+ All Categories
Home > Documents > Modeling local flotation frequency in a turbulent flow...

Modeling local flotation frequency in a turbulent flow...

Date post: 18-Oct-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
13
Modeling local flotation frequency in a turbulent flow field Margaritis Kostoglou , Thodoris D. Karapantsios, Kostas A. Matis Division of Chemical Technology, Department of Chemistry, Aristotle University of Thessaloniki, University Box 116, 541 24 Thessaloniki, Greece Available online 7 August 2006 Abstract Despite the significance of turbulent fluid motion for enhancing the flotation rate in several industrial processes, there is no unified approach to the modeling of the flotation rate in a turbulent flow field. Appropriate modeling of the local flotation (bubbleparticle attachment) rate is the basic constituent for global modeling and prediction of flotation equipment efficiency. Existing approaches for the local flotation rate are limited to specific set of conditions like high or low turbulence. In addition, the combined effects of buoyant bubble rise and/or particle gravity settling are usually ignored. The situation is even vaguer for the computation of collision and attachment efficiencies which are usually computed using the gravity induced velocities although the dominant mode of flotation is the turbulent one. The scope of this work is clear: the development of a general expression for the flotation rate in a turbulent flow field which will cover in a unified and consistent way all possible sets of the problem parameters. This is achieved by using concepts from statistical approach to homogeneous turbulence and gas kinetic theory. © 2006 Elsevier B.V. All rights reserved. Keywords: Flotation process; Mathematical modeling; Stochastic encounters; Turbulent flow field; Bubbleparticle attachment frequency Contents 1. Introduction ............................................................... 79 2. Theory ................................................................. 81 2.1. Deterministic encounters .................................................... 81 2.2. Stochastic encounters ...................................................... 83 2.3. Combined stochasticdeterministic encounters ......................................... 84 2.4. Towards a unified approach for flotation frequency ....................................... 84 2.5. Computational approach ..................................................... 86 3. Resultsdiscussion ........................................................... 87 3.1. Assessment of the collision and attachment efficiency sub-models ............................... 87 3.2. Analysis of the new unified flotation model .......................................... 88 4. Conclusions ............................................................... 90 References .................................................................. 91 1. Introduction Flotation is a very complicated process combining funda- mental hydrodynamics with many elementary physicochemical steps (bubbleparticle interaction forces, particleparticle interaction forces etc.). As for other processes of chemical industry, modeling of flotation is an important step for better understanding the process itself and also a necessary tool for equipment design and optimization. Modeling of the particular process is a very difficult task not only due to the large number of involved phenomena but also due to the wide disparity of their size scales. The statement that flotation is the encyclo- pedia of colloid science(see [1]) is not enough to describe its Advances in Colloid and Interface Science 122 (2006) 79 91 www.elsevier.com/locate/cis Corresponding author. Tel.: +30 2310 99 7767; fax: +30 2310 99 7759. E-mail address: [email protected] (M. Kostoglou). 0001-8686/$ - see front matter © 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.cis.2006.06.014
Transcript
Page 1: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

ace Science 122 (2006) 79–91www.elsevier.com/locate/cis

Advances in Colloid and Interf

Modeling local flotation frequency in a turbulent flow field

Margaritis Kostoglou ⁎, Thodoris D. Karapantsios, Kostas A. Matis

Division of Chemical Technology, Department of Chemistry, Aristotle University of Thessaloniki, University Box 116, 541 24 Thessaloniki, Greece

Available online 7 August 2006

Abstract

Despite the significance of turbulent fluid motion for enhancing the flotation rate in several industrial processes, there is no unified approach tothe modeling of the flotation rate in a turbulent flow field. Appropriate modeling of the local flotation (bubble–particle attachment) rate is the basicconstituent for global modeling and prediction of flotation equipment efficiency. Existing approaches for the local flotation rate are limited tospecific set of conditions like high or low turbulence. In addition, the combined effects of buoyant bubble rise and/or particle gravity settling areusually ignored. The situation is even vaguer for the computation of collision and attachment efficiencies which are usually computed using thegravity induced velocities although the dominant mode of flotation is the turbulent one. The scope of this work is clear: the development of ageneral expression for the flotation rate in a turbulent flow field which will cover in a unified and consistent way all possible sets of the problemparameters. This is achieved by using concepts from statistical approach to homogeneous turbulence and gas kinetic theory.© 2006 Elsevier B.V. All rights reserved.

Keywords: Flotation process; Mathematical modeling; Stochastic encounters; Turbulent flow field; Bubble–particle attachment frequency

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 792. Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

2.1. Deterministic encounters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 812.2. Stochastic encounters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 832.3. Combined stochastic–deterministic encounters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 842.4. Towards a unified approach for flotation frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 842.5. Computational approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

3. Results–discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873.1. Assessment of the collision and attachment efficiency sub-models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873.2. Analysis of the new unified flotation model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

1. Introduction

Flotation is a very complicated process combining funda-mental hydrodynamics with many elementary physicochemicalsteps (bubble–particle interaction forces, particle–particle

⁎ Corresponding author. Tel.: +30 2310 99 7767; fax: +30 2310 99 7759.E-mail address: [email protected] (M. Kostoglou).

0001-8686/$ - see front matter © 2006 Elsevier B.V. All rights reserved.doi:10.1016/j.cis.2006.06.014

interaction forces etc.). As for other processes of chemicalindustry, modeling of flotation is an important step for betterunderstanding the process itself and also a necessary tool forequipment design and optimization. Modeling of the particularprocess is a very difficult task not only due to the large numberof involved phenomena but also due to the wide disparity oftheir size scales. The statement that flotation is “the encyclo-pedia of colloid science” (see [1]) is not enough to describe its

Page 2: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

80 M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

complexity since phenomena like bubble–turbulence interac-tion (leading to particle scavenging) and two-phase flow hydro-dynamics (bubbly flows), which are crucial for the flotationprocess, reside outside colloidal science interests.

This work attempts to develop a general framework for themodeling aspects of flotation in order to present with clarity notonly the current situation but also the future requirements. Amultiscale approach to the problem is necessary since the co-existence of size scales from the intra bubble–particle thin film(b1 μm) to the macroscopic equipment size (order of meters)renders a direct tackling procedure impossible. Even the welldefined problem of a single phase flow is impossible to be solvedsimultaneously over all these size scales. The actual modelingproblem is decomposed to the following size/procedural scales.The first scale (Scale I) is the macroscopic scale. It includes themotion (hydrodynamics) of the three phase (liquid–bubble–particle) mixture at the equipment size scale. The tools forstudying this scale range from multiphase CFD codes ([2,3]) toconventional mixing models accompanied by experimentalcampaigns to assess the mixing characteristics of equipment([4,5]). The second scale (Scale II) is the so-called mesoscopicone. In fact, this is not a distinct size scale but it refers to theprocedure to transfer information from the particle microscopicscale to the equipment macroscopic scale through the appropri-ate descriptive equations. Mean field theories, such as thepopulation balance equation [6], seem to be appropriate tools forthis scale. The third scale (Scale III) is the microscopic one. Thisscale includes phenomena occurring at the bubble/particle sizescale. The particle–bubble collision efficiency, attachmentefficiency and particle–bubble aggregate stability are studiedin this scale. Albeit hydrodynamic aspects at this scale can be ingeneral modeled through a combination of first principles andstatistical theories, modeling of the surface physicochemicalaspects, which are equally important, is far more difficult. Forthis reason, they must be properly parameterized in a way thatpermits the experimental identification of undetermined para-meters from relatively simple experiments. A complete model ofa particular flotation process (industrial or in the laboratory)must include the above three scales.

Several scientific groups have been active over the last yearsin the subject of modeling the flotation process using differentapproaches and putting their attention at different size scalesaccording to the above classification. Yoon and coworkers ([7–9]) are mainly (but not only) focused on the colloidal aspects ofthe flotation process. This group attempts to quantify directly thespecific colloidal force (usual DLVO forces plus the so-calledhydrophobic force) which determines the bubble–particleaggregation instead of using a lumped parameter, i.e. inductiontime, to account for this. Another very important contributioncomes from Nguyen and his coworkers (e.g. [1,10]). The mainfocus of these authors is on the hydrodynamics of scale III. Inaddition, Nguyen has also studied extensively the stability of thefilm between a bubble and a particle as a separate sub-scale [1].A very extensive discussion on the relation between film andbubble scale and the ways of their theoretical description can befound in [11]. Yet, another interesting contribution was made byKing [12] who attempted to expand existing knowledge from

scale III (with somemodifications and improvements) to scale II.For instance, he showed how to tackle problems containingcomplications such as particle size distribution, particle com-position distribution, distribution of induction times, etc. It isimportant to mention that the above authors are among thefounders of the theory of flotation and their contribution to thesubject is not restricted to the particular aspects examined here.A huge amount of work on scale III of flotation has beenperformed by Russian researchers (see Ref. [13]). The results ofthis work have been largely overlooked and many times re-discovered in the Western literature.

At this point it must be noted that the driving force forflotation is the relative motion between bubbles and particles.This can be of deterministic nature (buoyancy driven motion ofbubbles and gravity settling of particles) or of stochastic nature(random motion due to a turbulent flow field). In the generalcase of an industrial flotation process, the two mechanismscoexist and this is the case we investigate in this work.

The approaches described above refer strictly to the case ofdeterministic driving forces for the bubble–particle collision. Inpractical applications, though, the stochastic driving force(induced by turbulent flow) is important and must beconsidered, too. A complete model (including all three scales)has been developed in a series of papers by Bloom and Heindel([14–18]) in order to simulate a laboratory flotation deinkingprocess. Both flotation mechanisms are taken into account bythese authors but their scale II and scale I models are overlysimplistic: a simple population balance is used for scale II and acompletely mixed semi-batch reactor model is used for scale I.An also simple scale II model, based on population balances, iscombined with a very detailed model of scale I, based onmultiphase CFD codes, in the work of Koh and Schwarz ([2,3]).The drawback in the latter work is that only the turbulentflotation mechanism is considered invoking the high level ofturbulence in the employed flotation equipment without,however, any numerical checking of this assumption.

Despite the abundant knowledge that has been accumulatedover the years on the modeling of flotation, a generally acceptedprocedure has not yet been established. The purpose of thepresent paper is to set the basis for this by developing a theoryfor the computation of the “local” flotation rate (scale III) asgeneral as possible which will incorporate the most acknowl-edged of the existing theories. This “local” flotation rate will beused in the future for the development of models for the othertwo scales (I and II).

The structure of the paper is the following: At first, the mostrecent models for bubble–particle collision and attachmentefficiencies under deterministic conditions are reviewed andassessed in order to incorporate them in the general model to bedeveloped. Then, the existing theories for the turbulent en-counter rate are presented and their inconsistencies are pointedout. Composite laws for the turbulent rms (root mean square)velocity are derived next, that should be consistent with allturbulent encounter mechanisms at the corresponding limits. Allthe above approaches are unified using techniques from thekinetic theory of gases. Computational considerations of theproposed model are discussed next. Finally, the appropriate

Page 3: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

81M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

models for the two efficiencies (collision and attachment) arechosen and several aspects of them regarding their contributionto the flotation frequency are extensively discussed.

2. Theory

The scope here is to develop an expression for the “local”particle–bubble aggregation rate under the influence of gravityand an arbitrary level of turbulence. To do this, we will start byreviewing the existing approaches, first for the case ofdeterministic collisions between a particle and a bubble andthen for the more complicated and less studied case of stochasticturbulence induced collisions. The analysis here will berestricted to the case of inertialess particles and immobile(completely retarded) bubble surface. It is noted that only themost recent editions of the several existing approaches for thedeterministic collisions will be examined. A detailed history andreview of most of these approaches can be found in [1,11,13].

2.1. Deterministic encounters

The unit problem here is to find the frequency of aggregationbetween bubbles having a deterministic velocity Ub and radiusRb and particles having a deterministic velocity up (in theopposite direction) and radius Rp. According to well knownconsiderations from the theory of colloidal particle aggregation,the aggregation (flotation) frequency (which can be transformedinto rate by multiplication with bubble and particle concentra-tions) is given as:

KdðRb;Rp;Ub; upÞ ¼ PkðRb þ RpÞ2ðUb þ upÞ ð1Þ

where P is the fractional flotation efficiency i.e. the fraction ofparticles contained at the fluid volume that are scavenged by thebubble which will be permanently attached to the bubble. Theefficiency P consists of two components i.e. P=PcPa. The first(Pc: collision efficiency) expresses the probability of one of theaforementioned particles to actually collide with the bubble (inthe sense that only a thin liquid film separates the particle fromthe bubble after the impact). The second (Pa: attachmentefficiency) expresses the probability of the collided particle todrain the thin liquid film and eventually aggregate with thebubble. According to the standard approach, the efficiency Pc isrelated to particle scale hydrodynamics whereas the efficiencyPa is related to the film scale hydrodynamics and particle–bubble physicochemical interactions.

The collision between a bubble and a particle is the result of acombination of the interception phenomenon (finite particle size)and the tendency of particles to deviate from fluid streamlines(due to their velocity up and/or their inertia). To calculate thecollision efficiency, the fluid flow field around the bubble isneeded and the way of computing it is a crucial part of thecalculation of Pc.

The most recent theories for the computation of Pc and Pa

are the following:i) modification of the classical Yoon-Luttrell [19] approach

by King [12], (henceforth model i).

The Reynolds number for the flow around a bubble (denotedas Reb) is usually of an intermediate value between 1 and 100.This means that the analytical solutions for the flow field forReb=0 (Stokes flow) and for potential flow (good representationfor large values of Reb since then only the flow field towards thebubble contributes to the collision efficiency) cannot be used inrealistic situations. But an analytical flow field is a prerequisiteto obtain the efficiencies Pc and Pa in analytical form, as well.So, Yoon and Luttrell determined an analytical flow field for anarbitrary Reb by interpolating between the above asymptoticanalytical flow fields. The interpolation parameter was obtainedby fitting the analytical flow field to published experimental datafor the streamlines. Using their analytical flow field, Yoon andLuttrell [19] estimated Pc and Pa solely due to interception, inthe limit of a small ratio Rp/Rb. The collision efficiency forcombined interception and gravity and an arbitrary ratio Rp/Rb

using the Yoon and Luttrell's flow field was provided by King[12]:

Pc ¼ 1þ upUb

� �−12 1þ Rp

Rb

� �−2FðRb þ RpÞ þ up

Ub

" #ð2Þ

where

FðxÞ ¼ 12

xRb

� �2−3a4

xRb

þ 3a4−12

� �Rb

xð3Þ

a ¼ exp −4Re0:72b

45

x=Rb−1x=Rb

� �ð4Þ

The parameter α used by King is not equal to the original oneof Yoon and Luttrell which, however, can be restored byexpanding the exponential term of Eq. (4) in a Taylor series andretaining only the first term (i.e. x /Rb−1b1). It must bementioned that the model of King refers only to the case of animmobile bubble surface (because the fluid flow profilecorresponds to a zero slip velocity). The justification of thisassumption is that due to the large amount of surfactants em-ployed in practical flotation processes, bubbles behave (from afluid dynamics point of view) as solid particles.

The attachment efficiency is computed as follows: First, thecumulative distribution of the residence time (sliding time inflotation language) of the particles on the bubble surface iscomputed. Then, the efficiency Pa is given as the fraction of thecolliding particles having a sliding time larger than the inductiontime tind. The induction time is a parameter which collectivelyincludes all physicochemical interactions between a bubble anda particle and must be determined experimentally or from empi-rical relations. The direct determination of the induction time bydirect observation of the particle trajectories is very difficultespecially for the case of relatively small particles for which thetrajectory cannot be separated in two parts: one out of the bubbleand another sliding on the bubble. In any case, the induction timeis just a phenomenological parameter and it can be determinedby fitting the above models to experimental attachment effi-ciencies. After some algebra, the following expression for the

Page 4: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

82 M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

attachment efficiency can be derived based on equations (9.20)and (9.27) of [12]:

Pa ¼ 1−exp

2ðU⁎b þupÞtindRpþRb

� �−1

exp2ðU⁎b þupÞtind

RpþRb

� �þ 1

2664

37752

ð5Þ

where

U⁎b ¼ Ub

32

1−aþ aRp

Rb

� �2 !þ 1

2−3a4

� �Rp

Rb

� �2" #ð5aÞ

ii) model developed by Bloom and Heindel (henceforthmodel ii)

Heindel and Bloom [15] also expanded the approach of Yoonand Luttrell [19] to account for particle settling and arbitrary sizeratio Rp/Rb using the generalized flow field as given by Yoon andLuttrell. Their result for the collision efficiency is [15]:

Pc ¼ 11þ jGj

ð2L3 þ 3L2Þ2ð1þ LÞ3 þ 2

15Re0:72b

ðL3 þ 2L2Þð1þ LÞ4

" #

þ jGj1þ jGj ð6Þ

where G ¼ −upUb

and L ¼ Rp

Rb.

A completely different approach from the usual one describedabove, has been followed by these authors for the derivation ofthe attachment efficiency [16]. A detailed analysis of the forcesgoverning the sliding process of a particle on the surface of abubble was provided. By including the resistive force due to filmdrainage, the gravitational force and the flow force between thebubble and the particle, the evolution equations for the thicknessof the liquid film between bubble and particle were derived.After an extensive mathematical analysis, a closed formexpression for the attachment efficiency was given:

Pa ¼ exp −2kCB

L1þ L

� �gðLÞ−GjkðLÞj−G� �

ðhr−1Þ� �

ð7Þ

where

gðLÞ ¼ 1−3

4ð1þ LÞ−1

4ð1þ LÞ3 !

þRe0:72B

151

1þ Lþ 1

ð1þ LÞ3 −2

ð1þ LÞ4 !

kðLÞ ¼ −

"1−

32ð1þ LÞ þ

1

2ð1þ LÞ3 !

þ 2Re0:72B

151

ð1þ LÞ4−1

ð1þ LÞ3−1

ð1þ LÞ2þ1

ð1þ LÞ

!#

λ is a measure of the deviation of the particle friction factor fpfrom the Stokes flow i.e. fp=λ6πμRp. For Stokesian particles(zero particle Reynolds number, Rep) it is λ=1. The empirical

expression λ=1+0.216Rep1/2 +0.0118Rep which is accurate for

0bRepb1000 is used here [1]. According to this model, thebubble behaves at the bubble size scale as a solid particle so theemployed flow field corresponds to an immobile bubble surface.On the other hand, at the scale of the bubble–particle film, thesurface mobility of the bubble may be important so it influencesthe efficiency Pa through the parameter CB. This parametertakes values between 1 (rigid surface) and 4 (fully mobilesurface), depending on the degree of surface mobility.

In this model the well known induction time tind has beenreplaced by another parameter, hr, which contains all thephysicochemical features of the system. This parameter is theratio of the initial thickness (at the moment of collision) of thebubble–particle liquid film to its thickness at the moment ofdisruption and formation of a particle–bubble aggregate. It is apurely empirical parameter and must be obtained (like tind) fromexperiments or empirical relations.

iii) model developed by Nguyen [20] (henceforth model iii)There are two distinct shortcomings in the composite analytical

flow field of [19]. First, it is derived by interpolation between twoflow fields with fore and aft symmetry so it exhibits the same typeof symmetry, too. But this symmetry is not actually obeyed by thereal flow field. The collision angle (the maximum angle of attackon the bubble's surface where a particle can collide) is incorrectlyassumed equal toπ /2 invoking the fore and aft symmetrywhereasthe actual collision angle is smaller. This error in the collisionangle can lead to a serious error in the collision and attachmentefficiency computation. A second important weak point is that theYoon and Luttrell flow field refers to the case of an isolatedbubble. But in practical applications the volume fraction ofbubbles (gas phase) in the liquid is finite, leading to a densificationof liquid streamlines and hence to an increase of the flotation rate.Both the above drawbacks were overcome by Nguyen in aningenious way. He solved numerically the Navier–Stokesequations around a bubble using a particle in “cell” approach(introduced in the context of flotationmodeling in [21]) to accountfor the finite gas holdup. Then the fluid velocities were expandedinTaylor serieswith respect to variable r /Rb−1 and the expansioncoefficients up to the second order were obtained by employingasymptotic results and fitting the numerical solutions of the flowfield for several values of Rb and gas holdup φ (for details see[22]). In this way, an approximate flow field valid in the regionclose to the bubble can be given in closed form as a function ofbubble Reynolds number Rb and gas holdup φ. Using this flowfield, the following relation for the collision efficiency due tointerception and particle settling velocity was obtained:

(8)

Pc ¼ 11þ jGj

Rp

Rb

� �2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðX þ CÞ2 þ 3Y 2

q−ðX þ cÞ

13:5Y 2

�ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðX þ cÞ2 þ 3Y 2

qþ 2ðX þ cÞ

� �2where

c ¼ upUb

Rb

Rp

� �2

ð8aÞ

Page 5: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

83M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

X ¼ 32

1þ 3Reb=16

1þ 0:309Re0:694b

� �þ ð37:515þ 0:006Re1:367b Þu

ð8bÞ

Y ¼ 3Reb=8

1þ 0:217Re0:518b

þ ð0:466Reb−0:443Re0:96b Þu ð8cÞ

The attachment efficiency is derived based on the conceptof induction time but using the new flow field and takinginto account the fact that the collision angle can be differentthan π / 2. The final expression, after a transformation oftrigonometric relations to algebraic for computational conve-nience, is

Pa ¼ 1−p2a1−p2c

� �X þ cþ YpaX þ cþ Ypc

� �ð9Þ

where

pc ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðX þ cÞ2 þ 3Y 2

q−ðX þ cÞ

3Yð9aÞ

and pa is given from the solution of the following transcen-dental equation

ð1−paÞð1−BÞ=2ð1þ BpaÞBð1þ paÞð1þBÞ=2 ¼ Z ð9bÞ

where

Z ¼ ð1−pcÞð1−BÞ=2ð1þ BpcÞBð1þ pcÞð1þBÞ=2 exp −

Ubð1−B2ÞAtindRp þ Rb

� �ð9cÞ

A ¼ upUb

þ Rp

RbX þ Rp

Rb

� �2M2

ð9dÞ

B ¼ Rp

Rb

YAþ Rp

Rb

� �2 N2A

ð9eÞ

X, Y have been already given in Eqs. (8b,c). M, N arefunctions of Reb and φ:

M ¼ −92−27Reb32

þ 0:4531Re1:1274b

−ð71:312þ 2:156Re0:954b Þu1:367

−ð370:374þ 44Re−212:032b Þu1:912

ð9f Þ

N ¼ −0:8748Re1:0562b þ 7:65Re0:993b u0:434−8:755Re0:982b u0:618

ð9gÞ

The above relations were derived for the case of animmobile bubble surface. The corresponding relations for afully mobile bubble surface were also derived in [23] but theyare not given since they will be not used here. Nguyen alsostudied the effect of particle inertia to the collision andattachment efficiencies [1]. For the special case of small

Stokes number, he was able to incorporate the inertia effect inclosed form relations for Pc and Pa. Inertia effects will not beconsidered here since it can be shown that for the case of animmobile bubble surface and particle diameters smaller than40 ìm they are not important. The contribution of inertia to theflotation rate defines the distinction between flotation andmicroflotation which is actually considered here [13]. Anextensive analysis of the effect of the partial retardation(partially immobile bubble) to the flotation process can befound in [13]. The opposite to the present case (significantinertia, mobile bubble surface) has been studied extensively in[24] and [25].

2.2. Stochastic encounters

Despite the significance of stochastic (turbulence in-duced) encounters in practical flotation applications, thismode has not been studied from the modeling point ofview at the same depth as the deterministic mode. In orderto understand and quantify the influence of turbulence onthe encounter rate between bubbles and particles, thephenomenological statistical theories of turbulence will beemployed [26,27]. From this standpoint, turbulence inducesencounters between suspended particles/bubbles in twodistinct ways. Particles with size smaller than the smallesteddy of the flow field follow exactly the fluctuating localfluid velocity and this motion leads to encounters of firstkind. The frequency of these encounters (turbulent mech-anism I) is given as

KtIðRp;RbÞ ¼ 1:3ðRp þ RbÞ3 em

� �1=2ð10Þ

where ε is the turbulent energy dissipation rate and ν is thekinematic viscosity of the fluid. The above expression hasbeen derived for the first time by Saffman and Turner [28].Latter, several other researchers used it with a differentnumerical constant in place of 1.3.

Particles of larger size (spanning several eddies) exhibit inertiawith respect to turbulent flow fluctuations leading to a motion ofparticles different to that of the fluid. This motion constitutes thesecond mechanism of turbulent aggregation (turbulent mecha-nism II). The corresponding frequency was derived by Abra-hamson [29] as

KtIIðRp;RbÞ ¼ 5ðRp þ RbÞ2ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiu2tp þ U 2

tb

qð11Þ

where utp, Utb are the roots of the mean values of the squares(rms) of the particle–fluid and bubble–fluid relative velocities,respectively. These velocities are given according to [30] as:

utp ¼ 0:685e4=9R7=9p

m1=3qp−qfqf

� �2=3

ð12aÞ

Utb ¼ 0:685e4=9R7=9b

m1=3qf−qbqf

� �2=3

ð12bÞ

Page 6: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

84 M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

where ρp, ρb, ρf are the respective particle, bubble and fluiddensities. The two turbulent flotation mechanisms (I and II)are usually considered to dominate at low and high turbulenceintensity, respectively. The attempts to incorporate thecollision and attachment efficiencies to a turbulent flotationmodel are very limited in literature. Recently, Koh and Swarz([2,3]) used the efficiency functions of Yoon and Luttrell(derived for deterministic velocities) computed at thecharacteristic turbulent velocities given above (Eqs. (12a,b)),combined with the encounter frequency (Eq. (11)). Pyke et al.[31] used Eqs. (12a,b) for turbulent flotation (with coefficient0.565 instead of 0.685) but they computed the efficiencyfunctions using the deterministic velocities. These authorswonder about their approach and the combination of turbulentencounter model with deterministic efficiency functions as itis obvious in their characteristic phrase: “As far as we areaware, the literature is silent both experimentally andtheoretically on these issues at present, so that there is atask defined for the future”. Sherrell and Yoon [32] computedcollision efficiencies in a turbulent flow field based on theenergetic approach (comparison between kinetic and interac-tion energies). This approach is very different from theparticle trajectories approach employed by the otherresearchers.

It is important to note that the turbulent inertia encountermechanism should not be confused with the deterministicinertia encounter mechanism: the first one is a source ofrelative particle–bubble motion whereas the second one justenhances the deterministic collision efficiency by permittingparticles to cross fluid streamlines around the bubble.

Eqs. (12a,b) were derived based on the balance betweenthe inertial subrange acceleration and Allen's drag law. Thisbalance may be appropriate at least for large bubbles (butwith a constant value 0.83 instead of 0.685, according to[1]) but it cannot be used for particles. The appropriatebalance for the particles is between dissipative subrangeacceleration and Stokes drag law leading to ([1]):

utp ¼2eR3

p

135m2qp−qfqf

� �2=3

ð13Þ

The above relation can also be used for the case of smallbubbles instead of Eq. (12b):

Utb ¼ 2eR3b

135m2qf−qbqf

� �2=3

ð13aÞ

2.3. Combined stochastic–deterministic encounters

The effect of the combined turbulent (mechanism II) anddeterministic particle motion on the aggregation frequency hasbeen addressed in [29]. Recently, Bloom and Heindel [17]rederived their final result for the particular application offlotation. The derivation was based on concepts and numericalintegration techniques borrowed from the kinetic theory ofgases. According to the latter authors, the frequency ofencounters between bubbles and particles having deterministic

velocities Ub and up and turbulent rms velocities Utb and utprespectively, are:

KtII;dðRp;RbÞ ¼ 5ðRp þ RbÞ2ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiu2tp þ U2

tb

qexp −

12

ðUb þ upÞ2U2

bt þ u2pt

!

þ kðRp þ RbÞ2ðUb þ upÞ2 þ U 2

bt þ u2ptUb þ up

!

� erfUb þ upffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2ðU 2

bt þ u2ptÞq

0B@

1CA

ð14Þ

There is a subtle point in the above equation. In the case ofzero turbulent velocities, the first term goes to zero and Eq. (14)degenerates to Eq. (1) as it should. For the case of zerodeterministic velocities, the first term is equal to Eq. (11) andaccording to [29] and [17] the second term goes to zero based onthe fact (as the above authors argue) that erf(x)→0 faster thanx→0. Yet, this is not true since the value of erf(x) /x in thelimit x→0 is the finite number 2 /π0.5. The correct asymptoticexpansion of Eq. (14) in the limit of zero deterministic velocitiesleads to an equation similar to Eq. (11) but with a coefficient 7.5instead of 5 (i.e. an error of 50%). This means that theintegration performed in [29] is not correct.

The only attempt known to us to incorporate the collisionand attachment efficiencies with deterministic and stochasticvelocities to a unified model for the flotation rate was made byBloom and Heindel [18]. They used Eq. (14) with their relationsfor the two efficiencies (Eqs. (6) and (7)) computed for thedeterministic velocities. The drawback of the procedure isobvious: their efficiencies refer to deterministic velocities evenin case where turbulent encounters dominate.

2.4. Towards a unified approach for flotation frequency

From the above it is apparent that despite the large amount ofwork devoted in developing expressions for the flotation rate, atthis moment there is no consistent relation of flotation rate forbubbles with diameters 100–1000 μm, particles with diametersup to 40 μm, arbitrary intensity of turbulence and arbitrary gasholdup. The above expressions suffer from inconsistencies,errors in derivation or restricted domain of validity. Using acoherent approach, the existing theories need be corrected,assessed and selected in order to be merged in a unifiedexpression for the flotation frequency.

The equation for the bubble turbulent rms velocity Eq. (12b)is valid only for large bubbles and high values of ε. As a result,it predicts unrealistically large values of bubble–fluid relativemotion for small bubble sizes. Also, the exponent of ε is smallerin Eq. (12b) than in Eq. (10) which is a paradox since theturbulent encounter mechanism I refers to low turbulenceintensity and mechanism II to high turbulent intensity. In orderto get a general expression for the velocity Utb valid for anybubble size and any turbulence intensity and leading to anencounter frequency smaller than that of turbulent mechanism I

Page 7: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

85M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

as ε decreases, the harmonic average of the relevant expressionsfor the two regimes (inertia and dissipative sub-ranges ofturbulence) is employed:

Utb ¼ 0:83e4=9R7=9b

m1=3qf−qbqf

� �2=3 !−1

þ 2eR3b

135m2qf−qbqf

� �� �−124

35−1

ð15Þ

Eq. (10) can be obtained in several ways. Its derivation in[28] was based on the mean flux of fluid entering a spherewith radius Rp+Rb. The relative velocity of a bubble–particlepair follows the mean motion of fluid. An alternativederivation was based on the diffusion equation together withthe appropriate expression for a spatially dependent turbulentdiffusivity [33]. Fortunately, a third derivation procedurebased on the kinetic theory of gases was proposed in [34]. Inthe latter, the stochastic quantities are not the relative particle/bubble–liquid velocities as in the case of mechanism II ofturbulent encounters but the relative particle–bubble velocity.Since both modes of turbulent encounters can be described interms of kinetic theory, this approach is appropriate to derive acomposite law. The turbulent rms velocity for the relativeparticle–bubble velocity is given (employing a statisticalapproach to turbulence) as [28]:

Wtr ¼ ðRp þ RbÞ e15m

� �1=2ð16aÞ

The above velocity refers to distances Rp+Rb smaller thanthe microscale of turbulence and leads to the encounter rategiven by Eq. (10). There is an additional inconsistency at thispoint. As the bubble radius increases, the velocity Eq. (16a)and the corresponding rate Eq. (10) increase without boundsleading to domination of the mechanism I encounters over themechanism II encounters. But this behavior is unacceptable onphysical grounds and it is due to the fact that the above velocitycan be used only for relatively small bubbles. A generalizationof the theory can be made by considering the structure factor ofturbulence for the inertia regime (distance Rp+Rb larger thanthe Kolmogorov microscale) and matching the resultingvelocities from the two regimes at the bubble–particle distanceequal to Kolmogorov microscale. Thus, the velocity Wtr isgiven by Eq. (16a) for (Rp+Rb)≤ν3/4 ε−1/4 and by

Wtr ¼ ðRp þ RbÞ1=3e1=3 115

� �1=2

for ðRp þ RbÞNm3=4e−1=4

ð16bÞ

Up to this point, consistent expressions for the rms velocities ofthree processes have been derived. Namely these are: (1) bubble–liquid relative motion, (2) particle–liquid relative motion (bothdue to bubble/particle inertia) and (3) bubble–particle relativemotion due to liquid small scale flow distribution. The threeprocesses can be assumed independent from each other, soconsidering a normal distribution of velocity for each one of thethree modes of motion and following the principles of kinetic

theory leads to the following total frequency of turbulentencounters:

Kt ¼ kðRp þ RbÞ2ð2kÞ9=2ðUtbutpWtrÞ3

Z Z Z Z Zl−l

Z Z Z ZC

exp −U2

x þ U2y þ U 2

z

2U 2tb

−u2x þ u2y þ u2z

2u2tp−W 2

x þW 2y þW 2

z

2W 2tr

!

dUxdUydUzduxduyduzdWxdWydWz

ð17Þ

where C is the total relative velocity between bubbles andparticles computed as

C ¼ ½ðUx−ux þWxÞ2 þ ðUy−uy þWyÞ2

þ ðUz−uz þWzÞ2�1=2ð18Þ

Although at first sight the above expression seems complex,it takes only an integration over the velocity distribution func-tions of velocity C in order to find the average total bubble–particle relative velocity. To our knowledge, this expression isthe most general one until now in literature for the frequency ofturbulent encounters and it is the starting point for addingmore mechanisms. The addition of the deterministic motion ofbubbles and particles with velocities Ub and up (gravity inducedmotion has only z-components) can be made easily by propermodification of the velocity magnitude C:

C ¼ ½ðUx−ux þWxÞ2 þ ðUy−uy þWyÞ2

þ ðUz þ Ub−uz þ up þWzÞ2�1=2ð19Þ

Abrahamson [29] in his derivation modified the probabilitydistribution of z-velocities, instead of the relative velocity C,to account for the deterministic motion. It can be shown thatboth approaches lead to the same results. On the other hand,Bloom and Heindel [17] modified both probability distribu-tions and C, deriving an incorrect expression for the encounterfrequency (their Eqs. (14), (21)–(23)). Strangely, their finalclosed form expression (their Eq. (24)) is the same with that ofAbrahamson and does not correspond to their previouslyderived equations.

Eq. (17) is not amenable to analytical integration, so anumerical integration procedure is employed. But the numericalintegration of a 9-fold integral is very cumbersome and a wayfor reducing the dimensionality must be found. The proposedprocedure here is to permit some degree of correlation betweenthe bubble–liquid and the bubble–particle modes of motion andto merge them in a single mode. The rms of this new combinedmode of motion is given according to the laws of statistics as:

Ucb ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiU 2

tb þW 2tr

qð20Þ

This assumption reduces the 9-fold integration to a 6-foldone and makes the numerical computation of the encounterfrequency manageable. Of course this assumption introduces anerror which becomes larger as the ratio of the rms of the twomerged variables gets closer to 1. It is very fortunate that in thepresent case the error of merging is negligible since the two

Page 8: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

86 M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

merged modes do not act simultaneously. For large bubbles Utb

is large whereas Wtr is very small and vice versa.The next step is to incorporate the theories for collision and

attachment efficiencies to the developed generalized frameworkfor encounter frequency between bubble and particles. Theseefficiencies depend on the particle and bubble velocities as theyapproach to each other and it is not clear what value of thesevelocities corresponds to the general case considered here. Bothdeterministic and turbulent rms velocities have been proposed([18,2] respectively) but a more general approach is needed. Thekinetic theory decomposes the random encounter bubble–particle velocity to random events with deterministic velocities.But each event must have its collision and attachment efficiencycorresponding to its particular velocities. In this way, the theoriesfor attachment and collision efficiencies under deterministicencounter conditions can be incorporated to the kinetic theoryapproach and integrated over the velocities distributions to givethe total flotation frequency.

The particle deterministic motion influences the two efficien-cies in a different way than the bubble motion. This is due tothe fact that the gravitational motion of a particle crosses thestreamlines of the flow field created by the motion of a bubble.The reason for which Wtr is decided to merge with the bubblevelocity Utp and not with the particle velocity utp is that itcorresponds to a liquid motion similar to that created by thebubble and not to a particle motion which can cross the liquidstream lines. This choice is critical for the correct evaluation of thecollision and attachment efficiencies. The collision efficiency forthe case of mechanism I is usually computed on the premise of anextensional flow field around the bubble ([35,36]) but the uniformflow field employed here permits a large degree of integrationwithout sacrificing much accuracy. Summarizing all the above,the final expression for the composite flotation frequency is:

Kc ¼ kðRp þ RbÞ2ð2kÞ3ðUcbutpÞ3

Z Z Z Zl−l

Z ZPC

exp −U2

x þ U 2y þ U 2

z

2U 2cb

−u2x þ u2y þ u2z

2u2tp

!

dUxdUydUzduxduyduz

ð21Þ

where

P ¼ Pcð½U2x þ U 2

y þ ðUz þ UbÞ2�1=2; ½u2x þ u2y þ ðuz−upÞ2�1=2Þ� Pað½U2

x þ U 2y þ ðUz þ UbÞ2�1=2; ½u2x þ u2y

þ ðuz−upÞ2�1=2Þ ð22Þ

C ¼ ½ðUx−uxÞ2 þ ðUy−uyÞ2 þ ðUz þ Ub−ðuz−upÞÞ2�1=2 ð23ÞAnyone of the deterministic models presented in the previous

sections can be employed for the computation of Pa and Pc.

2.5. Computational approach

The numerical computation of the integral in Eq. (21) isimpossible by direct discretization techniques due to the large

number of integration variables and their infinite range. AMonteCarlo integration approach is necessary. Even this approachconverges very slowly for the particular type of integrand whichtakes values among several orders of magnitude. Fortunately, thetype of integral in Eq. (21) is ideally suited for the easyapplication of an importance sampling technique instead of theslowly convergent random sampling [37]. The exponential termin the integrand in Eq. (21) is eliminated with the payoff to usenormally deviating random numbers. The procedure is asfollows: six velocities (Ux, Uy, Uz, ux, uy, uz) are chosen from sixnormal distributions having zero mean and rms (Ucb, Ucb, Ucb,utp, utp, utp), respectively. The value of the product PC iscomputed for these random velocities. The procedure is repeatedNMC times for different sets of random velocities and the averagevalue of PC is used for the computation of flotation frequencyfrom Kc=π(Rp+Rb)

2(PC)ave. The algorithm for random numbergeneration following the normal distribution is taken from [38].

The above procedure can lead to the computation of flotationfrequency with arbitrary accuracy but due to its stochastic natureand the large number of required computations it is not recom-mended for the estimation of the local flotation frequency in largescale flotation simulations. For this aim, approximate simplerexpressions for computing Kc must be derived. The followingapproach is proposed here: At first, the bubble velocities Ux, Uy

are merged to the new velocityUxy and the particle velocities ux, uyare merged to the velocity uxy. The procedure is the same as wasapplied to the merging of the bubble–liquid and bubble–particlemodes of motion but at the present case the error can be muchhigher since the two merging components attain similar values.The rms of the new velocities are Ucb,xy=2

βUcb and utp,xy=2βutp.

The nominal value of β is 0.5 but since a highly approximatingprocedure has been applied, it is better to find β from fitting theapproximate values of Kc to the exact ones. This is done byrequiring the 4 dimensional and the 6 dimensional approaches toKc (both computed via the Monte Carlo method) to yield the samevalue for zero deterministic velocities. The best choice is found tobe β=0.65. Having reduced the dimension of integration to 4, themultidimensional Hermitte integration scheme (of order NH) isemployed which automatically takes into account the infiniteintegration domain and the exponential weighting functions.

Kc;app

kðRp þ RbÞ2¼ k−2

XNH

i¼1

XNH

j¼1

XNH

k¼1

XNH

m¼1

wiwjwkwmPc;ijkmPa;ijkmCijkm

ð24Þwhere

Pc;ijkm ¼ Pcð½2h2i U 2cb;xy þ ð

ffiffiffi2

phjUcb þ UbÞ2�1=2;

½2h2ku2tp;xy þ ðffiffiffi2

phmutp þ upÞ2�1=2Þ

ð24aÞ

Pa;ijkm ¼ Pað½2h2i U 2cb;xy þ ð

ffiffiffi2

phjUcb þ UbÞ2�1=2;

½2h2ku2tp;xy þ ðffiffiffi2

phmutp þ upÞ2�1=2Þ

ð24bÞ

Cijkm ¼ ½ðffiffiffi2

phiUcb;xy−

ffiffiffi2

phjutp;xyÞ2 þ ð

ffiffiffi2

phkUcb

þ Ub−ðffiffiffi2

phmutp−upÞÞ2�1=2 ð24cÞ

Page 9: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

Fig. 1. Ratio between collision efficiency Pc from models (i) and (iii) to collisionefficiency of model (ii) versus particle diameter Dp for several bubble sizes.

87M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

where hi and wi (i=1,2,..NH) are the Hermitte integration pointsand weights, respectively, which can be found in tabular form forseveral values of NH (e.g. [39]).

The particle settling velocity is computed as [1]:

up ¼ up;Stokes 1þ Arp96

ð1þ 0:079Ar0:749p Þ−0:755� �−1

ð25Þ

where the Stokes settling velocity is given as

up;Stokes ¼2R2

pðqp−qf Þg9l

and the particle Archimedes number Arp is given as

Arp ¼8R3

pðqp−qf Þqfgl2

The bubble rising velocity is computed in a similar manner[1]:

Ub ¼ Ub;Stokes 1þ Arb96

ð1þ 0:079Ar0:749b Þ−0:755� �−1

ð26Þ

with Ub;Stokes ¼ 2R2bqfg9l

and Arb ¼ 8R3bq

2f g

l2

This expression is derived assuming the bubble as a sphericalsolid particle and, therefore, is not of general validity. Never-theless, for bubbles with a diameter smaller than 1000 μm(considered here) and for contaminated water it constitutes a verygood approximation [1]. It is noted that the above velocity refersto an isolated bubble. The finite gas holdup must be taken intoaccount as a correction to this velocity (e.g. via Richardson–Zakiindex).

3. Results–discussion

The general expression for flotation frequency given by Eq.(21) includes all the earlier expressions for frequencies (isolatedmechanisms) as asymptotic cases. For example, by setting theturbulent energy dissipation rate equal to zero, Eq. (21) issimplified to Eq. (1). In case of zero deterministic velocities, andwith collision and attachment efficiencies equal to 1, Eq. (21)degenerates to Eq. (11) for large bubbles and high turbulentintensity or to Eq. (10) for very small bubbles and low turbulenceintensity. Furthermore, for the expressions (10) and (11) it isassumed that the motion of both bubbles and particles obeys thesame mechanism whereas the generalized Eq. (21) can accountfor the case of a bubble size corresponding tomechanism II and aparticle size corresponding to mechanism I, a situation whichprevails under practical conditions. The consistent incorporationof all the mechanisms in one composite expression is the majoroutcome and novelty of the present work.

3.1. Assessment of the collision and attachment efficiency sub-models

Next, the question ariseswhich of the three cited recentmodels(referred as i, ii, iii) for the efficienciesPa andPc should be used in

Eq. (21). An assessment of these theories and comparison amongtheir results is performed here in order to make the choice.

It is noted that according to model (iii), Pc increases with nobound as Rp/Rb increases. This is evidently wrong since theefficiency cannot become larger than unity. The reason for thisinconsistency is that the model has been derived at the limit Rp/Rb≪1. In principle, this is not so bad since the whole theory forcollision efficiency (assuming particles in a flow field induced bya bubble motion and separate handling of bubbles and particles)rely implicitly on this assumption. On the other hand, the deri-vation at the above limit leadsmodel (iii) to have a large difference(of the order Rp/Rb) in the opposite direction compared with theother models even for values of the radii ratio as small as 0.1. Twosources of errors appear when it is attempted to use the model(iii) for larger values of the radii ratio. At first, the capture radius isRp+Rb and not Rb as it has been assumed in the model'sderivation. The error can be easily corrected by replacing Rb withRp+Rb in Eqs. (8,8a). This modification is assumed to be part ofthe model (iii) in what follows. With this simple correction, thevalues of Pa are restricted between 0 and 1. The second source oferror is the flow field which results from a near surface expansion.This means that although the employed flow field is indeed moreaccurate than the other models flow fields for Rp/Rb≪1, itsaccuracy diminishes as the ratio increases and gradually turns tobe even less accurate from the flow fields used by the othermodels. Unfortunately, there is no easy way to correct the flowfield as Rp/Rb increases.

In general, the collision efficiencies predicted by the threemodels are close to each other so in order to stress the dif-ferences, the ratios Pc,i/Pc,ii and Pc,iii/Pc,ii are shown in Fig. 1versus particle diameter Dp (1–40 μm) for several bubblediameters Db (between 100–1000 μm). In the calculations, theparticle density is taken as ρp=2 gr/cm

3. The models (i), (ii) givethe same Pc forDb=100 and 300 μm (small Reb number). As thebubble size and particle size increase the discrepancy betweenthe two models increases (due to the different form of the fittingparameter α for the composite flow field they use) but it does notexceed 3% for the parameters studied here. The model (iii) ingeneral is expected to predict smaller values of Pc than the othermodels since it takes into consideration the fact that the collisionangle can be different than π / 2. The expected behavior of themodel is reasonable on physical grounds for small particle size

Page 10: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

Fig. 3. Attachment efficiency versus bubble diameter Db: (1)model (i), t-

ind=0.1 s, Dp=10 μm (2) model (iii), tind=0.1 s, Dp=10 μm (3) model (ii),hr =1.98, Dp=10 μm (4) model (i), tind=0.02 s, Dp=30 μm (5) model (iii), t-

ind=0.02 s, Dp=30 μm (6) model (ii), hr =1.45, Dp=30 μm.

88 M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

but the increase of Pc,iii/Pc,ii as particle size increases is strangeand may be due to the localized (near surface expansion) natureof the employed flow field. The only comparison known to usbetween any of the models presented here is the one performedin [15] between models (ii) and (iii). The curves Pc vs Rb fromthe two models are compared for a case having Rp/Rb≪1 for allRb values. For small values of Rb, it is Pc,iiNPc,iii but theinequality is reversed as the bubble size increases. This behavioris fully compatible with the comparison shown in Fig. 1. Thesmaller values of Pc predicted by the model (iii) (due to acollision angle different than π / 2) are essential to fitexperimental results for collision efficiency (from [40]) whichcannot be fitted accurately using the model (ii). In addition,model (iii) is the only one which takes into account the finite(non-zero) gas holdup. This is a crucial requirement since thelocal gas holdup in the flotation equipment is usually not zero soflotation frequencies derived for isolated bubbles cannot beused. The modified model (iii) is chosen for the calculation of Pc

due to its important features of a finite gas holdup and a non π / 2collision angle. Nevertheless, a few percent error must beexpected as the ratio Rp/Rb takes values larger than 0.1.

The next step is to assess the models for the attachmentefficiency. Let us first compare the two models which have asfundamental parameter the induction time, tind. The attachmentefficiency computed by the models (i) and (iii) are shown versusthe induction time for two pairs of particle and bubble diametersin Fig. 2. As expected the attachment efficiency decreases as theinduction time increases since the time spent by the particle onthe bubble surface is not enough for permanent attachment. Forsmall bubbles the Reb is small and the actual flow field is close toexhibit a fore and aft symmetry. In this case the predictions of thetwo models for Pa are very close. However, for large bubbles(Db=1000 μm), the simplified flow field employed by model (i)is nomore valid and the twomodels lead to very different results.

In order to assess the model (ii) and to check whether it isequivalent to the other models (i.e., is hr just a function of tind?)or it exhibits a completely different behavior, the comparison ofthe predicted Pa among the three models versus Db forDp=10 μm (tind=0.1 s, hr =1.98) and Dp=30 μm (tind=0.02 s,hr =1.45) is shown in Fig. 3. Obviously, Pa,iii decreases as bubblesize increases due to the reasons already mentioned. Theefficiencies for the two other methods show a smaller variation

Fig. 2. Attachment efficiency Pa from models (i) and (iii) versus induction timetind for several pairs of bubble and particle diameters.

(in opposite directions) but it is sure that the two methods cannotbe assumed to be equivalent. So, the model (iii) which takes intoaccount the disappearance of the fore and aft symmetry of theflow field as Reb increases and leading to smaller attachmentefficiencies must be used. It is noted that the dependence of theattachment efficiency on the bubble diameter shown in Fig. 3 isnot a general feature but depends on the particular conditions.For example as the particle density increases this dependencedisappears gradually leaving the attachment efficiency almostinsensitive from Db.

3.2. Analysis of the new unified flotation model

The consistency of the expressions used for the turbulentvelocities will be presented through a simple example. Thevelocities (rms) Utb (Eq. (15)) and Wtr (Eq. (16a)) are shownversus bubble diameter for the case Dp=10 μm and three valuesof ε in Fig. 4. The anticipated behavior (not exhibited byprevious models) that as Db and ε increase, Utb dominates overWtr and vice versa is apparent. The particle inertia inducedvelocity utp range in this example from 1.85×10−5 m/s(ε=10 m2/s3) to 1.85×10−7 m/s (ε=0.1 m2/s3) and hasinsignificant contribution to the encounter rate.

The numerical aspects and requirements for the computationof flotation frequency will be examined next. The attachment

Fig. 4. Rms of relative bubble–fluid (Utb) and bubble–particle (Wtr) turbulenceinduced velocities versus bubble diameter Db for three values of turbulentenergy dissipation rate ε (Dp=10 μm).

Page 11: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

Fig. 6. Ratio between approximate (Eq. (24)) and exact (Eq. (21)) value of theencounter (i.e. P=1), collision (i.e. Pa=1) and attachment frequency Kc versusbubble diameter Db. (Parameters: Dp=10 μm, Ub=0, up=0, ε=1 m2/s3).

89M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

efficiency model (iii) is somewhat more complicated than theother two as it requires the numerical solution of atranscendental equation for the cosine of the critical angle ofattachment (pa in Eq. (9b)). This is not a problem for isolatedcomputations of Pa but for one computation of the flotationfrequency via Eq. (21) the required numerical solutions of thealgebraic Eq. (9b) are tenths of thousands so the stability androbustness of the solution method is very important. A methodwhich seeks the solution in a specified interval is appropriate inthis case. Here the bisection method is used utilizing theinformation that the critical angle of attachment is between zeroand the collision angle (i.e. pcbpab1). The computed value ofthe average generalized velocity (PC)ave (i.e., flotation frequen-cy divided by π(Rp+Rb)

2) for the case P=1, Ub=0, up=0,ε=1 m2/s3, ρ=2 gr/cm3, Rp=5 μm, Rb=500 μm is shown forseveral Monte Carlo realizations with several numbers NMC ofrandom points in Fig. 5(a). For this particular case (onlyturbulent encounters with P=1) the integration can beperformed analytically to give (PC)ave=Cave=0.3214 m/s.This value is presented as a solid line in the Fig. 5(a) forreference. Apparently, as the number NMC of computationalpoints increases the exact solution is approached with adecreasing scatter around it. For NMC=40,000 the error of asingle simulation is no more than 0.5%.

Next, the same case but with taking into account the collisionand attachment efficiency (tind=0.006 s) is considered and thecorresponding values of (PC)ave are shown in Fig. 5(b). Now,there is no exact solution for reference but the convergence of theMonte Carlo method as NMC increases is obvious. The fact thateven with only 1000 Monte Carlo points, a rational approxima-tion, ∼3%, to the flotation frequency can be found is due to the

Fig. 5. Computed value of average total bubble–particle relative velocity Cave

for several Monte Carlo realizations with different number NMC ofcomputational points (a) ignoring efficiencies (b) including efficiencies(Parameters: Dp=10 μm, Db=1000 μm, Ub=0, up=0, ε=1 m2/s3).

sampling technique used. The number of points that would beneeded for the same accuracy level using random samplingtechniques would be at least an order of magnitude larger. It isimportant to note that as the deterministic velocities increase, thecontribution of the stochastic part of the flotation frequencydecreases so the Monte Carlo integration converges faster (i.e.the most demanding case is that of zero deterministic velocitiesexamined above). Conclusively, a number NMC=40,000 is inany case sufficient for the computation of flotation frequency.

The approximate relation (24) offers a deterministic way forthe computation of flotation frequency. The number of terms NH

must be properly selected in order to keep within acceptablelevels both computational load and accuracy. A rational choiceNH=5 leads to a total of 625 integration points. This level ofcomputational cost permits the use of relation (24) for localflotation frequency in more demanding applications but thequestion about accuracy remains. To make the situation clear theratio of the approximate (Eq. (24)) to the exact (Eq. (21)) en-counter (i.e., P=1), collision (i.e., Pa=1) and attachment (fortind=0.006 s) frequencies for the system described above (forFig. 5), is shown versus bubble diameter in Fig. 6.

The error in the encounter frequency is almost independentfrom Db and it is the discretization error emerging from theattempt to approximate the function C in Eq. (21) in the interval(−∞,∞) using just five collocation points (i.e. the error associatedwith the dimensionality reduction has been already eliminated by

Fig. 7. Collision and attachment frequencies computed by Eq. (24) (exact) andby getting P out of the integral and computed it at the rms velocity values(approximate). (Parameters: Dp=10 μm, Ub=0, up=0, ε=1 m2/s3).

Page 12: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

90 M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

the proper choice of β). It is interesting that the approximationerror for the collision frequency is smaller. This may be due to thefact that the function PcC is more uniform with respect tovelocities than C, leading to larger integration accuracy. Finally,the attachment efficiency is a relatively sharp function of thevelocities so the integration error in the attachment frequencycomputation can be quite large (up to 40% for the present exampleas shown in Fig. 6). A similar behavior is found for other sets ofparameters. Keeping in mind again that the worst case (nodeterministic velocities) for the approximate model has beenexamined here, the final suggestion is that the approximateflotation frequency expression (Eq. (24)) can be used but only forthe case of hydrophobic particles (i.e. tind=0, Pa=1).

A basic feature of the model proposed here is that theefficiencies Pa, Pc are integrated over the velocity distributionsinstead of being computed at the turbulent velocity rms. To assessthe quantitative differences between the two approaches, thecorresponding collision and attachment frequencies for thesystem presented in the previous examples are shown in Fig. 7.The differences between the collision frequencies are in generalsmall but the differences between the attachment frequencies arelarge enough to justify the computational effort needed by theapproach proposed here. It must be noted that the approximationof computing the efficiency functions for a characteristic fluidvelocity (including stochastic and deterministic contributions)with an analytical integration of the remaining terms would be atempting alternative to be explored further but the analyticalintegration is not feasible. It must be recalled here that theintegration performed in [29] is incorrect. The purpose of thiswork is the derivation of the expression for the flotation frequencyin turbulent flowwhich covers (more or less) the region of particlediameter (2–40 μm) and bubble diameter 100–1000 μm and topropose procedures for its exact and approximate computation.Detailed parametric analysis of the model, incorporation of scaleII models and comparison with experimental results will follow.Also,more accurate and simpler approximate expressionsmust bederived since practical applications may require several levels ofadditional integrations [12] (i.e. particle size distribution, bubblesize distribution, induction time distribution, energy dissipationrate distribution).

It is noted here that the typical approach in flotation modelingliterature is the incorporation of a model for particle detachmentfrom the bubble, to the flotation frequency expression [1]. Butaccording to the integrated approach proposed in the introduc-tion, detachment cannot be assigned directly to the flotation(attachment) process. They are two separate processes which canoccur individually. The incorporation of the detachment in theflotation rate is possible for modeling spatial homogeneousprocesses, but this is not so for other cases, e.g., for a region of aflotation equipment with high turbulent intensity and very smallparticle concentration. Obviously in this case, detachment mayoccur without attachment so the detachment rate cannot beconsidered just as a modifier of the flotation rate but it entersdirectly into the macroscopic balances as an individual process.The two phenomena of attachment and detachment will becombined in scale II (using our terminology) and not in scale I asit is the usual practice.

The unified model presented here is based on well knownphenomenological theories on the structure of the turbulence andits effect on bubble–particle interaction. These theories of courseare followed by a large list of assumptions some of which can berelaxed in the future. Any future modification/improvement ofthe partial theories (sub-models) can be directly incorporated tothe proposed general model. Three special points will bediscussed in what follows. As it has been already said the shapedeformation of bubbles smaller than 1 mmmoving by buoyancyis insignificant. But for high level of turbulence the turbulentacceleration may lead to the deformation of even smallerbubbles. This deformation must be taken into account to thetheory of turbulent bubble–particle encounters for the case ofhigh turbulence levels (mechanism II). A second point concernsthe amount of turbulent energy accessible by each mechanism ofstochastic bubble–particle relative motion. Whereas the wholeturbulent energy is available for the mechanism I, it is veryreasonable to consider that the energy of scales smaller than thebubble size does not contribute to the bubble acceleration(mechanism II). This issue needs further development but itsincorporation in a CFD code using the present flotationfrequency model is not an issue: The local turbulent energydissipation rate given by the CFD will be used directly for themechanism I rate but for mechanism II it will not be used directlybut for the reconstruction of the local turbulent energy spectrumwhich will be integrated partially to give the amount of energyparticipating to the mechanism II. A third point is that the abovetheories do not take into account the effect of the existence ofbubbles and particles in the structure of turbulence. This is not sosignificant since regarding mechanism I this effect actually isconsidered in two levels: the macroscopic effect is considered bythe CFD codes and the bubble level effect is accounted throughthe consideration of the flow field around the bubble. As regardsthe mechanism II the effect of the existence of the third phase isnot taken into account but it is not really important since thecontribution of the turbulent inertia motion of particles (stronglyinfluenced by the existence of the bubble) to the flotation rate isinsignificant and the turbulent inertia motion of bubbles onlyslightly influenced by the existence of much smaller particles.

4. Conclusions

In the present work, a composite expression for the flotationfrequency (frequency of successful bubble–particle collisions)in a combined gravitational and turbulent flow field is derived.For bubble and particle diameters in the ranges 100–1000 μmand 2–40 μm, respectively, the new expression incorporatesmost appraised earlier existing theories which, however, have alimited extent of validity and to which the new expression candegenerate as particular asymptotic cases. The particular pro-cesses taken into account is the settling and buoyancy motion ofparticles and bubbles, respectively, the bubble/particle–liquidrelative motion due to bubble/particle inertia and its inability tofollow the high frequency turbulent velocity fluctuation and theparticle–bubble relative velocity due to the small scale structureof turbulence. In addition, the collision and attachment effi-ciencies are based on a flow field around the bubble which may

Page 13: Modeling local flotation frequency in a turbulent flow fieldusers.auth.gr/users/8/3/010438/public_html/tdk/Publications/files/... · general expression for the flotation rate in a

91M. Kostoglou et al. / Advances in Colloid and Interface Science 122 (2006) 79–91

not exhibit fore and aft symmetry and take into account a finitegas holdup. Numerical methods for the exact and approximatecomputation of the composite flotation frequency have beendeveloped and assessed. The proposed expression can be usedfor the computation of local flotation rates in simulating complexflotation process equipment.

References

[1] Nguyen AV, Schulze HJ. Colloidal science of flotation. New York: MarcelDekker; 2004.

[2] Koh PTL, Schwarz MP. CFD modeling of bubble–particle collision ratesand efficiencies in a flotation cell. Miner Eng 2003;16:1055–9.

[3] Koh PTL, SchwarzMP. CFDmodeling of bubble–particle attachments in aflotation cell. Proceedings of centenary of flotation symposium, Brisbane,Australia, June; 2005. p. 201–7.

[4] Mavros P, Lazaridis NK, Matis KA. A study and modeling of liquid phasemixing in flotation column. Int J Miner Process 1989;26:1–16.

[5] Yianatos JB, Bergh LG, Diaz F, Rodriguez J. Mixing characteristics ofindustrial flotation equipment. Chem Eng Sci 2005;60:2273–82.

[6] Ramkrishna D. Population balances. Theory and application to particulatesystems in engineering. San Diego: Academic Press; 2000.

[7] Yoon RH, Mao L. Application of extended DLVO theory, IV: derivationof flotation rate equation from first principles. J Colloid Interface Sci1996;181:613–26.

[8] Mao L, Yoon RH. Predicting flotation rates using a rate equation derivedfrom first principles. Int J Miner Process 1997;51:171–81.

[9] Yoon RH. The role of hydrodynamic and surface forces in bubble–particleinteraction. Int J Miner Process 2000;58:129–43.

[10] Phan CM, Nguyen AV, Miller JD, Evans GM, Jameson GJ. Investigationsof bubble–particle interactions. Int J Miner Process 2003;72:239–54.

[11] Ralston J, Dukhin SS, Mishchuk NA. Wetting film stability and flotationkinetics. Adv Colloid Interface Sci 2002;95:145–236.

[12] King RP. Modeling and simulation of mineral processing systems.London: Butterworth-Heinemann; 2001.

[13] Dukhin SS, Kretzschmar G, Miller R. Dynamics of adsorption at liquidinterfaces, Studies in Interface Science, Series editors D. Mobius and R.Miller, Elsevier, Amsterdam; 1995.

[14] Bloom F, Heindel TJ. A theoretical model of flotation deinking efficiency.J Colloid Interface Sci 1997;190:182–97.

[15] Heindel TJ, Bloom F. Exact and approximate expressions for bubble–particle collision. J Colloid Interface Sci 1999;213:101–11.

[16] Bloom F, Heindel TJ. An approximate analytical expression for theprobability of attachment by sliding. J Colloid Interface Sci 1999;218:564–77.

[17] Bloom F, Heindel TJ. On the structure of collision and detachmentfrequencies in flotation models. Chem Eng Sci 2002;57:2467–73.

[18] Bloom F, Heindel TJ. Modeling flotation separation in a semi-batchprocess. Chem Eng Sci 2003;58:353–65.

[19] Yoon RH, Luttrell GH. The effect of bubble size on fine particle flotation.Miner Process Extr Metall Rev 1989;5:101–22.

[20] Nguyen AV. The collision between fine particles and single air bubbles inflotation. J Colloid Interface Sci 1994;162:123–8.

[21] Rulev NN, Derjaguin BV, Dukhin SS. Flotation kinetics of small particlesby a group of bubbles. Kolloidn Z 1977;39:314–23.

[22] Nguyen AV. Hydrodynamics of liquid flows around air bubbles inflotation: a review. Int J Miner Process 1999;56:165–205.

[23] Nguyen AV. Particle–bubble encounter probability with mobile bubblesurfaces. Int J Miner Process 1998;55:73–86.

[24] Dai Z, Dukhin S, Fornasiero D, Ralston J. The inertial hydrodynamicinteraction of particles and rising bubbles with mobile surfaces. J ColloidInterface Sci 1998;197:275–92.

[25] Ralston J, Dukhin SS, Mishchuk NA. Inertial hydrodynamic particle–bubble interaction in flotation. Int J Miner Process 1999;56:207–56.

[26] Bachelor GK. The theory of homogeneous turbulence. London: Gam-bridge University Press; 1953.

[27] Baldyga J, Bourne JR. Turbulent mixing and chemical reactions. NewYork: Wiley; 1999.

[28] Saffman PG, Turner JS. On the collision of drops in turbulent clouds.J Fluid Mech 1956;1:16–30.

[29] Abrahamson J. Collision rates of small particles in a vigorously turbulentfluid. Chem Eng Sci 1975;30:1371–9.

[30] Schubert H, Bischofberger C. On the optimization of hydrodynamics inflotation processes. Proceedings 13th international mineral processingcongress, Warsaw, vol. 2; 1979. p. 1261–85.

[31] Pyke B, Fornasiero D, Ralston J. Bubble particle heterocoagulation underturbulent conditions. J Colloid Interface Sci 2003;265:141–51.

[32] Sherrell I, Yoon RH. Development of a turbulent flotation model.Proceedings of centenary of flotation symposium, Brisbane, Australia,June; 2005. p. 611–8.

[33] Gruy F, Saint-Raymond H. Turbulent coagulation efficiency. J ColloidInterface Sci 1997;185:281–4.

[34] Yuu S. Collision rate of small particles in a homogeneous and isotropicturbulence. AIChE J 1984;30:802–7.

[35] Melis S, Verduyn M, Storti G, Morbidelli M, Baldyga J. Effect of fluidmotion on the aggregation of small particles subject to interaction forces.AIChE J 1999;45:1383–93.

[36] Pnueli D, Gutfinger C, Fichman M. A turbulent-Brownian model foraerosol coagulation. Aerosol Sci Tech 1991;14:201–9.

[37] Elimelech M, Gregory J, Jia X, Williams R. Particle deposition andaggregation: measurement, modelling and simulation. London: Butter-worth-Heinemann; 1995.

[38] Press W, Flannery B, Teukolski S, Vetterling W. Numerical recipes. Theart of scientific computing 2nd edition. New York: Cambridge UniversityPress; 1992.

[39] Scheid F. Numerical analysis. New York: McGraw Hill; 1968.[40] Nguyen AV, Kmet S. Collision efficiency for fine mineral particles

with single bubble in a countercurrent-flow regime. Int J Miner Process1992;35:205–23.


Recommended