You are on page 1of 9

Int. J. Miner. Process.

79 (2006) 18 – 26
www.elsevier.com/locate/ijminpro

Effect of the bubble size on the dynamic adsorption of frothers


and collectors in flotationB
Anh V. Nguyen *, Chi M. Phan, Geoffrey M. Evans
Discipline of Chemical Engineering, School of Engineering, The University of Newcastle, University Drive,
Callaghan, New South Wales 2308, Australia
Received 9 September 2005; received in revised form 22 November 2005; accepted 22 November 2005
Available online 28 December 2005

Abstract

This paper presents an analysis on the dynamic adsorption of frothers and collectors onto air bubbles in flotation. The analysis
was carried out using the theory on the surfactant transfer by diffusion and adsorption. The bubble size was considered when
solving the diffusion equation for the dynamic surface excess. The adsorption kinetics was modelled employing the Langmuir
adsorption isotherm and its extension with the inclusion of the lateral interaction of the adsorbed molecules (the Frumkin
adsorption isotherm). The numerical computation was applied to solve the governing equations. The dynamic adsorption of the
surfactants was measured in terms of the dynamic surface tension for sodium dodecylbenzene sulphonate (SDBS) and Dowfroth
250, using pendant bubble tensiometry. The results show that, for air bubbles with diameter of the order of 1 mm, the bubble size
effect on the dynamic surface tension is small and the solution of the diffusion equation for a planar gas–liquid surface can be used
to describe the dynamic adsorption onto the air bubbles. The experimental data obtained with SDBS were well described by the
Frumkin adsorption isotherm and the molecular diffusion. The model description for the experimental data obtained with Dowfroth
250 required the kinetic equations. The paper is relevant for the analysis of the role of the adsorbed surfactants in the bubble–
particle attachment.
D 2005 Elsevier B.V. All rights reserved.

Keywords: flotation frothers; flotation collectors; flotation bubbles; flotation reagents; froth flotation; dynamic adsorption; dynamic surface tension

1. Introduction between particles and bubbles in the collection pro-


cess. For this reason, it is important to understand
Numerous inorganic and organic reagents (surface how the surfactants adsorb onto the interfaces.
active agents or surfactants) are employed in flotation When surfactants react with water, the water
to improve the selectivity and recovery of valuable dipoles combine readily with the polar groups and
minerals. The presence of these surfactants influences hydrate them, but there is practically no reaction
the properties of both the air–water and solid–water with the non-polar hydrocarbon group, the tendency
interfaces, and ultimately determines the interaction being to force the latter into the air phase. Thus, the
heteropolar structure of the surfactant molecule leads
B to its adsorption. Frother molecules concentrate at the
Presented at Centenary of Flotation Symposium, Brisbane, Aus-
tralia, 5–9 June 2005.
surface layer with the non-polar groups oriented to-
* Corresponding author. Fax: +61 2 49216920. wards the air and the polar groups towards the water.
E-mail address: Anh.Nguyen@newcastle.edu.au (A.V. Nguyen). Molecules of ionising collectors strongly adsorb at the
0301-7516/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.minpro.2005.11.007
A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26 19

water–mineral interface with the non-polar groups molecules from the subsurface layer onto the interface.
oriented towards the water and polar groups towards The modelling will be developed based on these two
the mineral surface. Therefore, the major role of the characteristic steps.
collectors is to make the mineral surface less water
wetted (hydrophobic). Alcohols and related com- 2.1. Surfactant molecular diffusion
pounds such as the glycol ethers are most widely
used as frothers, largely because of their inability to The molecular diffusion of surfactants is well de-
adsorb at mineral particles. Ionic collectors are in scribed by the diffusion equation (Fick’s law)
many respects chemically similar to frothers and, in-
Bc
deed, many of the collectors, such as oleates, sul- ¼ Dj2 c ð1Þ
phates and sulphonates, are powerful frothers, being Bt
in fact too powerful to be used as efficient frothers,
where c is the surfactant concentration, D is the mo-
since the froths which they produce can be too stable
lecular diffusivity, t is the reference time and j2 is the
to allow efficient transport to further processing. The
Laplace operator.
adsorption of both the frothers and collectors at the
For a planar gas–liquid interface, the partial differ-
gas–liquid interface leads to:
ential Eq. (1) reduces to a one-dimensional case, which
can be analytically solved. In terms of the dynamic
! Decrease in the bubble size, resulting in the increase
surface excess, C(t), of the adsorbed surfactant mole-
in the specific area of the gas–liquid interface for the
cules, the solution of the one-dimensional diffusion
particle attachment. Adsorbed frother molecules pre-
equation can be used to obtain the celebrated Ward
vent bubbles from coalescence.
and Tordai (1946) expression, which is widely used
! Decrease in the bubble rise velocity (decrease the
in the modelling of the dynamic surface tension
interface mobility, the bubble size) and the bubble
(Chang and Franses, 1995; Eastoe and Dalton, 2000).
residence time.
The celebrated Ward and Tordai equation gives
! Increase in the stability of the froth phase for carry-
ing hydrophobic particles to the concentrate. rffiffiffiffi Z t 
D pffi / ðt Þ
Cðt Þ ¼ 2cb t  pffiffiffiffiffiffiffiffiffiffi ds ð2Þ
In addition, the adsorbed frother molecules at the p 0 ts
gas–liquid interface can also increase the efficiency of
the bubble–particle attachment as shown some time where c b is the surfactant concentration in the bulk
ago by Leja and Schulman (1954) and Leja (1982). solution, /(t) is the surfactant concentration at the
Bubble–particle interactions, gas dispersion into fine subsurface layer and s is the integration variable.
bubbles, and froth stability and drainage involve many The use of Eq. (2) for adsorption at planar surfaces
transient processes and, for these reasons, it is instru- in the modelling of the dynamic adsorption of surfac-
mental to characterize the dynamic adsorption behav- tant molecules onto a flotation air bubble can be ques-
iour of the frothers and collectors at the gas–liquid tioned since the gas–liquid interface is strongly curved.
interface. This paper focuses on the modelling of the To check the applicability of Eq. (2) for flotation sys-
dynamic adsorption of the surfactants onto flotation air tems, below we solve Eq. (1) for air bubbles and
bubbles of order of 1 mm in diameter. The dynamic establish the asymptotic conditions of the modelling
adsorption is modelled based on both molecular dif- results.
fusion and adsorption kinetics. The suitability of the
modelling is investigated by measuring the dynamic 2.1.1. Surfactant diffusion towards a spherical air bub-
surface tension as a function of time and comparing ble surface
the results with the model predictions. The bubble geometry and the boundary conditions
used to solve Eq. (1) are shown in Fig. 1. The spherical
2. Theory coordinate system with the origin located at the bubble
centre can be conveniently employed to solve Eq. (1)
The dynamic adsorption of surfactant onto the gas– since the diffusion process is spherically symmetrical.
liquid interface is determined by two steps, involving In the coordinate system, Eq. (1) reduces to
(1) the molecular diffusion of the surfactant molecules  
Bcðr; t Þ D B 2 Bcðr; t Þ
from the bulk solution to the subsurface layer at the ¼ 2 r ð3Þ
interface and (2) the actual adsorption of the surfactant Bt r Br Br
20 A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26

r=

8
c = cb
Applying the boundary conditions given by Eqs. (10)
and (11) yields
r=R
c = φ (t) a¼0 ð14Þ
r
  
c0 pffiffiffi
b ¼ R /˜ ðpÞ  exp R k : ð15Þ
p
The distribution of the surfactant concentration
Bubble around the bubble surface now can be determined by
taking the inverse Laplace transform of Eq. (12). How-
ever, the dynamic surface excess, C(t), of the surfactant
Fig. 1. Bubble geometry and domain for the physical modelling. can be better determined in the following way.

where r is the radial coordinate. The initial condition is 2.1.2. Surface excess of adsorbed surfactant molecules
described by at the bubble surface
The dynamic surface excess, C(t), can be determined
cðr; 0Þ ¼ cb ð4Þ from the mass balance at the interface, which is des-
The boundary conditions are described by cribed by
 
cðl; t Þ ¼ cb ð5Þ dC Bc
¼D ð16Þ
dt Br r¼R
cð R; t Þ ¼ /ðt Þ ð6Þ
Since C(0) = 0, taking the Laplace transform of Eq. (16)
where R is the bubble radius. The partial differential yields
Eq. (3) can be analytically solved employing the  
dc̃c ðr; pÞ
Laplace transform given by C ð pÞ ¼ D
pC̃ ð17Þ
Z l dr r¼R
cðr; t Þ ¼ c̃c ðr; pÞexpð  pt Þdp ð7Þ Inserting Eq. (12) into Eq. (17), one obtains
0
" sffiffiffiffi! sffiffiffiffiffi!#
Z l D D ˜ D D
c̃c ðr; pÞ ¼ cðr; t Þexpð  pt Þdt ð8Þ C ð pÞ ¼ 
C̃ þ / ð pÞ þ cb þ
0
Rp p Rp2 p3

Applying the Laplace transform to Eq. (3) gives ð18Þ


 2 
d c̃c 2 dc̃c Taking the inverse Laplace transform of Eq. (18) gives
D þ ¼ pc̃c ðr; pÞ  cb ð9Þ rffiffiffiffi
dr2 r dr Z t 
D pffi /ðsÞ
The Laplace transforms of the boundary conditions Cðt Þ ¼ 2cb t  pffiffiffiffiffiffiffiffiffiffi ds
p 0 ts
given by Eqs. (5) and (6) give  Z t 
Z l D
þ tcb  /ðsÞds : ð19Þ
c̃c ðl; pÞ ¼ ept cðl; t Þdt ¼ cb =p ð10Þ R 0
0
Z l
The first term on the right-hand side of Eq. (19) is
c̃c ð R; pÞ ¼ ept /ðt Þdtu/˜ ðpÞ ð11Þ the same as that of the celebrated Ward and Tordai
0 equation, Eq. (2). The second term of Eq. (19) is a
The general solution of Eq. (9) is as follows function of the bubble radius and describes the correc-
 pffiffiffi  pffiffiffi tion to the Ward and Tordai equation due to the curva-
aexp r k þ bexp  r k cb ture effect of the spherical bubble surface.
c̃c ðr; pÞ ¼ þ ð12Þ
r p
2.2. Adsorption of surfactant molecules from the sub-
where a and b are the integration constants and k is
surface layer
defined by
p The adsorption of surfactant molecules from the
k¼ ð13Þ
D subsurface layer is usually described by the Langmuir
A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26 21

adsorption isotherm (Adamson and Gast, 1997), i.e. with the subsurface concentration. Therefore, the dy-
namic adsorption is controlled by the surfactant diffu-
Cðt Þ
/ðt ÞK ¼ ð20Þ sion from the bulk solution to the interface and is
C m  C ðt Þ often referred to as the diffusion-controlled dynamic
where K is the (equilibrium) adsorption constant and adsorption.
C m is the maximum of C. The dynamic surface excess In practice, the adsorption from the subsurface layer
can be described in terms of the experimentally mea- need not be instantaneous and may be controlled by the
sured dynamic surface tension by the Gibbs adsorption adsorption kinetics itself. For this case, the local equi-
equation (Adamson and Gast, 1997), which can be librium is no longer applied and the dynamic adsorption
described by is known as kinetics-controlled adsorption. The gener-
alized adsorption kinetics gives (Miller and Kretzsch-
c dr mar, 1980; Borwankar and Wasan, 1983)
C¼  ð21Þ
nRg T dc

dCðt Þ Cðt Þ
where r is the surface tension, R g is the universal gas ¼ ka /ðt Þ 1 
dt Cm
constant, n = 1 for non-ionic surfactants and n = 2 for

ionic surfactants, and T is the absolute temperature. Cðt Þ Cðt Þ


 kd exp  2b ð25Þ
Applying Eq. (21), with c being the subsurface con- Cm Cm
centration, /(t), and inserting Eq. (20) into Eq. (21) where k a and k d are the adsorption and desorption rate
yields constants, respectively. Note that the equilibrium ad-

sorption constant is equal to K = k a/k d and, if k aYl
C ðt Þ
r0  rðt Þ ¼  nRg T Cm ln 1  ð22Þ (adsorption is instantaneous), the kinetics-controlled
Cm
model reduces to the diffusion-controlled model.
where r 0 is the surface tension of the clean surface
(pure solvent). Eq. (22) can be used, together with 2.3. Numerical computation methodology
Eqs. (20) and (19), to validate the dynamic adsorption
theory. Numerical computation is required to solve the gov-
There exist a number of extensions to the Langmuir erning Eqs. (19), (20), (22)–(25) for the dynamic ad-
adsorption theory. Firstly, the interaction among the sorption process. Firstly, the time domain (0,t) can be
adsorbed molecules can be included, and leads to the partitioned into i small intervals by Dt j = t j  t j1, where
Frumkin adsorption isotherm (Lyklema, 1991), which j = 1, 2, 3. . .i. Eq. (19) is then discretised by applying
is described by the trapezoidal rule approximation to the integral, lead-
  ing to
Cðt Þ Cðt Þ
/ðt ÞK ¼ exp  2b ð23Þ
C m  C ðt Þ Cm gðti Þ ¼ A  /ðti ÞB ð26Þ
where b is a parameter characterizing the lateral inter- where
action among adsorbed surfactant molecules. The gðti Þ ¼ Cðti Þ=Cm ð27Þ
corresponding surface tension isotherm determined pffiffiffiffiffiffiffiffiffi (
from the Gibbs adsorption equation is then D=p pffiffiffi Xi1 

A¼ 2cb ti þ / tj þ / tj1
Cm
C ðt Þ j¼1
r0  rðt Þ ¼  Rg T Cm ln 1  )
Cm pffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffi

2  ti  tj  ti  tj1  /ðti1 Þ Dti
Cðt Þ
 bRg T Cm : ð24Þ (
Cm D X
i1 
þ 2ti cb  / tj þ / tj1
2Cm R
Eqs. (24), (23) and (19) can also be used to validate j¼1
)
the dynamic adsorption theory, in relation to non-
ideal behaviour and lateral interaction of the adsorbed  Dtj  /ðti1 ÞDti ð28Þ
molecules.
Both Eqs. (20) and (23) imply that the adsorption pffiffiffiffiffiffiffiffiffi
from the subsurface layer is considered instantaneous D=p pffiffiffiffiffiffi D
B¼ Dti þ Dti ð29Þ
and the surface concentration is always in equilibrium Cm 2Cm R
22 A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26

For diffusion-controlled dynamic adsorption, Eq. A Millipore Milli-Q water purification system was
(26) can be used in conjunction with either Eq. (20) used to rinse all components and to make all solutions.
or Eq. (23) to numerically solve for the transient sub- The Millipore system provides water with a surface
surface concentration, /(t), and the dynamic adsorption tension of about 72 mN/m. All glassware and compo-
excess, C(t). The initial conditions are /(0) = 0 and nents used in the experiments were soaked for several
C(0) = 0. hours in 6% (by volume) solutions of Deconex 15E
For kinetics-controlled dynamic adsorption, Eq. (25) (Borer, Switzerland), rinsed in the Milli-Q water and
can be further discretised by applying the implicit dried in a cabinet to protect them from dust. The
scheme to give needles were also cleaned using a dichromate mixture
and then sonicated repeatedly in the Milli-Q water. All
gðti Þ  g ðti1 Þ surfactant solutions were freshly prepared and used
¼ ka /ðti Þ½1  g ðti Þ
Dti immediately.
 kd g ðti Þexp½  2bg ðti Þ: ð30Þ
3.2. Apparatus
Eqs. (30) and (26) can be simultaneously solved for
the dynamic adsorption excess, C(t), by applying the The dynamic adsorption was experimentally inves-
initial conditions described previously. tigated by measuring the dynamic surface tension ver-
sus concentration using the pendant bubble method
3. Experimental (Adamson and Gast, 1997) with a fully computer-con-
trolled apparatus OCA 20 from DataPhysics (Ger-
3.1. Materials many). The experimental set-up for the pendant
bubble apparatus is shown in Fig. 2.
The surfactants used to study the dynamic adsorp- A Perspex cell was filled with the surfactant solution
tion included sodium dodecylbenzene sulphate (SDBS) and placed in the path of a collimated light beam. A
(Aldrich, Germany), which is used as an anionic col- bubble was formed on an inverted needle immersed in
lector, and Dowfroth 250 (Dow Chemical, USA), the solution. The light beam cast a bubble shape onto a
which is a mixture of polyglycols used as a flotation CCD camera. The image was recorded and stored in a
frother. The surfactants were used without any further computer for further image analysis. The bubble was
purification. rapidly formed using a gas-tight microsyringe, which

Camera

Fig. 2. Experimental set-up for the pendant bubble tensiometry measurements.


A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26 23

75

65

Surface tension (mN/m)


Planar interface

55
R = 100 μm
R = 1 μm R = 10 μm
45

35

25
0.1 1 10 100 1000
Time (ms)

Fig. 3. Theoretical predictions for the influence of the bubble size on the dynamic surface tension (C m = 5  10 6 mol/m2, c b = 5  10 4 mol/L,
K = 7.0 m3/mol, b = 0 and D = 5  10 10 m2/s).

was motorized and fully controlled by a computer. The time difference between the first recorded image
Typically, a droplet with volume of approximately and the starting time (t = 0) of the bubble (theoretical
11 AL was formed within a time less than 0.5 s. The moment of the interface formation) was determined by
camera was set to strobe, thereby obtaining the bubble fitting the model parameters to the data. The difference
shape as a function of time. Images of the bubbles were was then used to obtain the btime-correctedQ surface
taken at the frequency of 50 Hz. The bubble shapes tension values.
were digitised and analysed off-line. The digitised bub-
ble profiles were fitted numerically with the Young– 4. Results and discussion
Laplace equation to obtain the surface tension value.
The surface tension corresponding to each bubble 4.1. Effect of the bubble size
image was found. The system was calibrated using
the outer diameter of the inverted needle. Checks on Typical results for the dependence of dynamic sur-
the calibration were performed at regular intervals by face tension on the bubble size are shown in Fig. 3. The
measuring the surface tension of purified Milli-Q water. results were obtained using Eqs. (23), (24) and (26),

75
Surface tension (mN/m)

Experimental data
65
Diffusion & planar interface
Diffusion & spherical interface
55
Diffusion-kinetics & spherical interface

45

35
0 10 20 30 40
Time (s)
Fig. 4. Experimental (points) and predicted (lines) dynamic surface tension for a 5  10 4 mol/L SDBS solution. The bubble radius used in the
experiment and modelling was R = 1.24 mm. The parameters obtained by the best fit are summarised in Table 1.
24 A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26

Table 1
Adsorption model parameters obtained by the best fit for 5  10 4 mol/L SDBS solution
Model K (m3/mol) C m (mol/m2) b D (m2/s) R (mm) k d (1/s)
6
Diffusion-controlled/planar interface 7.66 4.76  10 0.057 2.37  10 11 – –
Diffusion-controlled/spherical interface 7.95 4.83  10 6 0.057 1.97  10 11 1.24 –
Kinetics-controlled/spherical interface 8.81 4.07  10 6 0.057 2.00  10 10 1.24 1.24

i.e., diffusion-controlled dynamic adsorption was con- governed by the molecular diffusion and SDBS can be
sidered. The model parameters as given in the text considered as a diffusive surfactant. In this study, the
capture to Fig. 3 are within the range for typical hy- adsorption dynamics of SDBS at the interface of an air
drocarbon-based surfactants (Chang and Franses, bubble is investigated using three models: (1) the
1995). The numerical value of 5  10 6 mol/m2 for diffusion-controlled model for a planar interface, (2)
C m gives the corresponding molecular cross-sectional the diffusion-controlled model for a spherical interface
area of ~ 33 Å2 at the maximum packing for adsorbed and (3) the kinetics-controlled (the kinetics-diffusion
surfactants. Similar results were obtained with the ki- mixed) model for a spherical interface. The best fit of
netics-controlled dynamic adsorption model using the the three model predictions to the experimental results
same values for model parameters. was carried out in Excel using the built-in function
Fig. 3 shows clearly that the bubble size significant- Solver and the user-defined macro written in the VBA
ly influences the dynamic surface tension and the dy- (Visual Basic for Application) programming language.
namic adsorption onto micrometer size bubbles. For It can be seen from Fig. 4 that all of the three model
typical flotation operations, where air bubbles have predictions match the experimental data. For all the
diameter of the order of 1 mm, the bubble size effect three models, the best-fit values for C m (given in
is not significant. The agreement between the our Table 1) for SDBS are within the normal range for
model with the inclusion of the bubble radius in Eq. surfactant which is from 1 to 10  10 6 mol/m2
(19) and the Ward and Tordai expression by Eq. (2) for (Chang and Franses, 1995). However, both the diffu-
a planar surface is within the width of the line shown in sion-controlled models for planar and spherical surfaces
Fig. 3. predict smaller diffusion coefficients. The difference in
the surface tension between the two models is small,
4.2. Comparison with experimental data for the dynamic which confirms the theoretical results presented in Fig.
surface tension 3. The diffusion-kinetic-mixed model predicts the ef-
fective diffusion coefficient comparable with the calcu-
A previous investigation using the pendant drop lated value for SDBS (Phan et al., 2004). The close
method (Phan et al., 2004) showed that the dynamic agreement of the model prediction for diffusion coeffi-
adsorption of SDBS onto the pendant drop interface is cient with the reported results indicates that an adsorp-

75

70 Experiment
Surface Tension (mN/m)

Planar interface
65
Spherical interface
60

55

50

45
0 100 200 300 400
Time (s)

Fig. 5. Experimental (points) and predicted (lines, kinetics-controlled model) dynamic surface tension for a 100 ppm (part-per-million concentra-
tion) Dowfroth 250 solution. The bubble radius used in the experiment and modelling was R = 1.36 mm.
A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26 25

Table 2
Adsorption model parameters obtained by the best fit for 100 ppm Dowfroth 250 solution
Model K (m3/mol) C m (mol/m2) b D (m2/s) R (mm) k d (1/s)
6
Kinetics-controlled/planar interface 16.713 3.97  10 0.61 2.25  10 12 – 0.106
Kinetics-controlled/spherical interface 13.086 4.14  10 6 0.34 6.55  10 11 1.36 0.110

tion barrier, even though very weak, still influences the and the surface tension gradient. For example, the
dynamic adsorption of SDBS. Therefore, SDBS is not a interface mobility is expected to change the liquid
perfectly diffusive surfactant. It requires the diffusion- flow at the bubble surface and to influence the capture
kinetic-mixed model for predicting the dynamic surface efficiency but it still remains to be quantified (Nguyen,
tension. 1999). The analysis given in this paper will provide
The effect of the bubble radius on the diffusion- important inputs for further developments of the flota-
kinetic-mixed mechanism for dynamic adsorption was tion theory, specifically the role of the adsorbed surfac-
also investigated by matching the models with the tants at the gas–liquid interface in the particle collection
dynamic surface tension measured for Dowfroth 250. phenomena.
The results are shown in Fig. 5. Dowfroth 250 is a
mixture of polyglycols, which have the slow adsorp- 5. Conclusions
tion characteristics of non-diffusive surfactants. The
diffusion-kinetic-mixed model was applied to the The dynamic adsorption of collectors and frothers
Dowfroth 250 experimental data in conjunction with onto the air bubble surface was theoretically analysed
the diffusion equations for the spherical and planar using the molecular diffusion and the surfactant adsorp-
interfaces. The best-fit predictions are shown in Table tion processes. The governing equations were solved for
2. The difference in the two best-fit values for the the dynamic surface excess of surfactants, which was
effective diffusion coefficients indicates a small effect described in terms of the experimentally measured dy-
of the bubble size, which is within the experimental namic surface tension. The adsorption kinetics was
error. The dynamic adsorption of Dowfroth 250 onto modelled using the Langmuir and Frumkin adsorption
air bubbles can be described using the Ward and isotherms. The dynamic adsorption of the surfactants
Tordai solution for surfactant diffusion toward a pla- was measured for the dynamic surface tension of
nar gas–liquid interface. sodium dodecylbenzene sulphonate and Dowfroth
250 frother using pendant bubble tensiometry. The
4.3. Relevance to the flotation theory measurements of the dynamic surface tension were
carried out using the pendant bubble method. The
Both the theoretical analysis and the experimental theoretical analysis using the bubble radius, R, be-
data presented in this paper show that the dynamic tween 1 Am and 10 mm, and the comparison with
surface tension and the dynamic adsorption onto the experimental data obtained using R = 1.24 and 1.36
micrometer size bubbles are not significantly influ- mm showed that the bubble size effect on the dynamic
enced by the bubble size. Therefore, the dynamic adsorption is negligibly small for air bubbles with the
adsorption of surfactants in flotation can be effective- diameter of the order of 1 mm, which is the typical
ly analysed using the theories applied to the planar range of the bubble size encountered in flotation.
gas–liquid interface. In particular, the last term on the Therefore, the dynamic adsorption onto the bubble
right hand side of Eq. (19) can be safely neglected, surface in flotation can be simply described using
giving the simpler expression for predicting the dynam- the Ward and Tordai equation derived from the diffu-
ic surface excess, C(t), by Eq. (2) of the Ward and sion equation for a planar interface. Furthermore, the
Tordai theory. At short adsorption time, the dynamic experimental data for SDBS were well described by
surface excess can be determined pffiffiffiffiffiffiffiffiffiffiby the Ward and the diffusion-controlled model with the Frumkin ad-
Tordai theory by Cðt Þ ¼ 2cb Dt=p, which does not sorption isotherm. However, the experimental data for
require the numerical computation. The simplified pre- the dynamic surface tension of Dowfroth 250 were
diction of the dynamic surface excess will allow a described by the kinetics-controlled model and the
number of important factors affecting the particle col- molecular diffusion. The presented results are signifi-
lection by air bubbles to be determined. The important cant for further analysis of the effect of the adsorbed
factors include the mobility of the gas–liquid interface surfactants at the bubble surface on the bubble–parti-
of rising air bubbles, the elasticity of the bubble surface cle attachment phenomena in flotation.
26 A.V. Nguyen et al. / Int. J. Miner. Process. 79 (2006) 18–26

Acknowledgement Leja, J., 1982. Surface Chemistry of Froth Flotation. Plenum Press,
New York, NY. 758 pp.
Leja, J., Schulman, J.H., 1954. Flotation theory: molecular interac-
The authors acknowledge the Australia Research tions between frothers and collectors at solid–liquid–air interfaces.
Council for financial support. Technical Publication 3693-B.
Lyklema, J., 1991. Fundamentals of Interface and Colloid Science,
vol. 1. Academic Press, San Diego. 673 pp.
References Miller, R., Kretzschmar, G., 1980. Numerical solution for a mixed
model of diffusion kinetics-controlled adsorption. Colloid and
Adamson, A.W., Gast, A.P., 1997. Physical Chemistry of Surfaces. Polymer Science 258, 85 – 87.
John Wiley & Sons, Inc., New York. 784 pp. Nguyen, A.V., 1999. Hydrodynamics of liquid flows around air
Borwankar, R.P., Wasan, D.T., 1983. The kinetics of adsorption of bubbles in flotation: a review. International Journal of Mineral
surface active agents at gas–liquid surfaces. Chemical Engineering Processing 56 (1–4), 165 – 205.
Science 38 (10), 1637 – 1649. Phan, C.M., Nguyen, A.V., Evans, G.M., 2004. Measurement and
Chang, C.H., Franses, E.I., 1995. Adsorption dynamics of surfactants modelling of the dynamic adsorption of flotation collectors and
at air/water interface: a critical review of mathematical models, frothers at the gas–liquid interface. Proc. 32nd Australasian
data and mechanisms. Colloids And Surfaces. A, Physicochemical Chemical Engineering Conference (Chemeca 2004), 27–29 Sep-
And Engineering Aspects 100, 1 – 45. tember, Sydney, Australia.
Eastoe, J., Dalton, J.S., 2000. Dynamic surface tension and adsorption Ward, A.F.H., Tordai, L., 1946. Time-dependence of boundary ten-
mechanisms of surfactants at the air–water interface. Advances in sions of solutions: I. The role of diffusion in time effects. The
Colloid and Interface Science 85, 103 – 144. Journal of Chemical Physics 14 (7), 453 – 461.

You might also like