You are on page 1of 16

Chemical Engineering Science 288 (2024) 119812

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Effect of surfactant concentration and surface loading on the dynamics of a


rising particle-laden bubble
Ai Wang , Edwin Banks , Geoffrey Evans , Subhasish Mitra *
ARC Centre of Excellence for Enabling Eco-beneficiation of Minerals, School of Engineering, University of Newcastle, Callaghan 2308, NSW, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Selective separation of both inorganic and organic particles by flotation process involving air bubbles has wide
Flotation, bubble rise velocity application areas which critically depends on the particle laden bubble rise velocity. In this study, the rising
Drag coefficient behaviour of particle-laden bubbles (particle diameter dp ~ 114 μm, bubble diameter dB = 2.88 ± 0.07 mm) at
bubble surface loading (BSL)
various surfactant concentrations (Sodium Dodecyl Sulphate, 0 % − 25 % cmc) was investigated. Bubble ve­
Surfactant
Bubble-particle aggregates
locity, aspect ratio and bubble surface loading (BSL) parameters were quantified by high-speed imaging. A
regime map was proposed to determine the effect of particles and surfactant on the observed bubble surface
mobility condition. Also, an unsteady-state mathematical model based on force balance approach with appro­
priate drag correction factor was developed to predict the rise velocity of a particle-laden bubble. The model
produced reasonable agreement with the experiment specifically in the high surfactant concentration cases
where the bubble shape became nearly spherical.

from bubbles due to system turbulence and bubble coalescence in the


froth phase.
1. Introduction The flotation type separation process described in above can be
envisaged as a typical reactor (pulp phase) and separator (froth phase)
Three-phase systems are ubiquitously encountered in the chemical system wherein the pulp-froth interface can be considered as the
engineering applications including separation processes and multiphase dividing boundary between these two components. Due to turbulence,
reactors. For example, a multitude of industrial solids-liquid separation not all particles in the bulk that attach to bubbles, can be successfully
processes utilise air bubbles to successfully collect and float particles. transferred to the froth phase. Successful recovery occurs when particles
These applications involve extraction of valuable base metals such as are carried by the bubbles to the froth phase without much detachment.
copper, lead, zinc by froth flotation (Schulze, 1989; Lee et al., 2009; Understanding rising behaviour of particle-laden bubbles of different
Jameson and Emer, 2019; Nanda et al., 2022), ion flotation for pre- solids loading levels therefore is critical for successful design and opti­
concentrating precious metals from dilute solutions (Chang et al., mization of these processes as the bubble surface loading (BSL)
2019), wastewater treatment by removing total suspended solids (TSS), parameter is directly related to particle recovery (Eskanlou et al., 2018a,
biological oxygen demand (BOD) and oil and grease contaminants by b). It is reported that BSL can reach up to ~ 0.2 in a mechanical flotation
dissolved air flotation (DAF) (Bennett and Peters, 1988), and raw water cell (Bradshaw and Connor, 1996; Koh and Schwarz, 2008). Neverthe­
purification by removal of microalgae (Chen et al., 1998). This diverse less, in the extreme case, a bubble loaded with heavy particles may
range of applications of the flotation process owe to its high separation indeed descend instead of rising (Uribe-Salas et al., 2003) due to loss of
efficiency, simplicity of the technology and favourable economy. buoyancy. Consequently, rise velocity reduction due to increased bubble
Central to particle collection mechanism in a typical flotation type surface loading decreases recovery rate (Wang et al., 2020). A detailed
separation process includes three sub-steps (Tao, 2004). The first step review of the measurement of BSL parameter and its effect on the bubble
involves close physical interactions of bubbles with particles in presence rising dynamics can be found in Wang et al. (2023).
of background turbulence, the second step involves selective physico­ Bubble rise velocity can be conveniently described by three non-
chemical attachment of hydrophobic particles to bubble interface and dimensional parameters - Galilei number (Ga = ratio of gravitational
rising of bubble-particle aggregates to the free surface into the froth force to viscous force), Eötvös number (Eo = ratio of gravitational force
phase and the third step involves physical detachment of some particles

* Corresponding author.
E-mail address: Subhasish.mitra@newcastle.edu.au (S. Mitra).

https://doi.org/10.1016/j.ces.2024.119812
Received 10 September 2023; Received in revised form 15 January 2024; Accepted 22 January 2024
Available online 24 January 2024
0009-2509/© 2024 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Nomenclature mBP mass of a particle-laden bubble (kg)


NP number of loaded particles (-)
ab semi-minor diameter (m)semi-major diameter (m) ReB, ReBP Reynolds number for a bare bubble and a particle-laden
AR aspect ratio (-) bubble (-)
BSL bubble surface loading (-) Sbubble the surface area of the bubble (m2)
CDB, CDB,P drag coefficient of a bare bubble and a particle-laden UB, UBP terminal velocity of a bare bubble and a particle-laden
bubble (-) bubble (m/s)
cmc critical micelle concentration (-) αBP,U, αBP,D velocity and drag correction factor (-)
Cs surfactant concentration (mol/L) µL dynamic viscosity of liquid (Pa⋅s)
Cv virtual mass coefficient (-) ρL, ρB, ρP liquid, bubble and particle density (kg/m3)
dB bubble diameter (m) σ surface tension (N/m)
dP particle diameter (m) ϕ packing factor of particles on bubble surface (-)
FB,B, FB,P buoyancy forces on bubble and loaded particles (N) Eo Eötvös number (-)
FD,BP drag force on a particle-laden bubble (N) Ga Galilei number (-)
FG,B, FG,P gravitational forces on bubble and loaded particles (N) Mo Morton number (-)
g gravitational constant (m/s2) St Stokes number (-)
J, H dimensionless variables in Clift’s rise velocity model (-)

to surface tension force), and Morton number (Mo = Eo3/Ga4) (Tripathi


et al., 2015). In a given liquid medium, rise velocity of a bare bubble (no
particles attached) depends on its diameter and is not known aprioi.
Determining bubble rise velocity in a typical flotation environment is
even more difficult which is governed by several parameters in addition
to bubble size which include such as particle size, level of bubble surface
loading (BSL), surfactant concentration, gas volume fraction, and the
background flow field. While it is challenging to include all these pa­
rameters in determining the particle laden bubble rise velocity, some of
these parameters such as surfactant concentration (Davis and Acrivos,
1966; Griffith, 1962; Takagi and Matsumoto, 2010) and bubble surface
loading (Yan et al., 2021; Wang et al., 2021) have been considered in the
literature.
In the presence of particles, it has been experimentally reported that
bubble rise velocity decreases significantly. For example, Yan et al.,
(2021) reported decrease in bubble rise velocity by ~ 18 % and ~ 31 %
with bubble surface loading (BSL) value ~ 0.2 and ~ 0.4, respectively.
The observed decrease in the rise velocity can be attributed to increased
bubble drag coefficient because bubble surface becomes more immobi­
lized and behaves similar to a rigid surface in presence of particles. It is
argued that the effect of loaded particles manifests similar to surfactant
in terms of immobilizing bubble surface (Huang et al., 2011; Ireland and Fig. 1. Decreasing trend of surface tension parameter of the aqueous medium
with increasing surfactant concentrations (error bar indicates one standard
Jameson, 2014; Wang et al., 2021, Wang and Brito-Parada, 2021).
deviation varying in the range from ~ 0.091 to 0.38).
For relatively large size bubbles (Ga > 40, Eo > 0.1), interface is
quite deformable which is reflected in surface oscillations when a bubble
rises and goes through various shape changes (Tripathi et al., 2015). In immobilized which increases drag coefficient and therefore reduces
presence of particles on bubble surface, these surface oscillations are bubble rise velocity (Legendre et al., 2009; Hessenkemper et al., 2020).
dampened for the same size of bubbles due to reduction in the apparent In our previous study (Wang et al., 2021), we examined the particle
surface tension of the interface. It is reasoned that decrease in the bubble laden bubble rise velocity in two limiting surfactant concentration cases
surface area in the initial interface deformation phase causes surface - pure RO water (no surfactant) and 20 % cmc of SDS (Sodium Dodecyl
pressure induced by the particles, which in effect decreases the apparent Sulphate) surfactant. In the 20 % cmc case, the effect of surfactant was
surface tension of the interface (Wang and Brito-Parada, 2021). noted to be dominating over the bubble surface loading, and conse­
Huang et al. (2011) showed the suitability of partial surfactant- quently the drag force modification factor was found insensitive to the
contamination model (Sadhal and Johnson, 1983) in predicting the bubble surface loading parameter.
drag coefficient of a particle-laden bubble in pure water. Other re­ It however remains unclear how the combined effect of surfactant
searchers adopted the similarity of particle-laden bubble to a rigid and bubble surface loading affect the interface rigidity and therefore the
bubble and suggested empirical correction factors to the well-known drag coefficient and rise velocity (Yan et al., 2021) when the surfactant
Schiller-Naumann drag coefficient model (Schiller et al., 1935; Wang effect is not too dominating. The current work, therefore, examines the
et al., 2019a,b; Wang et al., 2021) which has been shown to linearly intermediate surfactant concentrations range between 0 and 25 % cmc
correlate with the bubble surface loading level in pure water (Wang of SDS for the combined effect of surfactant and bubble surface loading.
et al., 2021). This is expected to aid to the modelling studies of particle-laden bubble
Adding surfactant in pure water is a common way to generate small dynamics for better prediction of flotation kinetics.
bubbles by reducing surface tension and preventing bubble coalescence With this overarching aim, the specific objectives of this study were
(Clift et al., 1978; Nguyen and Schulze, 2004; Mitra et al., 2021). In the set to i) quantify rise velocity for different initial bubble surface loading
presence of surfactant, bubble surface becomes partially, or fully levels and surfactant concentrations; ii) propose a regime map for the

2
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 2. Schematic diagram of the experimental setup for determining bubble rising behaviour.

dominant mechanisms behind the bubble surface rigidity; iii) propose motor speed (see Fig. 2). The chosen particle size was typical to the
both drag coefficient and rise velocity correction factor in the identified coarse end of the particle size distribution commonly encountered in the
regimes; and finally iv) quantify the transient rise velocity for particle- conventional mineral flotation process which was convenient for
laden bubbles using a force balance based mathematical model incor­ handling and visualisation. The glass particles were rendered a contact
porating the proposed drag correction factor sub-model. angle of ~ 53◦ after hydrophobicisation using 2 % hydrophobic solution
prepared by mixing a commercial hydrophobic agent NG1010 in RO
2. Experimental water (Wang et al., 2021).
Once the required bubble surface coverage level was attained, the
2.1. Materials screw feeder was stopped, and air injection was resumed by turning on
the syringe pump to dislodge the particle-laden bubble from the nozzle
Six different solutions in aqueous medium were prepared using So­ tip. Details of the bubble surface loading level from the captured images
dium Dodecyl Sulphate (SDS) surfactant with concentration ranging are presented in section 2.3.
from 0 to 25 % cmc where 100 % cmc of SDS equals to 8.2 × 10-3 mol/L High-speed imaging was performed to capture the rising dynamics of
(Markarian et al., 2005) to modify bubble surface mobility. Surface a particle-laden bubble using a Phantom v311 camera in shadowgraphy
tension parameter of these solutions was measured by a Sigma Force mode at 1000 frames per second and 65 µs exposure time. Proper
tensiometer 700 (Biolin Scientific) which varied in the range from 70.9 backlight illumination was provided by placing a diffuser screen in front
mN/m at 0 % cmc to 47.6 mN/m at 25 % cmc (592 ppm). The surface of a bright halogen light source (12 V, 50 W) placed at a sufficient dis­
tension as a function of surfactant concentration cs was fitted for inter­ tance from the apparatus to avoid any heating yet ensuring high contrast
polation purpose for any intermediate concentration as follows (see images. The field of view (FoV) in the images was kept at 512 × 800
Fig. 1): pixels with a pixel resolution of ~ 29 µm/pixel.

σ = 5 × 10− 5 c2s − 0.0639cs + 68.619 (1)


2.3. Data processing
where cs is in the unit of ppm.
The captured images were processed using an in-house MATLAB
2.2. Experimental setup and procedure code to quantify some important parameters such as bubble diameter,
shape (aspect ratio), surface loading, and rise velocity. The image pro­
A single bubble was generated using a precision digital syringe pump cessing method involved marking the boundary of a particle-laden
(Adelab Scientific) and injected through the tip of a nozzle into a sur­ bubble based on both pixel intensity contrast between the bubble
factant solution kept in a transparent acrylic tank (7 cm length × 7 cm (foreground object) and the background and size difference between the
width × 9.5 cm height) as shown in Fig. 2. The air injection process was bubble and the detached particles (see Fig. 3).
set at a flow rate of 0.1 ml/min and stopped once the height of the The clean bubble surface was then identified (top blue boundary
pendant bubble reached a desired level. above the two red dots in Fig. 3) by isolating the particle covered
Glass ballotini particles (Potter Mix, Australia) of density ρp ~ 2500 interface based on the bubble interface-particle coverage intersection
kg/m3 with a narrow size distribution (d32 = 114 μm) were fed above the points represented by the two red dots and subsequently fitted by an
stationary bubble attached to the nozzle in a controlled manner using an ellipse. The best fitted semi-minor diameter a, and semi-major diameter
in-house developed miniature screw feeder system by controlling its b were used together with the bubble centre to reconstruct the whole

3
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 3. Identification of bubble surface loading from a raw image of a rising particle laden bubble (Wang et al., 2021).

ellipsoidal bubble. Details of the image processing algorithm for surface The bubble rise velocity UBP was determined by dividing the shift in the
loading identification and ellipsoidal surface reconstruction procedure bubble centroids in the two successive images by the corresponding time
can be found in Wang et al. (2021). difference as follows:
Subsequently, the equivalent bubble diameter from the raw images
was calculated as: yB,i+1 − yB,i
UBP = (4)
Δt
( )1/3
dB = 2 ab2 (2)
Determining change in the bubble centroids calculated using Eq. (4)
Following bubble detachment from the nozzle, bubble size measure­ involved around two pixels error. For a pixel resolution of ~ 29 µm/
ments were performed once the initial surface oscillations subsided. The pixel, and time resolution Δt of 5 × 10-3 s, the measurement error in the
average diameter of the bubbles generated in various surfactant solu­ estimated bubble rise velocity UBP was ~ 0.0116 m/s.
tions was dB = 2.88 ± 0.07 mm. It is important to mention that this Assuming a monolayer of particles at bubble interface, bubble sur­
bubble size was found convenient to generate in a consistent manner. face loading (BSL) parameter was determined as a ratio of the surface
Bubbles of this size were quite stable at the nozzle tip allowing sufficient area of the particles covered bottom ellipsoidal cap (marked by the
time for particles to settle and coat a layer of particles on the bubble dashed magenta line) Scap to that of the whole surface area of the bubble,
surface. Generating any smaller bubble size using a smaller size nozzle Sbubble:
was very difficult to stabilise at the nozzle tip for the same duration in Scap
BSL = (5)
the presence of surfactant. Utilising the major and minor diameter ob­ Sbubble
tained from the images, the bubble aspect ratio AR was calculated as
a wherein the surface area of the bubble Sbubble, was expressed as:
AR = (3)
b

4
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 4. Temporal evolution of rise velocity for bare bubbles in various SDS surfactant concentration.

( )1/1.6 medium, hydrophobic end of the surfactant molecules adsorb on the


2(ab)1.6 + (bb)1.6
Sbubble = 4π (6) bubble surface. As the bubble rises, surfactant molecules are swept off
3
the front part of the bubble and accumulate in the rear part leading to a
surface tension gradient along the interface. Consequently, an interfacial
3. Results and discussion shear stress is generated which opposes the internal circulations of the
bubble leading to increasing drag force and consequently reduces its rise
3.1. Rising dynamics of bare bubbles velocity.
It can be also noted that terminal rise velocity also decreased with
The rising behaviour of the bare bubbles was studied first as the increasing surfactant concentration. For example, in the steady state
baseline case. A numerical model based on force balance approach was velocity region (t > 40 ms), bubble terminal velocity decreased from ~
also developed to predict the bubble rise velocity which was validated 0.29 m/s in a no-surfactant case (Cs = 0 % cmc) to ~ 0.23 m/s at Cs = 15
by the experimental data. Fig. 4 compares the time varying rise velocity % cmc indicating a decrease of approximately 19 %. This behaviour can
of bare bubbles both in the absence and presence of different surfactant be attributed to a slight reduction in the bubble diameter with the
concentrations (0 to 25 % cmc). It can be seen that rise velocity in all decreasing surface tension parameter as well as the increasing surface
cases first increases due to upward acting buoyancy force which is immobility (partial to no-slip condition) contributing to higher drag
gradually balanced by the downward acting drag force and finally rea­ force (Jávor et al., 2016).
ches a terminal rise velocity state indicated by the plateau region. The observed reductions both in the oscillatory nature of the rise
Bubble rise velocity trend in the absence of surfactant was quite fluc­ velocity profiles as well as velocity magnitude, were not pronounced
tuating and did not reach a steady plateau state in a strict sense. These beyond the 15 % cmc case which is evident from the almost overlapping
fluctuations are attributed to the observed shape oscillations of a rising velocity profiles noted in the 20 % cmc and 25 % cmc cases. This
bare bubble due to periodic exchange between the surface energy and behaviour can be attributed to complete coverage of the bubble surface
kinetic energy (Wang and Brito-Parada, 2021). Due to reduction in the by surfactant molecules at the 20 % cmc condition and any further in­
surface tension parameter in the presence of surfactant (Fig. 1), bubbles crease in the surfactant concentration does not affect the mobility con­
were noted to transition from the ellipsoidal state to a spheroidal state dition prevailing at the bubble interface.
with increasing surfactant concentration. This was noted to have a sig­ The experimentally determined bare bubble rise velocity obtained in
nificant effect on the corresponding bubble rise velocity profiles. Fig. 4, was first compared with Clift et al. (1978) model which is given as
With increasing surfactant concentration in the system, earlier follows:
observed fluctuations in the rise velocity profiles were noted to consis­
μL
tently decrease. A true plateau state in the rise velocity profiles was UB = 1.05 Mo− 0.149
(J − 0.857) (7)
ρL d B
attained at a relatively higher surfactant concentration (15 % cmc). This
can be explained by the Marangoni effect which is caused by a
wherein μL = liquid viscosity, ρL = liquid density and Morton number Mo
nonuniform concentration distribution of surfactant over the bubble
is expressed as:
surface. When a bubble is generated in a surfactant rich aqueous

5
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 5. Validation of the terminal velocity model (Eq. (7)) of bare bubbles with the experimental data.

gμ4L (ρL − ρB ) experimental data (deviation < 10 % and the root-mean-square devia­
Mo = (8)
ρ2L σ 3 tion (RMSD) = 0.020). Error bars for each surface tension case were also
plotted herein but those are not distinguishable because of small errors
wherein σ = gas–liquid surface tension, ρB = bubble density. involved (standard deviations ~ 1.2 × 10-4 to 6.2 × 10-4). Based on the
The variable J in Eq. (7) is a stepwise function which is written as: agreement obtained, Eq. (7) was used as the base model in Section 3.2
{ for development of a velocity correction factor to account for the pres­
0.94H 0.757 (2 < H⩽59.3)
J= (9) ence of particles on bubble surface.
3.42H 0.441 (H > 59.3)
A small jump in the experimentally measured terminal rise velocity
can be noted as the surface tension parameter increases from 0.0533 to
wherein the variable H is expressed as:
0.0566 N/m corresponding to SDS concentration decreasing from 15 %
4 0.14
(10) cmc to 10 % cmc. This behaviour could be attributed to the transition of
H = EoMo− 0.149
(μL /μw )−
3 bubble surface mobility from rigid to partially rigid state with less
adsorption of surfactant on the surface in the 10 % cmc case causing
ρ gd2
where μw = liquid viscosity at wall and Eötvös number Eo = L σ B . larger shape oscillations. The rise velocity correlation model in Eq. (7) in
Fig. 5 shows good agreement of the model prediction with the the 10 % cmc case however predicted a smoother velocity increase
without capturing this jump pattern which is attributed to inaccuracy in
the Clift model itself to capture variability in the rise velocity data for
Table 1
ellipsoidal shape bubbles of size around 3 mm due to oscillatory
Empirical models for drag coefficients of a bare bubble.
behaviour in the bubble shape.
Researchers Models To predict the temporal velocity of a bare bubble as it ascends from
its stationary state, a rise velocity model was proposed based on New­
{ [ 24 18.7 ]
Peebles and Garber CD,B = max max , ,
ReB Re0.68 ton’s second law comprising gravity, virtual mass, buoyancy and drag
(1953) [ ]} B
0.0275EoWe2 , force as follows:
min
0.82Eo0.25 We0.5
{ 24 ( [ ]} ∫ Ut ∫t
) 8 2 FB,B − FG,B − FD,B
Ishii and Chawla (1979) CD,B = max 1 + 0.1Re0.75 , min , Eo0.5
ReB B
3 3 dUB = dt (11)
{ 24 ( ) 8 Eo
} 0 0 m′B
Tomiyama et al. (1998) CD,B = max 1 + 0.15Re0.687
B ,
ReB 3 (Eo + 4)
[( )2 ]0.5 where the modified bubble mass m′B is expressed as sum of the actual and
Zhou et al. (2020) CD,B = CD,Mei + (CD (Eo) )2 CD,Mei =
virtual mass of a single bubble:
{ [ ( )] }
1
16 8 1 3.315 −
1+ + 1 + 0.5 CD (Eo) = 4 4
ReB ReB 2 ReB m′B = πR3B ρB + Cv πR3B ρL (12)
⎧ ⎫
3 3
(− 1.23log(Eo)+0.37log(Mo)+1.6 )

⎨ 10 , ⎪

max [
4Eo 8Eo
] where virtual mass coefficient was selected as Cv = 0.5.
⎩ min
⎪ ⎪
,
(Eo + 9.5) 3(0.8726Eo + 4.887)

FB,B in Eq. (11) denotes the bubble buoyancy force which is given as:

6
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 6. Comparison of theoretical and experimental rise velocity of a bare bubble in a) 0 % cmc; b) 10 % cmc; c) 15 % cmc; and d) 25 % cmc. Water densityρL = 998
kg/m3, Bubble densityρB = 1.18 kg/m3.

4 ~ 12 to 230 times the maximum Reynolds number (ReB = 863) obtained


FB,B = π R3B ρL g (13)
3 in the bare bubble cases in this study.
Considering drag coefficient is inversely related to bubble Reynolds
RB in Eq. (13) is bubble radius and ρL is the density of liquid. FG,B in Eq. number, the constant model underestimates drag coefficient and
(11) represents the bubble gravitational force: therefore overestimates bubble rise velocity. At low surfactant concen­
4 trations (below 15 % cmc) Zhou et al. model appears to be the best fit for
FG,B = π R3B ρB g (14) the experimental data with a minimum RMSD 0.016. This indicates that
3
Zhou et al. model is suitable for pure or slightly contaminated system. As
where ρB is bubble density. The drag force of a bare bubble in Eq. (11) is the concentration of SDS increased above 15 % cmc the Tomiyama et al.
expressed as: model provides best fit for the experimental data with a minimum RMSD
0.012 compared to Bo than Zhou et al. (2020) and Ishii and Chawla
FD,B = 0.5CD,B ρL πR2B UB2 (15)
(1979) model.
where the drag coefficient CD,B in surfactant contaminate solution has
been extensively modelled in literature and briefly summarized in 3.2. Particle-laden bubbles
Table 1.
The temporal bubble velocity UB in pure water as a bare bubble as­ 3.2.1. Rise regime
cends in different surfactant concentrations was predicted using Determining the bubble surface loading level (BSL) in flotation
different drag coefficient models (Table 1) and results are presented in process is important when bubbles reach a terminal rise velocity state
Fig. 6. The drag coefficient model developed by Yan et al. (2018) in due to its strong connection to final product recovery. The BSL param­
contaminated system however was not included in here due to current eter is also included in the calculation of the number of loaded particles
bubble Reynolds number being outside the applicable range of Yan et al. in the force balance model presented in this study.
(2018) which caused significant deviation in predicted bubble velocity The experimental images of rising particle-laden bubbles with
(Zhou et al., 2020). different initial BSL values for two different surfactant concentrations
In Fig. 6, Peeble and Garber (PC) model overestimated the terminal are presented in Fig. 7. Bubbles released from the nozzle initially un­
velocity by approximately twice of the experimental value. This is derwent shape transitions from spheroidal to ellipsoidal state.
because of the higher Eötvös number obtained in this study (Bo ~ 1.18 Two regimes were noted - an initial interface deformation phase after
to 1.61) which is not suitable for PC model (Tomiyama et al., 1998). The pinch-off from the nozzle and subsequent shape oscillations similar to
constant drag coefficient 0.48 can only be achieved in turbulent flow the earlier observations reported in Wang and Brito-Parada (2021).
when ReB is in the range of 1 × 104 – 2 × 105 (Clift et al., 1978) which is Wang and Brito-Parada (2021) determined the frequency and the
damping rate of surface oscillations by decomposing the bubble shape

7
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 7. Rising of particle-laden bubbles in pure water and 15% cmc case for different initial bubble surface loading (BSL) values.

into multiple spherical harmonics. Their results showed that the pres­ rear vortices of rising bubble due to low particle Stokes number (St =
ence of particles at bubble surfaces decreases the damping rate of the ρP UB d2P /(9μL dB ) ~ 0.25 to 0.36) and are randomly dispersed in the
dominant harmonic significantly while having a minor influence on the trailing vortices. These factors together contribute to the observed de­
√̅̅̅̅̅̅̅
oscillation frequency which can be scaled approximately as ρ σd3 . Since viation between the initial and final BSL values.
The temporal evolution of bubble surface loading levels reported in
L B

the oscillation frequency is not affected by the presence of particles, the


Fig. 7 was calculated using Eq. (5) and an example involving 0 % and 15
surface tension parameter, σ of the particle-laden bubbles remains the
% cmc surfactant concentration is presented in Fig. 8. It can be seen that
same as the uncoated bubbles.
both in the absence and presence of surfactant, BSL level initially re­
It can be noted that initial BSL values prior to detachment (t = 0)
duces as the surface area of bubble as well as particle cap area at bubble
were somewhat different than the average BSL values obtained over the
bottom also decreases due to change in the bubble shape until ~ 15 ms
rise duration e.g., 0 % cmc: initial BSL: 0.24, average BSL: 0.23; initial
(Fig. 7). Several particles were observed to detach from a bubble during
BSL: 0.29, average BSL: 0.27; initial BSL: 0.76, average BSL: 0.49; 15 %
this period which also contributed to reduction in BSL.
cmc: initial BSL: 0.2, average BSL: 0.18; initial BSL: 0.27, average BSL:
However, as a bubble further rose, particle detachment ceases and
0.25; initial BSL: 0.58, average BSL: 0.40. The observed deviations in the
both the bubble surface area and the particle cap area increase leading to
initial and average BSL values varied in the range from ~ 4 to 36 %.
an increase in the BSL value. This behaviour however is different for the
Visually, in the absence of surfactant, aspect ratio (AR) of particle-
pure water (0 % cmc) and 15 % cmc case, respectively. In absence of any
laden bubbles increases with increasing BSL as the bubbles transition
surfactant, both bubble surface area and the particle coated area in­
from an ellipsoidal to spheroidal shape depending on the surface tension
crease leading to increase in the BSL value (Fig. 8a). However, in the
and particle loading. Also, detachment of particles from bubble surface
presence of surfactant (15 % cmc case), surface area of bubbles remains
occurs in this transition period as well as re-organisation of particles at
almost constant (Fig. 8b) due to the acquired surface rigidity which
the rear end of the bubble due to interactions with the counter rotating
leads to relatively less steep rate in the BSL increase compared to the
trailing vortices. It can be seen the detached particles are trapped in the

8
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 8. Evolution of BSL over time in a) 0% cmc case (initial BSL: 0.29, average BSL: 0.27) and b)15% cmc case (initial BSL: 0.27, average BSL: 0.25).

pure water case. Such increase in the BSL parameter is ascribed to the The summary of bubble terminal velocity obtained at various BSLs
interaction of the particle layer with the trailing vortices which stretches and surfactant concentrations in Fig. 10 confirms this observation. The
a bubble in the equatorial direction (Wang et al., 2021). decreasing terminal velocity profiles with BSL as an independent
The lower increasing rate of BSL value in the presence of surfactant parameter can be categorized into three groups according to surfactant
can be attributed to Marangoni effect. When bubbles rise, the absorbed concentration: Cs = 0 % cmc, 5 % to 10 % cmc, and 15 % cmc and above,
surfactant on bubble surface is swept down to the rear part increasing respectively. In the first category (Cs = 0 % cmc) wherein the bubble
surfactant concentration. The resultant concentration gradient gener­ surface is mobile, Fig. 10 shows a non-linear decreasing trend of ter­
ates Marangoni stress in the tangential direction which tends to retard minal velocity with the BSL parameter. This decreasing trend is attrib­
surface motions. Consequently, particles at the interface slow down uted to higher drag force resulting from the decreasing buoyancy or due
when they are pushed towards the bubble equator by the trailing to presence of particles on bubble surface. In the second category (Cs = 5
vortices (Wang et al., 2021) leading to relatively lower expansion of the % to 10 % cmc), such decreasing trend of the rise velocity profiles be­
instantaneous BSL boundary. comes more linear wherein slope (UBP/BSL) is directly affected by the
In general, it is noted that BSL value increases for the initial BSL level BSL parameter. Increasing bubble surface loading levels decreases
< 0.4 and gradually becomes steady for the initial BSL > 0.4 (not bubble rise velocity due to increased gravity and drag forces. It can be
shown). The threshold value BSL ~ 0.40 obtained in this study is in noted that the lowest rise velocity occurred in the 10 % cmc case which
agreement with our previous study (Wang et al., 2021) involving pure had the largest BSL ~ 0.55 among all other cases.
water (no surfactant case) which confirms that such threshold is inde­ Although the degree of contamination on bubble surface by surfac­
pendent of the surfactant concentration. tant is difficult to measure (Clift et al., 1978; Legendre et al., 2009), it
The time evolution of aspect ratio (AR) and rise velocity of particle- can be postulated from these trends that bubble surface is partially rigid
laden bubbles UBP quantified in three different surfactant concentrations (partial slip condition) in the presence of surfactant, and the surface
are shown in Fig. 9. In the absence of surfactant (Fig. 9a), The aspect rigidity increases with more particles loaded onto bubble surface leading
ratio of the particle-laden bubble at a high BSL value ~ 0.42 remains to increased drag force. The third category involves a lower slope of rise
steady at ~ 0.75, significantly higher than the aspect ratio value ~ 0.35 velocity at higher surfactant concentration which indicates bubble sur­
at BSL = 0.00 and 0.23. In all cases, bubble rise velocity is observed to be face becomes more rigid from adsorption of larger amount of surfactant
inversely correlated to its aspect ratio and bubble surface loading. High (no-slip condition) and consequently bubble velocity becomes less
BSL values also suppress oscillations in the bubble rise velocity profiles. dependent on the BSL parameter.
As the surfactant concentration increases (Fig. 9b-c), the effect of BSL Given the importance of bubble surface rigidity in affecting the
parameter on the steady state aspect ratio (AR) becomes somewhat bubble rise velocity in the three surfactant concentration categories
insignificant, fluctuating around a steady value of 0.80 which indicates described in Figs. 9 to 10, a regime map of bubble surface rigidity in the
bubble surface elasticity achieved by the adsorption of surfactant, is not presence of surfactant and loaded particles is proposed in Fig. 11. In
amenable to any further shape change by bubble surface loading. It can Regime I at both BSL = 0 and Cs = 0 % cmc, bubble surface is free of
be also observed that once BSL reaches beyond ~ 0.24, with increasing surfactant and particles and therefore is mobile. In Regime II at BSL <
surfactant concentration Cs, the difference between bubble terminal 0.5 and 0 < Cs < 15 % cmc, bubble becomes partially rigid under the
velocity at different BSLs in the steady region reduces (Fig. 9b) and combined effect of loaded particles and surfactant. In Regime III bubble
becomes almost negligible at 15 % cmc (Fig. 9c). surface can become rigid from the contributions of loaded particles at

9
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 9. Rise velocity and instantaneous aspect ratio of particle-laden bubble in a) pure water; b) 10% cmc and c) 15% cmc case.

10
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 10. Terminal velocity of particle-laden bubbles vs bubble surface loading for different surfactant concentrations.

Fig. 11. Domains of dominant mechanisms for bubble surface rigidity.

11
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 12. Velocity modification factor as a function of bubble surface loading in various surfactant concentrations: a) 0% cmc; b) 5% − 10% cmc; c) 15% − 25% cmc.

BSL ≥ 0.5 and 0 < Cs < 15 % cmc, or from the adsorbed surfactant at 3.2.2. Bubble rise velocity model
BSL < 0.5 or combinedly from contribution of both particles and A velocity correction factor αBP,U was introduced to modify the rise
adsorbed surfactant at BSL ≥ 0.5 and Cs ≥ 15 % cmc. velocity of a bare bubble of same diameter dB in the presence of a sur­
Although surfactant plays an important role in controlling bubble factant of concentration Cs in the aqueous medium. Applying Eq. (7), the
rise velocity (Davis and Acrivos, 1966; Griffith, 1962; Takagi and Mat­ rise velocity of a particle-laden bubble was written as:
sumoto, 2010) as shown in Figs. 9 to 11, developing an analytical model μL
from first principles to determine the bubble rise velocity including UBP = αBP,U 1.05 Mo− 0.149
(J − 0.857) (16)
ρL dB
surfactant effect is a challenging problem. This is due to the combined
complexity posed by the distribution of surfactant on bubble surface and where the Morton number Mo in Eq. (7) is correlated to ρBP the density
the corresponding slip condition, deforming gas–liquid interface and of bubble-particle aggregate, namely the air bubble with the attached
bubble Reynolds number which usually falls in the intermediate flow particles. Mo was calculated as follows:
regime rather than the ideal flows defined by the Stokes or potential
flow field (Wang et al., 2021) for which no straightforward solutions Mo =
gμ4L (ρL − ρBP )
(17)
exist. ρ2L σ3
To the best of the authors’ knowledge, limited studies are reported
on the analytical modelling of rise velocity of a particle-laden bubble in ρBP was calculated by dividing the volume of single bubble into the sum
the presence of a surfactant. The model proposed by Yan et al. (2021) is of the bubble mass and loaded particle mass, as shown in the following
only suitable for pure water condition due to the assumption that the equation:
loaded particles have the same effect on the bubble surface mobility as a ρB dB3 + NP ρP dP3
surfactant. Nevertheless, this study shows that in presence of surfactant, ρBP = (18)
dB3
bubble surface becomes partially or even fully rigid even without any
particles however both parameters affect the bubble rise velocity when As the particles form a monolayer on the surface of the bubble, the
surfactant concentration is below a threshold value. In this context, the change in total volume of the bubble has very little difference between
models of rise velocity and drag coefficient of a single bubble accounting the bare bubble case and loaded bubble > 1 %. The volume of particles
for both particles and surfactant concentrations are presented in the therefore was ignored when determining the volume of the particle-
following sub-sections. laden bubble.
The number of loaded particles, NP in Eq.(18) was determined ac­
cording to Wang et al., (2020) as:

12
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 13. Drag correction factor as a function of bubble surface loading for various surfactant concentrations: a) 0% cmc; b) 5 to 10% cmc; c) 15 to 25% cmc.

4πdB2 (ϕBSL)
NP = (19) 3.2.3. Drag coefficient model
πdP2
The drag correction factor αBP,U which represents the effect of bubble
surface loading on
where ϕ is the packing factor of the particle loading region and was
assigned a value of 0.8 (Gallegos-Acevedo et al., 2006; Huang et al., 24 ( )
CD,BP = αBP,D 1 + 0.15Re0.687 (21)
2011). ReBP BP

The correction factor αBP,U in Eq. (16) was determined by dividing


the experimentally measured rise velocity of a bare bubble by the cor­ where the bubble Reynolds number ReBP in aqueous medium calculated
responding particle-laden bubble rise velocity. The model of αBP,U was from the terminal rise velocity reported in Fig. 9 and bubble diameter
then developed by fitting it to the measured BSL data. determined using Eq. (2), takes the following form:
αBP,U values are presented in Fig. 12 for the three different bubble ρL dB UBP
surface regimes – mobile surface, partially immobile surface and fully ReBP = (22)
μL
immobile surface. The parameter αBP,U exhibits distinctive behaviour in
these three regimes due to the competition between the adsorbed sur­ The force balance model depicting the terminal rise velocity state (no
factant and loaded particles. In the pure water condition (Fig. 11a) net force) for a particle-laden bubble was written by balancing the drag
which features the mobile bubble surface, αBP,U exhibits a non-linear force with buoyancy force with gravity and drag force as follows:
dependency on the BSL parameter. The dependency of αBP,U on BSL
clearly decreases with increasing surfactant concentration. In the sur­ FB,B + FB,P − FG,B − FG,P − FD,BP = 0 (23)
factant concentration range Cs = 5 to 10 % cmc (Fig. 11b), αBP,U linearly
where the drag force of the particle-laden bubble was determined as:
decreases with BSL with a slope ~ 0.52, however once the surfactant
concentration exceeds 15 % cmc (Fig. 11b), it becomes insensitive to FD,BP = 0.5CD,BP ρL πR2B UBP
2
(24)
bubble surface loading.
The velocity correction factor in these three different regimes The buoyancy force FB,B and gravitational force FG,B for a single bubble
describing the effect of bubble surface loading and surfactant concen­ were recalculated using Eqs. (13) and (14), respectively while the
tration can be best fitted by the following equations: buoyancy and gravity force for the particle layer were computed as
⎧ follow:
⎨ 1.00 + 0.33 BSL − 1.64 BSL2 (Cs = 0%, in pure water)
αBP,U = 1.00 − 0.42 BSL (0% < Cs < 15% cmc) (20) 4

0.93 (Cs ⩾15% cmc) FB,P = πR3P ρL gNP (25)
3

13
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Fig. 14. Comparison of predicted and experimental rise velocity of particle-laden bubbles in a) 0 % cmc; b) 10 % cmc; c) 15 % cmc; and d) 25 % cmc of SDS solution.
Liquid densityρL = 998 kg/m3, bubble densityρB = 1.18 kg/m3.

4 however suggests that Model Eq. (27) predicts the correction factor
FG,P = π R3P ρP gNP (26) reasonably well even for bubble diameter as small as ~ 1 mm obtained
3
from Huang et al. (2011). Although no other experimental data is
where the number of loaded particles, NP, was calculated using Eq. (19). available to verify other bubble diameters, the reasonable agreement of
The force balance model presented in Eq.(23) was used to determine the proposed correlation down to 1 mm size bubble indicates that the
the drag coefficient for particle laden bubble (CD,BP) using the experi­ models developed in this study is also applicable within the range 1 ≤dB
mentally measured rise velocity (UBP). Then, the drag correction factor ≤ 2.88 mm and aspect ratio AR ≥ 0.5. It should be noted that the data for
(αBP,D ) was determined by comparing CD,BP with the Schiller-Naumann bare bubble in pure water was not included in Fig. 13 as its aspect ratio
drag model (Schiller et al., 1935) in Eq. (21) by replacing the bare was lower (~0.3) than the threshold selected for Eq. (27). This proposed
bubble rise velocity (UB) with the particle laden bubble rise velocity model of drag coefficient factor presented in Eq. (27) covers a reason­
(UBP). The model of the empirical drag correction factor αBP,D was then able range of surfactant concentration and serves as an extension to the
developed by fitting it to the measured BSL data. It needs to be noted drag correction model earlier reported in Wang et al., (2021).
that use of experimentally measured bubble rise velocity in determining The rise velocity model for the bare bubble case presented in Eq. (11)
the velocity and drag correction factor decouples the interdependency of was suitably modified to account for the presence of particles by
these parameters which are otherwise related to each other through the including the drag coefficient correction factor αBP,D as follows:
force balance model. ∫ Ut ∫t
FB,B + FB,P − FG,B − FG,P − FD,BP
The drag modification factor αBP,D is presented in Fig. 13 for mobile- dUBP = dt (28)
surface bubble (Cs = 0 % cmc Fig. 13a), partial rigid bubble (Cs = 5 % 0 0 m′BP
and 10 % cmc, Fig. 13b) and rigid bubble (Cs ≥15 % cmc, Fig. 13c). where the modified mass m′BP is equal to the addition of the mass of
Similar to the velocity correction factor, the drag correction factor αBP,D loaded particles to the modified mass of a bare bubble m′B calculated
also shows distinctive behaviours in the three different bubble surface using Eq. (12) in Section 3.1 as below:
regime for near-spherical bubbles (BSL≤0.50 and AR ≥ 0.5) which can
be best fitted by the following expressions: 4 4 4
m′BP = πR3B ρB + πR3P ρP NP + Cv πR3B ρL (29)
⎧ 3 3 3
⎨ 0.29 + 2.56BSL (Cs = 0%, in pure water)
αBP,D = 1.11 + 0.81 BSL (0% < Cs < 15% cmc) (27) The rise velocity of a particle-laden bubble UBP predicted using Eq. (28)

1.44 (Cs ⩾15% cmc) for various surfactant concentrationCs = 0 to 25 % cmc is presented in
Fig. 14. Overall, a reasonable prediction was obtained from the model
It can be noted that in this study, only a single nozzle size was used to
when compared with all experimental cases (average RMSD ~ 0.028). In
produce bubbles of averaged diameter 2.88 ± 0.07 mm. Fig. 13a
the transition period, relatively larger deviations were observed in the

14
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

surfactant concentration range from 0 % to 15 % cmc. The observed Data availability


deviation is attributed to the complex bubble shape transitioning from
pendant state to ellipsoidal state which is different from the spherical Data will be made available on request.
shape assumed in the model. As surfactant concentration was further
increased to 25 % cmc, the deviation significantly reduced as immobility Acknowledgements
in the bubble surface increased due to the adsorption of surfactant and
the surface became more rigid (no slip condition) leading the bubble The authors would like to acknowledge the funding support from the
becoming more spherical. In the steady region where the terminal rise Australian Research Council for the ARC Centre of Excellence for
velocity was reached, the model predicted velocity agrees quite well Enabling Eco-Efficient Beneficiation of Minerals, grant number
with the experimental data for the examined range of surfactant CE200100009.
concentration.
References
4. Conclusion
Bennett, G.F., Peters, R.W., 1988. The removal of oil from wastewater by air flotation: A
In this study, rising dynamics of particle-laden bubbles were exam­ review. Crit. Rev. Environ. Control 18 (3), 189–253. https://doi.org/10.1080/
ined in solutions of different SDS concentrations (0 to 25 % cmc) for 10643388809388348.
Bradshaw, D.J., Connor, C.T., 1996. Measurement of the sub-process of bubble loading in
various levels of bubble surface loading (BSL). It was found that both flotation. Miner. Eng. 9 (4), 443–448. https://doi.org/10.1016/0892-6875(96)
surfactant and particles contribute to modification of bubble surface 00029-5.
mobility and affect the terminal rise velocity. The key findings are Chang, L., Cao, Y., Fan, G., Li, C., Peng, W., 2019. A review of the applications of ion
floatation: wastewater treatment, mineral beneficiation and hydrometallurgy. RSC
summarised below: Adv. 9 (35), 20226–20239. https://doi.org/10.1039/c9ra02905b.
Chen, Y.M., Liu, J.C., Ju, Y.-H., 1998. Flotation removal of algae from water. Colloids
• For the bare bubble (BSL = 0) case, terminal rise velocity decreased Surf. B Biointerfaces 12, 49–55. https://doi.org/10.1016/S0927-7765(98)00059-9.
Clift, R., Grace, J.R., Weber, M.E., 1978. Bubbles, drops, and particles. Academic Press,
with increasing surfactant concentration due to adsorption of sur­
New York.
factant on the bubble surface resulting in increased surface rigidity Davis, R.E., Acrivos, A., 1966. The influence of surfactants on the creeping motion of
and increased drag force. bubbles. Chem. Eng. Sci. 21 (8), 681–685. https://doi.org/10.1016/0009-2509(66)
80017-9.
• In the presence of particles, bubble terminal velocity generally
Eskanlou, A., Khalesi, M.R., Abdollahy, M., Chegeni, M.H., 2018b. Interactional effects of
reduced as the BSL parameter increased due to higher drag force bubble size, particle size, and collector dosage on bubble loading in column
resulting from the layer of particles at the interface. The dependency flotation. Journal of Mining & Environment, 9(1), 107-116. 10.22044/
of terminal velocity on BSL however decreased as the surfactant jme.2017.5815.1393.
Eskanlou, A., Khalesi, M.R., Mirmogaddam, M., Hemmati Chegeni, M., Vaziri Hassas, B.,
concentration was increased. This occurred because a bubble surface 2018. Investigation of trajectory and rise velocity of loaded and bare single bubbles
attains more rigidity by adsorbing surfactant in a high concentration in flotation process using video processing technique. Sep. Purif. Technol. 54 (11),
environment compared to the presence of particles at the interface. 1795–1802. https://doi.org/10.1080/01496395.2018.1539104.
Gallegos-Acevedo, P.M., Pérez-Garibay, R., Uribe-Salas, A., 2006. Maximum bubble
• A regime map showing three types of bubble surface rigidity – mobile loads: Experimental measurement vs. analytical estimation. Miner. Eng. 19 (1),
(deformable), partially rigid (partially immobile) and fully rigid 12–18. https://doi.org/10.1016/j.mineng.2005.04.002.
(immobile) in the presence of surfactant and loaded particles was Griffith, R.M.A., 1962. The effect of surfactants on the terminal velocity of drops and
bubbles. Chem. Eng. Sci. 17 (12), 1057–1070. https://doi.org/10.1016/0009-2509
presented with appropriate correction factors applied to both rise (62)80084-0.
velocity and drag coefficient. Hessenkemper, H., Ziegenhein, T., Lucas, D., 2020. Contamination effects on the lift force
• The correction factors showed higher dependency on BSL in the of ellipsoidal air bubbles rising in saline water solutions. Chem. Eng. J. 386 https://
doi.org/10.1016/j.cej.2019.04.169.
lower surfactant concentration regime which gradually diminished
Huang, Z., Legendre, D., Guiraud, P., 2011. A new experimental method for determining
as the surfactant concentration was increased. The force balance particle capture efficiency in flotation. Chem. Eng. Sci. 66 (5), 982–997. https://doi.
model for particle laden bubble rise velocity accounting for these org/10.1016/j.ces.2010.12.006.
Ireland, P.M., Jameson, G.J., 2014. Collision of a rising bubble–particle aggregate with a
correction factors produced quite reasonable agreement with
gas–liquid interface. Int. J. Miner. Process. 130, 1–7. https://doi.org/10.1016/j.
experiment especially in the higher surfactant concentration cases minpro.2014.05.002.
where bubble shape was nearly spherical. Ishii, M., Chawla, T. C., 1979. Local drag laws in dispersed two-phase flow. https://www.
osti.gov/biblio/5512664.
Jameson, G.J., Emer, C., 2019. Coarse chalcopyrite recovery in a universal froth flotation
Future work in this area will address the effect of gas volume fraction machine. Miner. Eng. 134, 118–133. https://doi.org/10.1016/j.
on the drag coefficient and rise velocity by studying a swarm of particle mineng.2019.01.024.
laden bubbles of varying number density. Jávor, Z., Schreithofer, N., Heiskanen, K., 2016. Multi-scale analysis of the effect of
surfactants on bubble properties. Minerals Eng. 99, 170–178. https://doi.org/
10.1016/j.mineng.2016.09.026.
CRediT authorship contribution statement Koh, P.T.L., Schwarz, M.P., 2008. Modelling attachment rates of multi-sized bubbles with
particles in a flotation cell. Miner. Eng. 21 (12–14), 989–993. https://doi.org/
10.1016/j.mineng.2008.02.021.
Ai Wang: Writing – original draft, Data curation, Formal analysis, Lee, K., Archibald, D., McLean, J., Reuter, M.A., 2009. Flotation of mixed copper oxide
Investigation, Methodology, Software, Validation. Edwin Banks: Data and sulphide minerals with xanthate and hydroxamate collectors. Miner. Eng. 22 (4),
curation, Formal analysis, Investigation, Methodology, Software, Vali­ 395–401. https://doi.org/10.1016/j.mineng.2008.11.005.
Legendre, D., Sarrot, V., Guiraud, P., 2009. On the particle inertia-free collision with a
dation. Geoffrey Evans: Writing – review & editing, Funding acquisi­
partially contaminated spherical bubble. Int. J. Multiph. Flow 35 (2), 163–170.
tion, Project administration, Resources, Supervision. Subhasish Mitra: https://doi.org/10.1016/j.ijmultiphaseflow.2008.10.002.
Conceptualization, Investigation, Project administration, Resources, Markarian, S.A., Harutyunyan, L.R., Harutyunyan, R.S., 2005. The Properties of Mixtures
of Sodium Dodecylsulfate and Diethylsulfoxide in Water. J. Solution Chem. 34 (3),
Supervision, Writing – review & editing.
361–368.
Mitra, S., Hoque, M.M., Evans, G., Nguyen, A.V., 2021. Direct visualisation of bubble-
Declaration of competing interest particle interactions in presence of cavitation bubbles in an ultrasonic flotation cell.
Miner. Eng. 174, 107258 https://doi.org/10.1016/j.mineng.2021.107258.
Nanda, S., Kumar, S., Mandre, N.R., 2022. Flotation behavior of a complex lead-zinc ore
The authors declare that they have no known competing financial using individual collectors and its blends for lead sulfide. J. Dispers. Sci. Technol.
interests or personal relationships that could have appeared to influence 1–8, 2036185. https://doi.org/10.1080/01932691.2022.2036185.
the work reported in this paper. Nguyen, V.A., Schulze, H.J., 2004. Colloidal science of flotation. Marcel Dekker, New
York.
Peebles, F.N., 1953. Studies on the motion of gas bubbles in liquid. Chem. Eng. Prog. 49
(2), 88–97.

15
A. Wang et al. Chemical Engineering Science 288 (2024) 119812

Sadhal, S.S., Johnson, R.E., 1983. Stokes flow past bubbles and drops partially coated Wang, P., Cilliers, J.J., Neethling, S.J., Brito-Parada, P.R., 2019a. The behavior of rising
with thin films: Part 1. Stagnant cap of surfactant film – exact solution. J. Fluid bubbles covered by particles. Chem. Eng. J. 365, 111–120. https://doi.org/10.1016/
Mech. 126, 237–250. https://doi.org/10.1017/S0022112083000130. j.cej.2019.02.005.
Schiller, L., Naumann, Z., 1935. In: A Drag Coefficient Correlation, pp. 77–318. Wang, P., Cilliers, J.J., Neethling, S.J., Brito-Parada, P.R., 2019b. Effect of Particle Size
Schulze, H.J., 1989. Hydrodynamics of bubble-mineral particle collisions. Miner. on the Rising Behavior of Particle-Laden Bubbles. Langmuir 35 (10), 3680–3687.
Process. Extr. Metall. Rev. 5, 43–76. https://doi.org/10.1080/08827508908952644. https://doi.org/10.1021/acs.langmuir.8b04112.
Takagi, S., Matsumoto, Y., 2010. Surfactant effects on bubble motion and bubbly flows. Wang, A., Hoque, M.M., Moreno-Atanasio, R., Evans, G., Mitra, S., 2020. Development of
Annu. Rev. Fluid Mech. 43, 615–636. https://doi.org/10.1146/annurev-fluid- a flotation recovery model with CFD predicted collision efficiency. Miner. Eng. 159,
122109-160756. 106615 https://doi.org/10.1016/j.mineng.2020.106615.
Tao, D., 2004. Role of Bubble Size in Flotation of Coarse and Fine Particles—A Review. Wang, A., Hoque, M.M., Moreno-Atanasio, R., Doroodchi, E., Evans, G., Mitra, S., 2021.
Seperation Sci. Technol 39 (4), 741–760. https://doi.org/10.1081/SS-120028444. Effect of bubble surface loading on bubble rise velocity. Miner. Eng. 174, 107252
Tomiyama, A., Kataoka, I., Zun, I., Sakaguchi, T., 1998. Drag coefficients of single https://doi.org/10.1016/j.mineng.2021.107252.
bubbles under normal and micro gravity conditions. JSME Int. J. Series B 41 (2), Wang, A., Evans, G., Mitra, S., 2023. A review of bubble surface loading and its effect on
472–479. https://doi.org/10.1299/jsmeb.41.472. bubble dynamics. Miner. Eng. 199, 108105 https://doi.org/10.1016/j.
Tripathi, M.K., Sahu, K.S., Govindarajan, R., 2015. Dynamics of an initially spherical mineng.2023.108105.
bubble rising in quiescent liquid. Nat. Commun. 6 (6268), 1–9. https://doi.org/ Yan, X.K., Zheng, K.X., Jia, Y., Miao, Z.Y., Wang, L.J., Cao, Y.J., Liu, J.T., 2018. Drag
10.1038/ncomms7268. Coefficient Prediction of a Single Bubble Rising in Liquids. Ind. Eng. Chem. Res. 57
Uribe-Salas, A., de Lira-Gómez, P., Pérez-Garibay, R., Nava-Alonso, F., Magallanes- (15), 5385–5393. https://doi.org/10.1021/acs.iecr.7b04743.
Hernández, L., Lara-Valenzuela, C., 2003. Overloading of gas bubbles in column Yan, X., Zheng, K., Su, W., Wang, L., Zhang, H., Cao, Y., Guo, C., 2021. Predictions of
flotation of coarse particles and effect upon recovery. Int. J. Miner. Process. 71 (1–4), terminal rising velocity, shape and drag coefficient for particle-laden bubbles. Miner.
167–178. https://doi.org/10.1016/S0301-7516(03)00036-X. Eng. 173, 107188 https://doi.org/10.1016/j.mineng.2021.107188.
Wang, H., Brito-Parada, P.R., 2021. Shape deformation and oscillation of particle-laden Zhou, Y., Zhao, C., Bo, H., 2020. Analyses and modified models for bubble shape and
bubbles after pinch-off from a nozzle. Chem. Eng. J. 412, 127499 https://doi.org/ drag coefficient covering a wide range of working conditions. Int. J. Multiph. Flow
10.1016/j.cej.2020.127499. 127, 103265. https://doi.org/10.1016/j.ijmultiphaseflow.2020.103265.

16

You might also like