the bubbles 2 N - Number of the plateau borders pb N - Number of films that rupture in the bubble coalescence process rupture p - Capillary pressure c P- Flotation probability P - Probability of attachment a P - Probability of collision c P - Probability of detachment d P- Probability of bubble-particle aggregate surviving the froth phase f P- Probability of bubble-particle aggregates transferring from the pulp to the froth t r - Radius of the particle 1 r - Radius of the bubble 2 R- Overall flotation recovery R - Froth recovery due to entrainment e R - Pulp phase recovery p R- Froth phase recovery f R - Film radius film R - Maximum fraction of the particles entering the froth phase max R - Maximum theoretical water recovery w R’- Gas constant Re- Reynolds number S - Bubble surface area flux b t- Retention time t - Critical rupture time of the thin liquid film cr t - Drainage time d T- Absolute temperature 𝑢̅ - Particle RMS velocity 1 𝑢̅ - Bubble RMS velocity 2 U- Liquid drainage velocity U - Radial velocity of the particle approaching a bubble 1 U - Velocity of a particle approaching a bubble at the critical rupture distance Hc V - van der Waals interaction energy D V - Electrostatic interaction energy E V - Hydrophobic interaction energy H V - Superficial gas rate g W - Work of adhesion a x- Fractional surface liberation Z - Collision frequency between particles and bubbles 12 α- Fitting parameter in froth recovery calculation β- Drag coefficient ε- Energy dissipation rate ε - Energy dissipation rate at bubble generation zone b ε ’- Liquid fraction at the base of a foam (or froth) b xi
Virginia Tech
Chapter 1: INTRODUCTION 1.1 Background 1.1.1 Flotation History and Application Flotation is undoubtedly the most important and versatile industrial process for the separation and concentration of minerals [1]. It is an amazing separation process that enables minerals denser than water to float to the top as bubble-particle aggregates for collection. This is achieved by exploiting the differences in surface properties between valuable minerals and gangue. In 1860, the first hint that minerals can be separated from each other according to the differences in their surface properties appeared in a patent awarded to William Haynes [2], who claimed that sulfides could be floated by oil and non-sulfide minerals could be removed by washing in a powdered ore. Bessel brothers built the first commercial flotation plant in Dresden, Germany, which was used to purify the graphite ore. The CO bubbles were applied in their plant, which were 2 generated by the reaction of lime with acid. The first flotation plant to process sulfide ores was based on Carrie Everson’s patent. The year 1885 was important in the flotation history due to the patents by the Bessel brothers and Carrie Everson. True industrialization of the flotation process, from being a research topic in the lab to a more commercially valuable tool, occurred in the early twentieth century [3]. In 1901, the immediate problem occurred at Broken Hill, Australia, which was finding a way to recover sphalerite fines from the waste dumps. Several flotation processes and machines were studied there by engineers in different programs. The results of these programs were that froth flotation was developed as an industrial process for concentrating sulfides and was used to recover zinc from millions of tons of slime tailings. Froth flotation was used in the United States for the first time in 1911. The first flotation plant in the US was installed by James M. Hyde in Basin, Montana [4], who understood and verified that the use of rougher-cleaner closed circuits could remove entrained gangue particles from concentrates. The success of this plant was a milestone which represented flotation was poised to take off [5]. During 1925-1960, the introduction of chemical reagents and the trend of selective flotation brought about more widespread application of flotation process as an economic tool. To meet the increased demand for minerals by flotation, the capacity of flotation plant increased a lot. In the following decades, with the improvement of flotation cells and development of on- stream analysis (OSA) systems, the mineral production from flotation increased rapidly and accurate control of flotation circuits was achieved. 1
Virginia Tech
Over the past decades, flotation has been used not only in the mineral processing industries, but also in the food industries, e.g., removing solids in butter and cheese. It is also commonly used for removing the contaminant from water so that purification can be achieved. Other areas in which flotation can be applied are de-inking of recycling paper, paint manufacturing, and paper industry [5]. The flotation process as it exists today remains essentially the same as it was in Broken Hill [6]. A feed of slurry is pre-treated with suitable reagents in a tank where it is agitated to keep the solids in suspension before being pumped into a series of flotation cells. In the flotation cell, the slurry is agitated by an impeller, where air is also injected to generate fine bubbles. As bubbles rise in the slurry, they collect hydrophobic particles selectively and enter the froth phase on the top. The forth laden with hydrophobic particles overflows the cell lip and recovered as a concentrate. 1.1.2 Flotation Process Flotation is a process for separating finely divided solids from each other using air bubbles under hydrodynamic environment. The process is based on separating hydrophilic particles from hydrophobic ones in the slurry by attaching the latter selectively onto the air bubble surfaces [3]. Specific chemicals, which are known as collectors, are added to the slurry before flotation to increase the differences in hydrophobicity of the minerals to be separated. In general, the recovery and selectivity of flotation increases with increasing hydrophobicity difference. Thermodynamically, for bubble-particle attachment to occur, the change of Gibbs free energy (ΔG) must be less than zero. The changes of free energy in bubble-particle attachment can be described as the changes in the interfacial tensions at the solid-liquid, solid-air and air- liquid interfaces [7]. By applying Young’s equation, one can obtain the following relationship for the change of Gibbs free energy, 2    c o s  1    G lv (1) where γ is the interfacial tension and θ is the contact angle at the three phase contact point. Eq. lv (1) shows that G < 0 when  > 0, and that the higher the contact angle, the more negative ΔG becomes. Bubble-particle aggregates rise through the pulp since the overall density of the aggregates are lower than the density of the slurry. At the pulp/froth interface, some of the air bubbles loaded with various particles enter the froth phase, while others may drop off depending on the bubble size, particle size, and bubble loading. Froth is a complex three-phase system, which contains air bubbles, particles, and liquid films. At the bottom of a froth phase, bubbles are separated by thick water films. The liquid films become thinner as the bubbles rise in the froth phase, creating thin liquid films (TLFs) (or lamellae films). Three lamellae films meet at a plateau border (PB), through which water drains. As the thickness of the lamella film between two bubbles reach a critical thickness (H ), the TLF ruptures and two bubbles become one, cr
Virginia Tech
which is referred to as bubble coarsening. As bubbles become larger, the surface area on which hydrophobic particles are attached (or ‘parked’) become smaller, forcing less hydrophobic particles to detach and return to the pulp phase [8]. Thus, the bubble coarsening provides a mechanism by which product grade improves. Two mechanisms, i.e., recovery due to attachment and recovery due to entrainment, contribute to the recovery in froth phase. The former represents true recovery based on hydrophobic interactions, while the latter represents unwanted recoveries due to hydraulic entrainment associated with the recovery of water or water split [9]. Fine particles with low inertia are prone to the hydraulic entrainment. 1.1.3 Role of Modeling Flotation Flotation is a complex physiochemical process involving solid, liquid, and gas phases; therefore, the number of parameters affecting the process is large. These parameters can be subdivided broadly into two groups, i.e., hydrodynamic and surface chemistry parameters. The former includes particle size, bubble size, energy dissipation rate, etc., while the latter includes contact angle (θ), -potential, Hamaker constants, and surface tension (γ). Many investigators developed flotation models in the past, most of which are based on the hydrodynamic parameters. On the other hand, the separation efficiencies of flotation are determined by control of surface chemistry parameters rather than hydrodynamic parameters, particularly the hydrophobicity of the particles to be separated, as has already been noted in the foregoing section. For this reason, Virginia Tech has been developing a flotation model using both the hydrodynamic and surface force parameters. The first principle model can, therefore, predict both the recovery and grade for the first time. Having a first principle model has many advantages, the most important aspects including predictive and diagnostic capabilities. There are some parameters that are difficult to be tested in experiment without affecting other parameters. For example, changing the pH to study the effect of -potentials of particles also affect the -potentials of air bubbles as well as the collector adsorption and hence the particle contact angles. A first principle model can easily study the effects of isolated process variables one at a time and, thereby, optimize a flotation circuit. In addition, the model-based simulator can be used design flotation plants with minimal input from experiment. 1.2 Literature Review 1.2.1 Flotation Modeling Flotation is a 3-phase separation process. Therefore, modelling flotation is difficult simply because of the large number of parameters involved. Furthermore, a bubble-particle interaction involves several different sub processes, which need to be modeled separately. A large number of factors and interactions between them need to be considered in the flotation model, since all of these can affect the flotation results in different ways. 3
Virginia Tech
Currently, there are several academic or commercial flotation models and simulators to predict the performance of flotation circuits or a flotation unit, such as the P9 Flotation Model, limn, USIM PAC simulator and SUPASIM flotation simulation program. The P9 Flotation Model has been developed at the Julius Kruttschnitt Mineral Research Center (JKMRC) over the past twenty years [6]. Both entrainment and true flotation are considered when it comes to simulating the recovery of particles. The recovery is given as [10]: 4 R i, j  1  P  i, P Sj i, b Sj R b fi, R j fi, 1 1j R  w R  w   E  N E T N Ri T i w R w  (2)  where subscript i represents a particle size class, and subscript j represents a surface liberation class. P is ore floatability, R is overall recovery, τ is residence time, R is water recovery, S is w b the bubble surface area flux in the pulp zone and ENT is degree of entrainment. Drawback of this model is that the parameters used in the model must be acquired from flotation tests data and surveying. Therefore, the collection of representative samples and good experimental data can determine that whether a simulation of a flotation circuit/unit is a successful one or not. Limn software is an Excel-based application that allows the user to draw and model a circuit. Limn software incorporates partition models for gravity separation and size separation, which is powerful in stream simulation. In the Limn, however, the flotation recovery is also simulated by the partition model, in which the Ep and Rho50 values are entered manually to fit the yield and grade data from the flotation tests. The lack of effective flotation model is the main disadvantage of the Limn software. SUPASIM flotation simulation model was developed in the mid-1980s by Eurus Mineral Consultants to predict plant performance from standard laboratory flotation tests [11]. The model is based on Kelsall’s unmodified equation in which two rate constants appear, R  1 0 0    1  e x p   k f t    1  e x p   k s t   (3) where Θ is slow floating fraction, t is flotation time, R is percentage recovery at time t, k is fast f floating rate constant and k is slow floating rate constant. Θ, k, and k are estimated from the s f s laboratory batch flotation tests. Based on these parameters, the continuous flotation process can be modelled by applying the scale-up algorithms. USIM PAC simulator is an easy to use steady-state simulation software developed by BRGM since 1986 [12]. It contains several flotation models which can be classified as “performance” models and “predictive” models [13]. Performance models are made for material balance calculation and definition, while predictive models are based on kinetic approach.
Virginia Tech
A Model with two kinetic rate constants considers that the feed is composed of three “sub-populations”, non-floating, fast floating and slow floating. Assuming that each cell is represented as a perfectly mixed reactor, flotation can be described as,   1    1  F  F Rinf 1  1 1  (4) fj j j  j  1ks j  j   1kf j  where F is flow rate of mineral j in the froth, F is flow rate of mineral j in the feed, Rinf is fj j j maximum possible recovery of j in the froth, φ is proportion of mineral j capable of floating and j which shows slow floating behavior, and τ is mean residence time. Another predictive model incorporates a distribution of kinetic constants according to particle size. The kinetic rate constant is calculated for each mineral j and each size class i as below, 5 k i, j  x 0i .5  1   x x l i j  1 .5  e x p    x 2 e x j i  2   (5) where x is average size in size fraction i, α is adjustment parameter for mineral j, xl is the i j j largest floating particle size for mineral j, and xe is the easiest floating particle size for mineral j. j In perfectly mixed condition, flotation can be described as,       F f i, j F i, j R i n f 1 1 1 k i, j (6) where Ffi,is flow rate of mineral j and size class i in the froth and Fi,j is flow rate of mineral j j and size class i in the feed. USIM PAC also includes an entrainment model, which is based on the reference [14]. Recovery due to entrainment is shown below, R i, j  P i, j R w (7) where P is recovery of mineral j in size class i in one cell and R is water recovery. i,j w The flotation models or simulators mentioned above all require basic input data from flotation tests, which causes the limitation of applying these models to predict flotation performance. However, the flotation model developed from first principles helps better understand and predict flotation process. A flotation model considering both surface chemistry parameters and hydrodynamic conditions was first proposed by Yoon and Mao, 1996. The model was further improved by other researchers at the Center for Advanced Separation Technologies
Virginia Tech
at Virginia Tech [3, 15, 16]. The first principle flotation model will be presented in the following chapter. 1.3 Research Objectives The objectives of the present research are to verify and to improve the first principle flotation model developed at Virginia Tech. The main focus of the improvement will be to incorporate the bubble coarsening model to the model, so that it can predict bubble size ratio in the froth phase. To validate the model and simulator, the results of a series of flotation tests conducted by other researchers and reported in the literature will be used as data base for model verification and simulation using the first principle flotation model. The model parameters will be adjusted so that the model prediction and simulation results will be in close agreement with the flotation test results. Once the model has been verified, the computer simulator based on the first-principle model will be used to predict the effects of various parameters that are critically important in industry. The parameters to be studied will include particle size, degree of mineral (or surface) liberation, circuit arrangement, and others. 1.4 Organization The body of this thesis consists of five chapters: Chapter 1 provides background information of flotation and flotation modeling. Several flotation models or simulators are introduced in this chapter that have been developed and applied in the mineral processing industry. This chapter also introduces the research objective of the present work. Chapter 2 presents the model equations for flotation and bubble coarsening in froth phase. The models are developed from first principles, which can help understand the various sub processes occurring in flotation. Chapter 3 presents the results of model validation. A computer simulator is used to validate the model against the results of a series of pilot-scale continuous flotation tests reported in the literature and against a laboratory-scale flotation tests. Chapter 4 presents the simulation results without experimental validation. The flotation model is used to study the effects of various process parameters such as particle size, contact angle, forth height, ζ-potential and others. Furthermore, the simulator is used to see the effect of changing flotation circuits as a means for optimization. Chapter 5 summarizes the results of the work presented in the foregoing chapters and suggests future work. 6
Virginia Tech
Chapter 2: FLOTATION MODEL BASED ON FIRST PRINCIPLES 2.1 Framework 2.1.1 Pulp Phase Recovery Flotation Kinetics Flotation process can be modeled as a first-order rate equation [17, 18], 7 d N d t 1   k N 1 (8) in which k is the rate constant and N is number of particles in a cell. Under a steady condition, k 1 can be determined by [19], k  1 4 S b P (9) where P is probability of flotation and S is bubble surface area flux. b Generally, P is composed of four parts as shown below,   P P P 1P P (10) a c d t where P represents the probability of attachment, P the collision probability, P the probability a c d of detachment in pulp phase, and P represents the probability of bubble-particle transfer at the t pulp-froth interface. In the past, flotation processes were often modeled as a first-order process with a single rate constant for the sub-processes of pulp and froth phase recoveries, which is overall rate constant (k). In the current model, however, the two sub-processes are considered separately and subsequently combined to obtain an overall flotation rate. Basically, flotation process should be considered a second-order process in that its rate should depend on concentration of particles (N ) and bubbles (N ). If one assumes N >>N or N 1 2 2 1 2 remains constant during flotation, the process may be considered a pseudo first-order process and may be represented as d N d t 1   k p N 1   Z 1 2 P (11) where k is the rate constant in the pulp phase, and Z is the collision frequency. Z can be p 12 12 calculated as below [20],
Virginia Tech
8 Z 1 2  2 3 / 2 1 / 2 N 1 N 2 d 21 2 u 21  u 22  (12) which was derived originally by Abramson [21] based on the assumption that particle velocities were independent of fluid flow. In Eq. (12), d is the collision diameter (sum of radii of bubbles 12 and particles), and 𝑢̅ and 𝑢̅ are the root-mean-square (RMS) velocities of the particles and 1 2 bubbles, respectively. Substituting Eq. (11) into Eq. (12), one can obtain, d N d t 1   2 3 / 2 1 / 2 N 1 N 2 d 21 2 u 21  u 22 P  (13) From Eqs. (11) and (13), one can obtain, k p  Z 1 N P2 1 (14) Bubble Generation Model The diameters of bubbles (d ) were calculated using the bubble generation model derived 2 by Schulze [22], d 2   2 .1 3 1 0b lv.6 6  0 .6  (15)  where γ is the surface tension of the water in a flotation cell, ρ is the density of the water, and lv 3 ε is the energy dissipation rate in the bubble generation zone. In the present work, it is assumed b that air bubbles are generated at the high energy dissipation zone in and around the rotor/stator assembly, which has 15-times larger energy dissipation rate than the mean energy dissipation rate (ε) of a flotation cell [22]. RMS Velocities The RMS velocity of the particles is calculated using the following empirical relation [20], 2/3 4/9d7/9    u 0.4 1  1 3  (16) 1 1/3      3 where ε is energy dissipation rate, d particle diameter, ν kinematic viscosity of water, ρ is 1 1 particle density, and ρ is the density of water. 3 On the other hand, the RMS velocity for bubbles is calculated using the following equation [23],
Virginia Tech
K rr V   132 1 2 (20) H 6H  r r  0 1 2 where H is the closest separation distance between bubble of radius r and particle of radius r 0 2 1 in water, and K is the hydrophobic force constant between the bubble and the particle [27]. It 132 has been shown that hydrophobic interaction between hydrophobic solid surface and bubble surface with different contact angles can be predicted using the combining law [19], K  K K (21) 132 131 232 where K is the hydrophobic force constant between two particles in a medium and K is the 131 232 hydrophobic force constant between two air bubbles in a medium [28]. Figure 2.1 [29] shows the relationship between K and particle advancing contact angle, from which one can clearly see 131 that K increases with the increase of the contact angle. In the present work, the values of K 131 131 has been determined using the data presented in Figure 2.1. To calculate P using Eq. (18), it is necessary to know the value of E . The kinetic energy a k may be calculated using following relation, 10 E k  0 . 5 m 1 U H c  2 (22) where m is the mass of the particle, and U is the velocity of the particle approaching a bubble 1 Hc surfaces at the critical rupture distance (H ). In the present work, the following equation has been c used to calculate U , Hc  U H c  U 1 / (23) where 𝑈 is radial velocity of the particle approaching a bubble and  is the drag coefficient in 1 the boundary layer of the bubble [30]. The values of U and  have been determined as described 1 previously [31], b) Probability of Collision (P ) c In the present work, Eq. (24) is used to determine the probability of collision, which is shown below,  3 3  Re d  P c  tanh2   2 1 16 10.249Re0.56     d1 2      (24) where d and d are bubble and particle diameters, respectively, and Re is the Reynolds number. 1 2 Eq. (12) represents a hard-core collision model, that is, bubble-particle collision occurs when the two macroscopic spheres approach each other within the collision radius (r ), which is effective 12 only under extremely turbulent conditions. For quiescent flow, however, the collision is affected
Virginia Tech
by the streamlines around bubbles. The truth may lie in between. Therefore, one may get correct Z by multiplying P [31]. 12 c c) Probability of Detachment (P ) d The probability of detachment is calculated using the following expression [19] 11 P d  e x p   W a E  'k E 1  (25) where W is the work of adhesion, and 𝐸′ is the kinetic energy of detachment. W can be a 𝑘 a obtained from the following relation, 21 1 c o s  2    W a  lv r  (26) where γ is the surface tension of water, r is the radius of the particle, and θ is the contact angle. lv 1 By using Eq. (25), 𝐸′ can be calculated using the following relation [15], 𝑘 ' 0 .5 1  1 2  /  2   E k  m d  d (27) where  is the energy dissipation rate and  is the kinematic viscosity. Pulp Recovery Calculation In a mechanically-agitated cell, the pulp phase recovery, R , can be calculated as below, p k t R  p (28) p 1k t p in which k is the flotation rate constant in the pulp phase. Eq. (28) is applicable for perfectly p mixed flotation cells as is the case with a mechanically-agitated individual cell in a flotation bank. For plug-flow reactors, one may use the relation below to calculate R , p R p  1  e  k tp (29) 2.1.2 Froth Phase Recovery It is well kwon that the forth recovery accounts for two independent mechanisms, i.e. recovery due to attachment and recovery due to entrainment. Fine particles are recovered by entrainment, while coarse and hydrophobic particles are recovered by attachment. Thus, the overall froth recovery (R) can be written as follows [16], f
Virginia Tech
12 R f  R   f   R e  m ax e x p (30) in which R is the maximum fraction of the particles entering the froth phase that is recovered max into a launder,  is the retention time of air in the froth, and R is the recovery of particles due to f e entrainment. Former researchers [15] have developed equations for calculating of R and τ. In e f addition, it can be readily seen that R should vary with bubble coalescence as follows, max R m ax  S S t b  d d 2 2 ,b ,t (31) where S and S are the bubble surface areas on the top and bottom of a froth phase, respectively, t b while d and d are the bubble diameters at the top and bottom, respectively. 2,t 2,b 2.1.3 Overall Recovery Figure 2.2 shows the diagram to calculate the overall flotation recovery [3], in which R p is the recovery in the pulp phase and R is the recovery in the froth phase. According to the f diagram, one can readily find that the overall recovery, R, can be calculated using the equation below, R  R p R R f p  R 1 f  R p (32) Figure 2.2: Diagram of mass balance of materials around a flotation cell. Soni, G., Development and Validation of a Simulator based on a First-Principle Flotation Model. 2013, Virginia Tech. Used under fair use, 2015.
Virginia Tech
Note here that one can calculate the overall rate constant (k) using the equation below under a perfectly mixed flotation cell, 13 k  k p R f (33) By combining Eqs. (28), (32) and (33), one can get the relationship between overall flotation rate constant (k) and overall flotation recovery (R), k t  1 R  R (34) which is the same as the overall rate constant equation under perfectly mixed condition developed by Levenspiel [32]. 2.2 Bubble Coarsening Model In the flotation froth, air bubbles become larger and larger with the increase of froth height, which is due to bubble coalescence phenomenon. Bubble coalescence has significant impacts on the flotation recoveries and grades, as it causes the bubble surface area to decrease along the vertical direction, which forces some particles to drop back to the pulp phase. In the calculation of froth recovery due to attachment, one needs to determine R , which depends on max the bubble size ratio (d /d ). In the past, an assumed value of bubble size ratio was used in Eq. 2,t 2,b (31). At present, a bubble coarsening model has been incorporated in the current flotation model to predict the bubble size ratio accurately. 2.2.1 Bubble Coarsening Foam Model In this section, a 2-phase bubble coarsening model will be introduced, which does not consider the effects of particle size and particle hydrophobicity on the stability of the froth. Therefore, the model to be described below may be applicable for foam rather than froth. Eq. (35) describes the drainage of liquid in a foam, 1  1 A U  gA  (35)  A x where U is the drainage rate; g is the gravitational acceleration; µ, ρ, and γ are the dynamic viscosity, density, and surface tension of water, respectively; A is the cross-sectional area of the plateau border (PB), and x is the distance from the top of the foam. At a steady state, the downward liquid drainage velocity (U) should be equal to the upward superficial gas velocity (-V ), g U   V g (36)
Virginia Tech
Combining Eqs. (35) and (36) and then integrating from the top to the bottom of a foam, one can obtain,       14 A t  V g g t a n  t a n  1  V g A g b    h 2 f g V g (37) where A and A are PB areas at the top and bottom of a foam, respectively, and h is the foam t b f height. In a dry foam, it is reasonable to assume that the number of PBs (N ) is proportional to pb the number of bubbles, which can be calculated by dividing the cross-sectional area (S) of a foam (or froth) column by bubble size, i.e., 4s/πd2. One can then write the following relation, N N p p b ,t b ,b  4 4 s s / / d d 22 ,t 22 ,b   d d 2 2 ,b ,t  2  (38)  where N and N , are the numbers of PB at the top and bottom of a foam, respectively, and d pb,t pb b 2,t and d are the corresponding bubble sizes. 2,b As a foam drains, A decreases with time, or the foam becomes drier. At the same time, the thickness (H) of the lamella films will also become thinner. As H becomes smaller, there will be a critical point where a lamellar film will rupture instantaneous, which is referred as critical rupture thickness of the liquid film (H ). Accordingly, it may be reasonable to suggest that cr bubbles begin to coalesce when A reaches A . As bubbles coalesce, N will decrease. In the cr pb present work, the changes of N is represented by the following relation, pb N p b  N 0 e x p   C A A c r  (39) where N is the number of PBs at the base of a foam, and C is an adjustable parameter. Eq. (39) 0 shows that N decreases exponentially with the square root of A . pb cr Substituting Eq. (39) into Eq. (38), one obtains the following relation, d  1 1  2,b  expC A    (40) d cr  A A  2,t  b t  which shows that bubble size ratio, or bubble coarsening, can be predicted if the values of A , A b t and A are known. cr