49
A Critical Review of the Growth, Drainage and Collapse of Foams Jianlong Wang, Anh V. Nguyen, Saeed Farrokhpay PII: S0001-8686(15)00217-1 DOI: doi: 10.1016/j.cis.2015.11.009 Reference: CIS 1600 To appear in: Advances in Colloid and Interface Science Please cite this article as: Wang Jianlong, Nguyen Anh V., Farrokhpay Saeed, A Critical Review of the Growth, Drainage and Collapse of Foams, Advances in Colloid and Interface Science (2015), doi: 10.1016/j.cis.2015.11.009 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

A Critical Review of the Growth, Drainage and Collapse of Foams378148/UQ378148... · 2019. 10. 11. · formation of foam by destroying the liquid film (Section 2.2.2). Solid particles

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • �������� �������

    A Critical Review of the Growth, Drainage and Collapse of Foams

    Jianlong Wang, Anh V. Nguyen, Saeed Farrokhpay

    PII: S0001-8686(15)00217-1DOI: doi: 10.1016/j.cis.2015.11.009Reference: CIS 1600

    To appear in: Advances in Colloid and Interface Science

    Please cite this article as: Wang Jianlong, Nguyen Anh V., Farrokhpay Saeed, A CriticalReview of the Growth, Drainage and Collapse of Foams, Advances in Colloid and InterfaceScience (2015), doi: 10.1016/j.cis.2015.11.009

    This is a PDF file of an unedited manuscript that has been accepted for publication.As a service to our customers we are providing this early version of the manuscript.The manuscript will undergo copyediting, typesetting, and review of the resulting proofbefore it is published in its final form. Please note that during the production processerrors may be discovered which could affect the content, and all legal disclaimers thatapply to the journal pertain.

    http://dx.doi.org/10.1016/j.cis.2015.11.009http://dx.doi.org/10.1016/j.cis.2015.11.009

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    1

    A Critical Review of the Growth, Drainage and Collapse

    of Foams

    Jianlong Wang1, Anh V Nguyen

    1* and Saeed Farrokhpay

    2

    1School of Chemical Engineering, University of Queensland, Brisbane, Queensland 4072,

    Australia 2JKMRC, University of Queensland, Brisbane, Queensland 4072, Australia

    *Correspondence: [email protected]

    Highlights

    • Mechanisms governing foamability and foam stability are reviewed.

    • Studies on interfacial properties and foam films are inadequate for foam properties.

    • Solid particles significantly affect film stability and foam drainage.

    • Foam growth and collapse are poorly quantified in the literature.

    Graphic abstract

    Abstract

    This review focuses on the current knowledge regarding (i) the mechanisms

    governing foamability and foam stability, and (ii) models for the foam column kinetics.

    Although different length scales of foam structure, such as air-water interface and liquid film,

    have been studied to elucidate the mechanisms that control the foamability and foam stability,

    many questions remain unanswered. It is due to the collective effects of different mechanisms

    involved and the complicated structures of foam sub-structures such as foam films, Plateau

    borders and nodes, and foam networks like soft porous materials. The current knowledge of

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    2

    the effects of solid particles on liquid film stability and foam drainage is also discussed to

    highlight gaps in our present level of understanding foam systems with solid particles. We

    also critically review and summarize the models that describe macroscopic foam behaviors,

    such as equilibrium foam height, foam growth and collapse, within the context of the

    mechanisms involved.

    Keywords: foam, foam stability, foam drainage, foamability, froth flotation

    Contents

    1. INTRODUCTION ............................................................................................................................................ 3

    2. MECHANISMS GOVERNING FOAMABILITY AND FOAM STABILITY ................................................................ 4

    2.1 EFFECTS OF INTERFACIAL PROPERTIES ON FOAM PROPERTIES ...................................................................................... 4

    2.1.1 Surface tension and foamability ........................................................................................................... 4

    2.1.2 Surface viscoelasticity and foam properties .......................................................................................... 7

    2.2 EFFECTS OF LIQUID FILM PROPERTIES ON FOAM PROPERTIES ..................................................................................... 11

    2.2.1 Intermolecular forces between two interfaces ................................................................................... 11

    2.2.2 Effects of antifoam behaviors on liquid film stability .......................................................................... 12 2.2.2.1 Entry barrier .................................................................................................................................................. 13 2.2.2.2 Role of oil spreading in antifoam performance ............................................................................................. 14 2.2.2.3 Effect of adsorption kinetics of surfactants on the antifoam performance .................................................. 15

    2.2.3 Effects of solid particles on liquid film stability ................................................................................... 16 2.2.3.1 Minimum energy to remove particle at interface,

    removeG and maximum capillary pressure for the onset

    of coalescence, maxcP ................................................................................................................................................. 17

    2.2.3.2 Bridging-Dewetting mechanism of liquid film rupture by solid particle ........................................................ 18 2.3 FOAM DRAINAGE FOR AQUEOUS FOAMS ............................................................................................................... 19

    2.4 FOAM DRAINAGE IN THE PRESENCE OF SOLID PARTICLES ........................................................................................... 20

    3. FOAM COLUMN KINETICS .......................................................................................................................... 25

    3.1 MODELS OF EQUILIBRIUM FOAM HEIGHT .............................................................................................................. 25

    3.1.1 Hrma et al.’s model ............................................................................................................................. 25

    3.1.2 Hartland et al.’s model ........................................................................................................................ 26

    3.1.3 Pilon et al.’s models ............................................................................................................................ 28

    3.1.4 Limitations of the models for equilibrium foam height ....................................................................... 32

    3.2 MODELS OF FOAM GROWTH AND COLLAPSE .......................................................................................................... 32

    3.2.1 Models without consideration of foam collapse ................................................................................. 33

    3.2.2 Empirical equations for foam growth and collapse ............................................................................ 35

    3.2.3 Foam growth model relating to critical Plateau border size ............................................................... 37

    3.3 SUMMARY ..................................................................................................................................................... 38

    4. CONCLUDING REMARKS ............................................................................................................................. 39

    ACKNOWLEDGEMENTS .................................................................................................................................. 40

    REFERENCES ................................................................................................................................................... 41

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    3

    1. INTRODUCTION

    Foams are highly concentrated dispersions of gas (dispersed phase) in a liquid

    (continuous phase) [1]. Due to their lightness and large specific surface areas, foams are

    widely applied in our daily lives and industry. Examples of applications in which foams are

    used [2] include food, cosmetics, cleaning, surface treatment, building materials, reducing

    pollution (e.g. wastewater treatment [3] and heavy metal removal [4] by foam fractionation)

    and extraction of nature resources, i.e., froth flotation.

    When producing a column of foam, different mechanisms either to produce and

    stabilize the foam column, such as the formation of liquid films and foams or to destroy it,

    such as coarsening of foams and drainage and rupture of liquid film and foams, get involved

    [2]. The combined effects of these mechanisms determine the lifetime of a foam or foam

    stability. The foamability of a solution is a measure of its capacity to produce a foam. During

    the foam formation, the surface energy E A increases with the creation of gas-liquid

    interfaces with the surface tension, and interfacial area A. The increased surface energy

    means the foam formation is not a spontaneous process, and the input of energy is

    indispensable to generate a column of foam. Based on the different ways to put energy into

    the liquid phase to generate foam, foaming techniques are categorized into physical, chemical

    and biological foaming [5]. We only focus on the mechanical foaming (the rotor-stator mixer

    method in Section 2.1.1 and the bubbling method in Section 3) that belongs to the physical

    foaming techniques in this review for the sake of its common application. Surfactants must

    also be present in the solution to promote the production of foam. The addition of surfactants

    stabilizes the liquid film and foams by altering the static surface tension (Section 2.1.1),

    adsorption kinetics (Section 2.1.1), surface viscoelasticity (Section 2.1.2) and interactions

    between two interfaces of a liquid film (Section 2.2.1). On the contrary, antifoams prevent the

    formation of foam by destroying the liquid film (Section 2.2.2). Solid particles also play an

    important role in the film stability (Section 2.2.3) and foam drainage (Section 2.4). Foam

    drainage for the aqueous foams (Section 2.3) is also discussed due to its importance for the

    foam growth and collapse. Finally, the models for the foam growth and collapse are reviewed

    (Section 3).

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    4

    2. MECHANISMS GOVERNING FOAMABILITY AND FOAM STABILITY

    Foamability and foam stability are two main foam properties of surfactant solutions.

    Foamability is an overall capacity of surfactant solution to produce foams, whereas foam

    stability refers to the lifetime of a foam column. These two terms are interrelated. For

    example, the foamability of a transient foam is believed to depend on its stability. Although

    these two terms are commonly used in the literature, there are no simple universal physical

    parameters to quantify them. Foam height and foam lifetime have been applied to

    characterize foamability and foam stability [6-8]. However, these two criteria are not

    satisfactory because they are not only dependent on the chemical composition of foaming

    solutions but also on the method of foam generation. Moreover, a general theory to explain

    the mechanisms of foam formation and stability for all types of foam system does not exist

    [9-11] because the magnitude and mutual importance of the different types of effects can vary

    significantly, depending on the stage of foam life and the conditions of its existence [10].

    Bearing in mind the complicated interplay of various mechanisms, this review focuses on the

    existing theories that describe the mechanisms governing the foamability and foam stability.

    2.1 Effects of interfacial properties on foam properties

    The adsorption of surfactant molecules on the air-water interface alters the interfacial

    properties, and the foam properties change accordingly.

    2.1.1 Surface tension and foamability

    If the external energy applied to generate the foam is constant, then the foam surface

    area is inversely proportional to the surface tension. Therefore, a lower surface tension will

    increase the foamability of a solution from the perspective of surface energy. The other factor

    that influences the foamability is the bubble breakup. Foam generation involves creation and

    deformation of gas-liquid interface. When the mechanical method applies to foam generation,

    bubbles deform due to the external forces (e.g., shear stress or pressure of turbulence) that are

    subjected to the bubble surface. If the deformation is significantly large, bubbles can be split

    into smaller ones, and foam generation is possible. The critical Weber number, which is a

    dimensionless ratio between the inertial force (causing the deformation) and the surface

    tension (restoring the bubble sphericity), has been used as the criterion of the bubble breakup

    [12-18]. The bubble breakup occurs if the external forces applied to the bubbles make the

    Weber number exceed its critical value, which is of the order of unity. A lower surface

    tension ends up a large Weber number, therefore, promotes the bubble breakup process.

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    5

    Although the role of surface tension in the foamability has been recognized regarding

    surface energy and bubble breakup process, the equilibrium surface tension is not reached

    instantaneously, and the dynamics of adsorption of surfactant molecules must be considered

    when the adsorption time is longer than the time scale of foam generation [10]. The evolution

    of the surface tension value is controlled by two processes: (i) the diffusion of surfactant

    molecules to the surface and (ii) the adsorption of surfactant molecules on the interface,

    which must overcome an associated energy barrier. Either of these processes may become the

    rate-determining step. The average lifetime of the bubbles at the solution/air interface during

    foam generation has been recommended as a rational choice of an adsorption reference time

    [19].

    Research efforts of using different designs have focused on the effect of dynamic

    surface tension on the foamability. For example, the relationship between foam formation and

    dynamic surface tension of non-ionic and anionic surfactants has been studied using a rotor

    test, in which air is introduced to the surfactant solutions by a special stirring device [19]. A

    good correlation between foam formation and dynamic surface tension values at t = 100 ms

    has been found (Figure 1). The relationship between foam formation and dynamic surface

    tension has also been investigated using the Ross-Miles test [20-22]. Recently, the foam

    formation by using a sparger was also related to the dynamic surface tension [23]. Figure 2

    shows a good correlation of foam height with the surface tension reduction rate, which is

    defined as follows [24]:

    01/2 *2

    mRt

    (1)

    where 0 is the surface tension of the solvent, m is the quasi-equilibrium surface tension at

    which the decrease rate of the surface tension is smaller than 1 mN/m per 30s, and *t is the

    time when the surface tension is equal to 0 / 2m . Although the correlation between

    foam height and surface tension reduction rate has been found, we note that the surface

    tension reduction rate in Eq. (1) [22, 24] is only an empirical equation and lacks a

    fundamental basis.

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    6

    Figure 1. The relative dynamic pressure at t = 100 ms and the rate of foam formation for the

    non-ionic surfactant C12(EO)6 (a, b) and the anionic surfactant C12SO3Na (a’, b’). The rate of

    foam formation refers to a rotor speed of 900 min-1

    . The graphs are reproduced from [19].

    Figure 2. Overview of the results from the paper by Rosen et al. [22], showing the relation

    between R1/2 and the foam height in a Ross-Miles test, reprinted from [23]. The solid line is

    obtained by fitting a logarithmic equation (R2 = 0.9).

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    7

    2.1.2 Surface viscoelasticity and foam properties

    A century ago the effect of surface tension gradients on foam and film stability was

    discovered and described by Gibbs and Marangoni [25]. In 1941, Levich developed a theory

    to relate surface tension gradient with surface dilational elasticity [26]. However, only until

    1970 [27] a new methodology for measuring the dilational rheology via a harmonically

    oscillating bubble has paved the way for the first commercial instrument for routine

    experimental determination of the dilational surface elasticity [28], based on oscillating drops

    and bubbles. Various aspects of surface viscoelasticity have been reviewed [29-33] and a

    book devoted to surface viscoelasticity has been published recently [34].

    The dilational elasticity modulus, E , is defined as the ratio of the surface tension

    change, , to the relative increase in surface area, /A A , i.e., / lnE d d A . Surface

    dilational elasticity and viscosity are expressed as 0' / cosE A A and

    0'' / / sinE A A , respectively, where , 0A , , are the phase shift between the

    surface tension and area oscillations, the mean surface area and the angular frequency (Figure

    3, bottom panel). For the film elasticity, its value should be two times the value of E because

    of the presence of two film interfaces. It is noted that an analogy between the Gibbs elasticity

    and the surface dilational modulus has been made [35]. The Gibbs elasticity refers to the

    increase in the film surface tension resulting from a decrease in the surfactant concentration

    within the interlamellar solution caused by the small extension of the film relative to the film

    size. In the Gibbs theory, the film elasticity originates from the deterioration of the interstitial

    surfactant solution with the assumption that the thickness of the lamellae is very small. The

    Marangoni elasticity, as related to the Marangoni effect, originates from the transport of

    surfactant molecules from the adjacent bulk phase to the interface for the case of the non-

    equilibrium state of thick foam films [36]. For soluble surfactants, the magnitude of the

    elasticity modulus depends on the frequency of external disturbances or the oscillation

    frequency. The adsorption layer will behave as an insoluble monolayer when the frequency is

    sufficiently high, and the Marangoni dilational modulus reaches its limiting value [37, 38]. At

    low frequencies, the adsorption layer behaves as a viscoelastic surface because of relaxation

    processes, i.e., diffusional exchange that occurs in and near the interface (The phase shift

    between surface tension and area changes in the bottom panel of Figure 3). Therefore, the

    modulus E has both elastic and viscous components. The surface viscoelasticity of

    adsorption layers and the origin of the dilational viscosity are analyzed in a recent review

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    8

    [39]. The dilational surface viscosity is categorized as 1) “true” dilational surface viscosity,

    which originates only from the transport of surfactant molecules at the interface, and 2)

    “apparent” dilational surface viscosity caused by the surfactant diffusion from the bulk

    solution.

    Surface viscoelasticity is related to the non-equilibrium state of the adsorption layer

    [40]. A distortion of its existing equilibrium state and the absence of adsorption equilibrium

    at a freshly created interface are the two main reasons underlying the non-equilibrium state of

    the adsorption layer. Because the interface has not yet attained its equilibrium adsorption

    coverage during foam generation, surface viscoelasticity plays a significant role in

    foamability. Assuming that the surface area of the film elements increases while its volume

    and the total amount of surfactants remain constant, and the diffusion of surfactant molecules

    can be ignored, the surface concentration of surfactant molecules decreases during liquid film

    thinning (Figure 3, top). The reduction in surface concentration induces a local increase in

    surface tension (Figure 3, bottom), which in turn resists the stretching of the liquid film.

    Consequently, bubble coalescence rate decreases. In practice, the film thinning process is

    simulated by oscillating a droplet using drop profile analysis tensiometer (Figure 3, middle).

    The surface tension gradients induced at the air-water interface can stabilize the foam films

    by retarding drainage [41].

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    9

    Figure 3. Schematic presentation of the adsorbed surfactant molecules during the thinning of

    a liquid film by stretching (top, reproduced from [42]), oscillation of a droplet using drop

    profile analysis tensiometer (middle) to simulate liquid film thinning and transient change in

    the surface tension and surface area of a droplet with a mixture of 0.5 mM SDS and 5mg/L

    dodecanol, oscillated at the frequency of 0.05 Hz (bottom). Amplitudes of surface tension and

    area, with a clear phase shift between surface tension and area changes. Reprinted from [43].

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    10

    Good correlations between the liquid film or foam stability and the surface

    viscoelasticity have been reported [38, 44-60]. However, many questions remain challenging:

    Is the good correlation between film stability and surface rheology indicative of a similar

    correlation with foam stability? Are measurements at the interfaces (i.e., surface

    viscoelasticity) and on a single film (i.e., film stability) sufficient to understand and predict

    the foamability and foam stability [2]? What is the appropriate frequency at which the surface

    viscoelasticity can correlate with the foamability and foam stability? For the first and second

    questions, because the foams cannot be considered as a simple combination of foam films,

    the correlation between film stability and surface rheology cannot be directly applied to foam

    stability. Moreover, it is challenging to correlate the foam column stability with the liquid

    film stability [61, 62]. This discrepancy has been explained by “the conditions of the

    existence of the foams and the single film were quite different and, therefore, the different

    types of forces could operate there [10].” For the last question, on the one hand, the

    frequency of the measurements of the surface viscoelasticity should be relevant to the surface

    age during foam generation (0.1–1 s) regarding foamability. On the other hand, the two main

    processes that determine the foam lifetime, that is, the Ostwald ripening and bubble

    coalescence, have been shown to be controlled by the low- and high-frequency surface

    elasticities, respectively [51]. The roles of surface viscoelasticity in the foamability and foam

    stability remain poorly understood. The link between surface viscoelasticity and foam

    properties are summarized as follows [2]:

    For a given applied stress, a large dilational surface elastic modulus results in a small

    strain, the foam film will stretch less, and is less likely ruptured, by disturbances.

    The film elasticity will restore the interface by bringing back surfactants and by limiting

    the stretch of the interface.

    The Marangoni effect will draw surfactants back to the interface and liquid into the film,

    thereby reducing the possibility of liquid film rupture.

    Low surface viscoelasticity facilitates stretching interfaces, resulting in liquid film rupture,

    but high surface viscoelasticity causes a solid-like response of the interface, causing

    fracture.

    Surface elasticity modifies the process of drainage in foam films and Plateau borders and

    controls the appearance of bell-shaped liquid drops that have a destabilizing effect on

    films.

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    11

    2.2 Effects of liquid film properties on foam properties

    2.2.1 Intermolecular forces between film interfaces

    Foams are made of liquid films and Plateau borders. Therefore, the link between liquid

    film properties and foam properties is critical to our understanding of their behaviors. Foam

    films inspired many pioneers like Boyle, Hooke [63], Newton [64], Plateau [65] and Gibbs

    [25, 66]. The history of foam and foam films is very rich as recently compiled [67]. The

    disjoining pressure and DLVO theory developed for describing the liquid film stability are

    considered as milestones in the theoretical development of foam films.

    When two bubbles meet in foam, forces act between the two interfaces. These forces

    determine the stability of the liquid film that separates the foam bubbles. The disjoining

    pressure, , was introduced by Derjaguin in 1930’s to describe the force per unit area

    between the two interfaces of a liquid film [68, 69]. The DLVO theory is based on the

    following attractive and repulsive interactions [17, 42, 70-74]:

    The van der Waals interactions originate from the electromagnetic fields between two

    dipoles. The attractive London-van der Waals disjoining pressure is considered:

    36van

    A

    h (2)

    Where A is the Hamaker constant and h is the film thickness. The van der Waals

    attractions always destabilize foam films. Therefore, repulsive forces produced by

    surfactant and electrolyte molecules are needed to balance the attractive force to

    stabilize foam films and foam dispersions.

    The electrostatic interaction becomes evident when the interfaces are electrically

    charged in the presence of ionic surfactants. The repulsion between the two charged

    interfaces stabilizes foam films. The presence of salts such as NaCl screens this

    repulsion. The Debye length defines the effective range of the Coulombic force on

    ions against their thermal diffusion at a charged interface:

    022

    Bk T

    nze

    (3)

    where 0 and are the dielectric constant of vacuum and the relative permittivity of

    the solution, Bk is the Boltzmann constant, T is the absolute temperature, n is the

    ion concentration in the solution, z is the valence of a symmetric z:z electrolyte and e

    is the elementary charge. The repulsion can be clearly observed when the film

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    12

    thickness h falls below 2 D . The contribution of electrostatic repulsion to the

    disjoining pressure approximately decays exponentially:

    exp /dl Dh (4)

    In addition to the DLVO disjoining pressure components, non-DLVO disjoining pressures,

    non-DLVO, such as the steric repulsion, hydration repulsion and hydrophobic attraction have

    also been considered [17], giving.

    van dl non DLVO (5)

    Figure 4. Schematic of a disjoining pressure isotherm for foam films that includes

    contributions from dl , van and steric for the non-DLVO pressure. Reproduced from [75].

    Shown in Figure 4 is an example of the disjoining pressure versus film thickness. A

    positive disjoining pressure in the film is fundamental to film stability and the existence of a

    foam. Therefore, a strong and long-range repulsive interaction in the film is necessary for

    good foamability and foam stability.

    2.2.2 Effects of antifoam behaviors on liquid film stability

    Antifoams are oils, hydrophobic solid particles or a mixture of both that are present in

    the solution and prevent the formation of foam [76, 77]. Antifoams have been widely used in

    many industrial applications, such as pulp and paper production, food processing, textile

    dyeing, fermentation, wastewater treatment, and the oil industry [76, 78-81]. In the froth

    flotation of naturally hydrophobic minerals, such as coal, graphite, sulfur and molybdenite,

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    13

    nonpolar collectors are used [82, 83]. They are commonly petroleum-based hydrocarbon

    liquids, such as diesel oil. Therefore, nonpolar collectors potentially exhibit antifoam

    behaviors and affect the froth stability. Moreover, some nonionic surfactants, such as

    dodecanol, can also influence the foam properties by exerting antifoam actions. In this

    section, the mechanisms of liquid antifoam actions and the effects of surfactants on antifoam

    activity are discussed. Systematic reviews discussing antifoams are available in the literature

    [76, 77, 79]. It should be noted that the terms “fast antifoams” and “slow antifoams” have

    been previously introduced [77]. The former term denotes “the antifoams whose globules can

    enter the surfaces of the foam films and to destroy these films in the early stages of film

    thinning”, whereas the latter indicates “antifoams whose globules first leave the foam films

    and destroy the foam after entering the walls of the Plateau borders.” [77]. Here, only the

    mechanisms of “fast antifoams” are discussed here.

    2.2.2.1 Entry barrier

    Antifoams must enter the liquid film to destroy a liquid film or foam layer. Antifoams

    with a low entry barrier completely collapse the foam in seconds, whereas antifoams with a

    high entry barrier require hours to destroy the foam. The interaction energy per unit area in an

    asymmetric oil-water-air film [84] and the so-called generalized entry coefficient [85-88]

    have been introduced to represent the entry barrier. However, the determination of their

    values is very difficult [77]. Alternatively, the capillary pressure of the air-water interface at

    the moment of oil drop entry, CRCP , has been proposed as a quantitative characteristic of the

    entry barrier because it is related to antifoam efficiency [89-97]. The film trapping technique

    (FTT) has also been developed to measure precisely the value of CRCP [98, 99].

    Because the antifoam activity strongly depends on the magnitude of the entry barrier

    [77], it is crucially important to understand the factors that affect the entry barrier. Here, the

    two main points discussed in the literature concerning the factors that affect the entry barrier

    of an oil droplet are summarized.

    Entry coefficient

    Antifoam activity has been correlated with the entry coefficient [76, 79]:

    AW OW OAE (6)

    where AW , OW and OA are the surface tensions of an air-water interface, oil-water

    interface, and oil-air interface, respectively. From the thermodynamic perspective, the

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    14

    condition for the emergence of an oil droplet at the air-water interface is 0E . However, it

    has been correctly stated that a positive value of E does not guarantee high antifoam

    performance because, from the perspective of kinetics, the entry barrier also plays a crucial

    role [76-80, 100-102]. Regarding their contributions to antifoam performance, the

    relationship between the entry coefficient E and the entry barrier is analogous to chemical

    thermodynamics and kinetics [77].

    Surfactant concentration

    The effect of surfactant concentration on the entry barrier has been studied previously

    [93]. Figure 5 shows that the entry barrier increases as the surfactant concentration increases.

    The oscillatory (structural) forces shown in Figure 5 is non-DLVO forces and appears in thin

    liquid films containing colloidal particles (surfactant micelles for this case). The origin of this

    structural forces has been discussed in [103]. It is noted that the effect of the surfactant

    concentration has not yet to be fully understood [77].

    Figure 5. Entry barrier of hexadecane drops, CRCP , as a function of the SDDBS concentration,

    SC [93]. All solutions contain 12 mM NaCl. Reproduced from [77].

    2.2.2.2 Role of oil spreading in antifoam performance

    When an oil droplet emerges at the air-water interface, depending on the sign of the

    spreading coefficient, AW OW OAS [104], it either spreads out ( 0S ) or bridges the

    two interfaces ( 0S ) [100-102, 105-112]. Both outcomes result in rupture of the liquid film.

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    15

    Similar to the entry coefficient, E , S is also a thermodynamic property. A positive initial

    spreading coefficient INS (defined using AW without the spreading oil) has been shown to

    contribute to antifoam performance [110, 113]. Because the calculation of S requires the

    value of the oil-water surface tension, OW , which is difficult to measure, the spreading

    pressure, i fAW , which is defined as the reduction in the equilibrium surface

    tension of the air-water interface caused by the addition of an antifoam to the aqueous surface,

    was introduced [112]. f and i are the equilibrium surface tensions of surfactant solutions

    with and without antifoams, respectively. A positive value of AW indicates that it is

    thermodynamically favorable for antifoams to spread on the surface of the surfactant solution.

    2.2.2.3 Effect of adsorption kinetics of surfactants on the antifoam performance

    Some antifoams have been found only to affect the foamability and exert no influence

    on foam stability [91, 92, 114]. For example, Figure 6 shows that in the presence of antifoams,

    the initial foam volume (foamability) generated by shaking AOT solutions is several times

    larger than that of APG solutions. However, the foam stability of APG foams is much higher

    than that of AOT foams. Solutions without antifoams exhibit both good foamability and foam

    stability. The different effects of antifoams on the foam stability of AOT and APG are

    explained by the different entry barrier values ( 125CRCP Pa for APG solutions, and

    3CRCP Pa for AOT solutions). The much higher entry barrier for APG solutions makes the

    entry of antifoams into the liquid film more difficult and results in more stable foams relative

    to AOT solutions. However, during foam generation, the antifoams become more active for

    the APG solutions because of the slow adsorption kinetics of APG molecules (Figure 7). The

    unsaturated adsorption layers make the entry of antifoams much easier relative to the fully

    saturated surfaces. It is noted that the possible higher surface viscoelasticity deduced from the

    slow adsorption kinetics of APG molecules may also contribute to the higher foam stability

    of APG foams relative to AOT foams. Additionally, the fast adsorption of AOT molecules on

    the air-water interface may facilitate foam formation and result in a higher foamability than

    the APG solutions (see Section 2.1.1).

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    16

    Figure 6. Foam volume vs. time for two surfactant solutions -10 mM AOT and 0.6 mM APG

    - containing 0.01 wt% PDMS-silica compound (Bartsch test). For comparison, in the absence

    of an antifoam, the initial foam volume was 180 ± 10 mL for AOT and 100 ± 10 mL for

    APG, and the foam was stable for the duration of this experiment [91]. Reproduced from [77].

    Figure 7. Dynamic surface tension of AOT and APG solutions measured by MBPM: 10 mM

    AOT and 0.6 mM APG were used in the foam tests (Figure 6). For comparison, the results for

    equal surfactant concentrations (i.e., 2.5 mM) are also shown [91], reproduced from [77].

    2.2.3 Effects of solid particles on liquid film stability

    Similar to surfactants, solid particles can adsorb to air-water interface and alter the

    interfacial properties [115-117]. Although effects of the hydrophobicity [118-122], size [123]

    and shape [118, 124] of solid particles on liquid film stability have been observed, the

    mechanisms that solid particles stabilize or destabilize liquid films are still not well

    quantified. It has also been found that solid particles influence the interfacial properties of

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    17

    liquid films and bubbles not only under equilibrium conditions but also under dynamic

    conditions (such as the bubble coalescence process) [116]. Despite the complicated effects of

    solid particles on the stability of liquid films, here we review the theory of maximum

    capillary pressure induced by solid particles to stabilize the liquid film and the bridging-

    dewetting mechanism of liquid film rupture by solid particles, respectively, to explain how

    solid particles can stabilize and destabilize the liquid film. We also discuss the effects of solid

    particles on foam drainage in Section 2.4 from a macroscopic perspective.

    2.2.3.1 Minimum energy to remove particle from interface, removeG and maximum capillary

    pressure for the onset of coalescence, maxcP

    To stabilize the liquid film, solid particles are required to stay at the air-water interface.

    The stability of a solid particle at the air-liquid interface is quantified by the minimum energy

    required to remove the particle from the air-water interface, removeG as follows [116, 125,

    126]:

    22 1 cosremoveG R (7)

    where R is the radius of the solid particle, is the surface tension of the air-water interface

    and is the contact angle of the liquid phase on a solid particle (the particle hydrophobicity).

    Sign “+” refers to particle removal into the gas phase while sign “-” refers to the removal of

    the particle into the liquid phase. Eq. (7) indicates that the highest stability of solid particle at

    the air-water interface occurs when the contact angle equals 90°.

    Bubble coalescence occurs only if the two interfaces of the liquid film contact. Suppose

    the solid particles form a single layer between the two bubbles, Kaptay [126] derived

    theoretical equations for the maximum capillary pressure for the onset of coalescence, maxcP ,

    inspired from the work of [127] as follows (Figure 8):

    max 2 coscP pR

    (8)

    where p is a parameter describes the effect of the packing and interfacial coverage by the

    particles on the capillary pressure. Eq. (8) indicates that the liquid film will be stabilized by

    the single layer of particles if the contact angle is lower than 90° ( maxcP is positive). On the

    contrary, the liquid film will be destabilized if the contact angle is larger than 90° ( maxcP is

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    18

    negative). Both Eqs (7) and (8) should be considered when the effect of solid particles on

    film stability is analyzed. Based on the combined analysis of Eqs (7) and (8), the optimum

    contact angle for the stabilization of liquid film by a single layer of particles is around 70°

    [126].

    Figure 8. Schematic of the packed single layer of spherical particles at the air-water-air

    interfaces, reprinted from [126].

    It is noted that the experimentally measured maximum capillary pressure for the solid-

    stabilized emulsions is just 0.25 of the calculated value based on Eq. (8) [128]. This

    discrepancy was explained by a sharp decrease of the capillary component of the film

    elasticity in the emulsions and by the defects of the particle layer packing.

    2.2.3.2 Bridging-Dewetting mechanism of liquid film rupture by solid particle

    Solid particles can act as antifoam to destroy the liquid films. The bridging-dewetting

    mechanism was established to explain the antifoam action of solid particles [76, 77, 79, 107,

    109, 118, 129]. First, the solid particle contacts with the two opposite interfaces of the liquid

    film during the film thinning and form a solid bridge between them. Then the particle surface

    is dewetted, and the three-phase contact lines come in direct contact with each other, which

    makes liquid film rupture (Figure 9). Evidently, for solid spheres the film rupture happens

    when the contact angle is larger than 90°, which is consistent with the physical meaning of

    Eq. (8).

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    19

    Figure 9. Schematic presentation of the bridging of liquid film by a solid particle, reprinted

    from [129]. For spherical particle, the film rupture happens when the contact angle is larger

    than 90°.

    2.3 Foam drainage for aqueous foams

    Foam drainage is the passage of liquid through a foam. Foam drainage is crucially

    important for foam stability. There are three different mechanisms governing the lifetime of a

    foam: (i) foam drainage caused by gravity, (ii) coarsening caused by the transfer of gas

    between bubbles induced by the capillary pressure differences, and (iii) bubble coalescence

    caused by the rupture of liquid films between neighboring bubbles [2]. The coarsening

    process cannot directly affect the foam lifetime. However, it has been found that the

    coarsening and foam drainage are strongly coupled since the coarsening rate is sensitive to

    the liquid fraction that is controlled by foam drainage, and conversely, foam drainage is

    strongly dependent on bubble size, which is controlled by coarsening [2, 130-132]. Therefore,

    coarsening can influence the foam lifetime via its effects on foam drainage. Among these

    three processes, foam drainage determines the liquid fraction of a foam, which is a key

    parameter for both coarsening and bubble coalescence [130, 133-145]. Different types of

    drainage regime (i.e., forced, free and pulsed drainage) have been observed [2, 9, 146-148]. A

    forced drainage experiment consists of pouring the foaming solution at constant flow rate into

    a foam that has already been allowed to drain freely [2, 9, 149, 150]. The theories developed

    for the foam drainage of aqueous foams are discussed below.

    During foam drainage, the liquid is confined in a network of channels or Plateau

    borders, which join at a node in four. Therefore, modeling foam drainage primarily focuses

    on the liquid flow in the two foam structures. Some researchers have also considered the

    contribution of the liquid film to the foam drainage [151, 152]. However, the liquid films

    have not been found to significantly contribute to the drainage process because of the small

    amount of liquid contained in films relative to that in the Plateau borders and nodes [151].

    Studies on foam drainage can be categorized into microscopic and macroscopic

    investigations [151]. The former refers to studies at the scale of a single Plateau border,

    whereas the latter refers to studies at the scale of at least several bubbles. The first study on

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    20

    the microscopic modeling of foam drainage only considers the contribution of Plateau

    borders to the foam drainage [153, 154]. The drainage rate is thought to be dependent upon

    the mobility of the Plateau border wall or the surface shear viscosity. A Plateau border-

    dominated approach to studying foam drainage was developed by subsequent researchers,

    who expanded on the initial microscopic modeling attempt [155, 156]. Nguyen improved on

    the work of [153] and provided a numerical solution for the liquid flow velocity in a single

    Plateau border as a function of surface viscosity [157]. The Plateau border-dominated

    approach was challenged when the contribution of nodes to the foam drainage was

    recognized, and mobile Plateau border walls were assumed [158]. Since then, the standard

    foam drainage equation [153] has been modified [159], and foam drainage models that

    consider viscous losses from both the Plateau borders and nodes have been proposed [146,

    160].

    On the macroscopic level, foam drainage has been analogous to the liquid flow

    through a porous medium [2, 161]. However, two key differences between liquid drainage

    through a foam and that through a porous medium must always be considered. First, in a

    foam, the size of the network (i.e., Plateau borders and nodes) through which the liquid flows

    is not fixed but is actually dependent on the flow itself: The bubbles can move apart to allow

    liquid to pass and then move back, and the interfaces are not completely rigid but are instead

    partially mobile in a foam, depending on the interfacial properties, such as surface shear

    viscosity. It has been demonstrated that the Darcy law that describes the fluid flow through a

    porous medium at low Reynolds numbers is also applicable to aqueous foams [133, 158, 161-

    163]. An alternative way to study foam drainage on the macroscopic level is to adopt

    dimensional analysis to compare the existing foam drainage data in a consistent manner and

    thus simplify the analysis [162]. One of the applications of foam drainage study is to

    calculate the liquid profile in pneumatic foam, which is a key factor to understand the liquid

    transport in a foam column. Different theories have been proposed to predict the liquid

    profile [164, 165]. However, further experiments are needed to validate these theories.

    2.4 Foam drainage in the presence of solid particles

    Despite the progress made in understanding aqueous foam drainage [2, 9], foam

    drainage in the presence of solid particles remains poorly understood. The study of the

    drainage of three-phase foams or froths is crucially important for many industrial applications.

    For example, in froth flotation, the wash water is commonly applied to the froth phase to

    flush the entrained gangue particles out of the froth and consequently increase the grade of

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    21

    hydrophobic particles. The reports on foam drainage in the presence of solid particles are

    reviewed here.

    Theories of foam drainage for aqueous foams have established benchmarks for

    studying three-phase foams in which solid particles are present. Logically, the following

    questions have been raised regarding foam drainage in the presence of solid particles: Can the

    foam drainage equations for aqueous foams apply to foams with solid particles? How does

    the presence of solid particles influence foam drainage?

    Not surprisingly, in the past 15 years, many studies on three-phase foams have focused

    on the transport or motion of solid particles in the flotation froth [166-168] or surfactant

    foams [169-172] because of their paramount importance for industrial applications. The solid

    particles in flotation froths have been divided into attached particles (i.e., hydrophobic

    particles), which follow the bubbles, and unattached particles (i.e., both hydrophobic and

    hydrophilic particles), which mainly follow the liquid [166]. On the one hand, attached

    hydrophobic particles adsorb to the air-water interface and act as a barrier to prevent bubble

    coalescence and impede the coarsening process [115, 173-175]. However, their effects on

    foam drainage remain unknown. On the other hand, studies on foam drainage in the presence

    of solid particles have primarily focused on the unattached hydrophilic particles and

    nanoparticles [167, 170-172, 176-183]. Nevertheless, these studies contribute to our

    understanding of drainage behaviors in the presence of solid particles. For example, by the

    scaling behavior (power law) between the drainage velocity and the imposed flow rate in

    forced drainage experiments [178], the presence of nanoparticles has been found to induce a

    foam drainage transition from a node-dominated regime to a Plateau border-dominated

    regime. Moreover, atypical results such as large foam permeability exponents and prefactors

    which have not been obtained in aqueous foams, have been recorded in three-phase foams

    [176]. It should be noted that these observations in three-phase foams cannot be explained

    simply based on the theories developed for aqueous foams. For example, the foam regime

    transition in aqueous foams is usually caused by a change in the surface viscosity or interface

    mobility [184, 185]. However, it is difficult to make the same claim in foams containing

    nanoparticles because the presence of hydrophilic solid particles can change the interfacial

    properties to only a small extent [178].

    To explain the foam drainage behaviors in the presence of hydrophilic particles, several

    mechanisms have been proposed, such as rheology of the powder suspension and clogging in

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    22

    the confined regions of the Plateau borders [171, 176, 179, 181-183]. Among them, one of

    the important parameters controlling the drainage behaviors of foams with hydrophilic

    particles is the confinement parameter, /p cd d , which relates the size of the particle, pd ,

    to the maximum diameter of the circle inscribed in the Plateau border cross-section, cd

    (Figure 10). The hydrophilic particles are trapped in the network of aqueous foams either by

    the mechanism of collective trapping – jamming – of the suspension for 1 [182], or by the

    individual capture of the particles by the foam constrictions for 1 [172, 179, 181, 183].

    The other important parameter that accounts for the effect of hydrophilic particles on foam

    drainage is the volume fraction of particles in the suspension, p . It has been found that the

    foam drainage velocity is reduced when p is sufficiently high even if 1 [182].

    Figure 10. Network of aqueous foams containing hydrophilic solid particles. Particles

    suspended in interstitial liquid can be either freely transported or trapped, depending on the

    confinement parameter, , that compares the particle size with the size of passage through

    those constrictions, cd , reprinted from [186].

    Very recently, a systematic study on the joint effects of the volume fraction of particles

    in the suspension, p and the confinement parameter on the foam drainage in the presence of

    hydrophilic particles, was reported [186]. In this study, all the regimes and transitions of

    reduced foam drainage velocity induced by the presence of hydrophilic particles are

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    23

    identified experimentally based on two parameters, p and . Particularly, the authors

    identified the particle inclusion/exclusion transition, which refers to the transition from full

    inclusion to full exclusion of solid particles with respect to the foam network. This transition

    makes the foam drainage velocity evolves from its minimal value (fully included particles) to

    its maximal one (fully excluded particles), due to the decrease in the fraction of particle

    surface area in contact with the foam network and consequently the reduction of the drag

    experienced by the draining liquid. Different regimes are divided by two particular values of

    the confinement parameter, c and * , that correspond to the end of the flowing suspension

    regime and the minimal foam drainage velocity, respectively. Specifically, five regimes are

    identified for the low and moderate particle volume fractions, i.e., 0.40p (Figure 11a):

    (i) Flowing suspension regime ( c ), where the drainage velocity is almost constant (for

    0.16p ) or just decreases slightly with ;

    (ii) Capture transition regime ( *c ), where the drainage velocity decreases

    significantly (for 0.40p );

    (iii) Maximum frictional drag regime ( * ), where the drainage velocity becomes minimal;

    (iv) Particle exclusion transition regime ( * 10 ), where the drainage velocity increases

    to the values that are close to the velocities in the first regime;

    (v) Minimal frictional drag regime ( 10 ), where the increase in the drainage velocity is

    much less pronounced than that of the previous regime.

    For 0.40p , the foam drainage stops in the range of moderate values of , but starts again

    for higher values of (Figure 11a). The authors also identified region of the jammed state

    by giving the following equations for 1 and * , respectively (Figure 11b):

    21 / 3 0.1

    0

    pack

    p

    pack

    p

    ( 1 ) (9)

    3

    6 0.72 11

    pack

    p

    C

    C

    ( * ) (10)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    24

    where is the gas fraction of the foam ( 0.90 in [186]), packp is the critical particle

    volume fraction beyond which the particle packing occurs [182], 0packp represents the

    value of packp for unconfined conditions and C is expressed as

    0.27 2.751 0.57 1 / 0.27 1 3.17 1C by numerical simulations of foam

    structures [172].

    Figure 11. (a) Reduced viscous drag (inverse of the reduced drainage velocity, 0/V V V ,

    defined as ratio of foam drainage velocities with particles and without particles) of foamy

    suspensions as a function of the confinement parameter , for several particle volume

    fractions. c and * , correspond to the end of the flowing suspension regime and the

    minimal foam drainage velocity, respectively. (b) Diagram of the reduced particle volume

    fraction as a function of the confinement parameter in the foam network. The jammed state

    domain is deduced from experiments. The red and black lines correspond to Eqs. (9) and (10),

    respectively. Reprinted from [186].

    Recognizing the dependence of foam drainage velocity on p and is significant not

    only for understanding the foam drainage with hydrophilic particles but also for the ones with

    unattached hydrophobic particles. The works from the Pitois group [186] could help the

    understanding of the complex froth behaviors in froth flotation.

    The following gaps have been identified by the current knowledge of foam drainage in

    the presence of solid particles:

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    25

    The effects of attached hydrophobic particles on the foam drainage remain unclear.

    The effects of solid particles on the stress state of the gas-liquid interface of foams

    remain poorly understood.

    It is unclear whether the foam drainage equations for aqueous foams can be applied to

    three-phase foams.

    3. FOAM COLUMN KINETICS

    Foam column kinetics describes the transient behaviors of foams, including the growth,

    drainage, and collapse of foams. Foam column kinetics is crucially important for many

    industrial applications. For example, the precise control over the froth phase becomes

    particularly important in froth flotation because of the strong dependence of its performance

    (product grade or selectivity and recovery) on the froth stability [11, 17, 187]. Foam column

    kinetics has been used to predict the froth rising velocity, which can be linked to a key

    parameter in froth modeling and plant operation, that is, the fraction of bubbles bursting on

    the top surface of the froth [188-190]. Foam column kinetics has also been applied in glass-

    melting furnaces [191-194]. We critically review the models for equilibrium foam height and

    foam growth. Then, we summarize the models for foam column kinetics that are available in

    the literature.

    3.1 Models of equilibrium foam height

    3.1.1 Hrma et al.’s model

    Bikerman introduced the concept of “unit of foaminess”, (s) as follows [6, 7]:

    max

    g

    H

    j (11)

    where max ( )H m is the equilibrium foam height under the superficial gas velocity,

    3 2( / ) ( ) / ( )gj m s Q m A m with volumetric gas flow rate Q and column cross-section area A .

    Despite its wide application to represent the foam or froth stability [188-190, 192, 195, 196],

    this expression is limited by its dependence on the gas flow rate [192, 197]. It is to note that

    an expression of Bikerman’s unit of foaminess, , regarding relevant physical quantities and

    the average bubble volume has been proposed [198]:

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    26

    31/2 1/2

    02

    1

    0.16H

    b

    f

    Vg

    (12)

    where and are liquid density and viscosity, is the surface tension, bV is the average

    bubble volume, g is the gravitational acceleration, f is a numerical factor that characterizes

    the boundary conditions (cross-section and surface mobility) for the flow of liquid through a

    Plateau border ( 49f for flow through a Plateau border under no-slip boundary conditions,

    i.e., the Poiseuille flow, and its value is reduced in the case of slip boundaries [151, 199]),

    and 0 and H are the liquid fraction at the bottom and the top of the foam column,

    respectively. However, the authors also pointed out that the product of the flow factor f and

    the liquid fraction H complicates the interpretation of any measured value for in

    Bikerman’s test [198].

    An equation has been proposed to express the equilibrium foam height as a function

    of the effective average bubble radius, r ; the critical superficial gas velocity, cj , beyond

    which the foam will grow without limit; the minimum superficial gas velocity required to

    generate foam, mj ; and the superficial gas velocity, j , as follows [192]:

    max 2 / /c m c mH r j j j j j j (13)

    Equation (13) is subjected to three conditions, namely, (i) m cj j j , (ii) the liquid film

    ruptures as soon as the critical film thickness is reached, and (iii) foam height is proportional

    to the gas phase flux. Only when cj j and m cj j , Eq. (13) can be expressed as

    max / 2 / mH r j j , where 2 / mr j such that a liquid with a low gas velocity threshold is

    more “foamable”.

    Hrma’s model successfully explains the experimental data qualitatively and gives

    useful insights into the mechanism of foam formation and stability [194]. However, the

    model cannot predict the equilibrium foam height due to the lack of expressions for cj and

    mj . Consequently, it is impossible to validate the model by comparing the model with

    experimental data.

    3.1.2 Hartland et al.’s model

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    27

    The first attempt to predict the equilibrium foam height dates back to 1974 [200]. In

    this model, the foam height, maxH , is related to the critical liquid film thickness, , the liquid

    density, , viscosity, , surface tension, , gas velocity, v , and bubble diameter, d . It is

    expressed by the following equation:

    5/4

    max

    1/4 7/4

    0.55 vdH

    d g

    (14)

    The authors also compared their model with experimental results by plotting against

    5/7 1/7 4/79/7

    max/v d g H as shown in Figure 12 with the following equation as the

    theoretical line that comes from Eq. (14):

    5/7 9/7

    1/7 4/7 4/7

    max

    0.71 v d

    g H

    (15)

    Because the regression line with a slope 0.62 is close to the theoretical line with a slope 0.71,

    it is claimed that little is lost by the assumption of the immobile interface.

    Figure 12. Comparison of experimental results with Hartland’s model, replotted from [200].

    The model successfully predicts that the foam height increases with the viscosity, gas

    velocity and bubble size increase but decrease as the density, surface tension and film

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    28

    thickness increase. However, the model assumes that the drainage of the films is represented

    by the axisymmetric drainage of liquid from between two flat discs, for which the Reynolds

    equation is applied. This assumption can be limited because the gas-liquid interfaces are

    mobile or partially mobile [146, 158, 160]. The assumption that the liquid film rupture occurs

    when the critical film thickness is reached is also problematic. It has been found that liquid

    films do not necessarily burst when their critical thickness is reached. Instead, the rupture of a

    liquid film occurs due to two independent and consecutive stages [191]. The first stage is the

    thinning of the liquid film due to drainage, and the second stage is the tear of the liquid film

    due to random molecular collisions [201] and fluctuations [143]. The characteristic time of

    liquid film rupture, f is the sum of two characteristic times: (i) the characteristic time of

    drainage, d , and (ii) the lifetime of the critically thin film, c , giving [191, 192]:

    f d c (16)

    Only when the foam is evanescent, that is, bubbles burst as soon as their films reach the

    critical thickness, is 0c . Even if the evanescent foam assumption is correct, in a real foam,

    the films are unlikely to burst at a single exerted pressure or critical thickness but will instead

    exhibit a distribution of bursting probabilities [202]. Also, the model does not consider the

    coupling between liquid film drainage and foam drainage (drainage along Plateau borders),

    whereby the pressure exerted on the film varies with position in the foam.

    3.1.3 Pilon et al.’s models

    Dimensional analysis has been applied to identify two dimensionless parameters, 1

    and 2 that include all physical variables relating to the equilibrium foam height [194]:

    2

    1

    Re

    m

    gr

    j j Fr

    (17)

    max max2

    mH j j HCa

    r r

    (18)

    A power-law type relation has been assumed between the two dimensionless parameters, i.e.:

    max Ren

    HCa K

    r Fr

    (19)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    29

    where K and n are constant parameters determined from experimental data. Re , Fr , and Ca

    are the Reynolds, Froude, and Capillary numbers, respectively, which are defined as:

    Re

    mj j r

    (20)

    2

    mj jFr

    gr

    (21)

    mj jCa

    (22)

    The authors found Eq. (19) can fit experimental data over a wide range of thermophysical

    properties with K = 2905 and n = -1.80 with a correlation coefficient R2 = 0.95, as shown in

    Figure 13.

    Figure 13. Correlation of dimensionless parameters 2 vs 1 , reprinted from [194] with

    experimental data from [203-208].

    Finally, the following relationship between the two dimensionless parameters has been

    determined:

    1.80

    max 2905

    Re

    H Fr

    r Ca

    (23)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    30

    It is noted that the relationship between 1 and (liquid volumetric fraction of foam) was

    established to compare the published forced drainage data [162, 209]. In these equations,

    is the liquid density, r is the average bubble radius, mj is the minimum superficial gas

    velocity for the onset of foaming [210], is the liquid viscosity, is the surface tension

    and g is the acceleration due to gravity.

    Because Eq. (23) neglects both bubble coalescence and inter-bubble gas diffusion and

    its application is limited to highly viscous liquids, a third dimensionless number to represent

    the Ostwald ripening that occurs in surfactant foams is needed to generalize this model to low

    viscosity fluids (e.g., aqueous surfactant solutions) [211]:

    3o

    m

    DS

    r j j

    (24)

    where D is the diffusion coefficient of the gas that diffuses across the liquid film from one

    bubble to the other, and oS is the Ostwald solubility coefficient. This dimensionless number

    can be interpreted as the ratio of the average contact time between bubbles in the foam,

    / mr j j and the characteristic permeation time, 2 / or DS .

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    31

    Figure 14. Correlation between dimensionless numbers 1.82 1 and 3 for aqueous foams

    made from different surfactant solutions and gases, reprinted from [211] with experimental

    data from [202, 211-217].

    The relationship among 1 , 2 and 3 is assumed to follow a power law as:

    2 1 3

    n mL (25)

    where L, n, and m are constants determined from experimental data. It has been found that

    both high and low viscosity fluids yield a value of n close to -1.80 via least squares fitting

    [194, 211-213]. The authors found that Eq. (25) can fit experimental data over a wide range

    of thermophysical properties with the parameters L = 118 and m = -0.96 with a coefficient of

    determination R2 = 0.95, as shown in Figure 14.

    Finally, the equilibrium foam height can be expressed in a dimensional form based on

    Eq. (25) [211]:

    1.760.8

    max 1.8 0.961.64118

    m

    o

    j jH

    r g DS

    (26)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    32

    Although dimensional analysis provides a simple and useful tool for the prediction of

    equilibrium foam height, there is no physical meaning for these fitting parameters in the

    models, such as K, L, m and n. Also, the theoretical model [210] developed for predicting the

    superficial gas velocity for the onset of foaming, mj is still needed to be validated by a

    consistent set of experimental data [218].

    3.1.4 Limitations of the models for equilibrium foam height

    All of the above models assume that there is no bubble coalescence on the top of the

    foam before the equilibrium foam height is reached. Based on this assumption, the foam will

    reach its equilibrium height as soon as the bubbles burst at the top of the foam. However, a

    gentler transition from not bursting to bursting is much more common in growing foam or

    froth [188-190, 202]. This relatively gentle transition indicates that rupture of liquid films or

    bubbles of foams do not occur at the same time but instead exhibit a distribution of bursting

    probabilities [202]. Therefore, predicting the change in foam height with time becomes more

    important than the equilibrium foam height.

    3.2 Models of foam growth and collapse

    Most modeling of foam stability has focused on foam collapse rather than growth [1,

    202, 219-222]. A mass balance equation for the bubbling gas in a foam column shown in

    Figure 15 has been proposed [1, 223]:

    2

    1

    11

    z

    z

    dzdGA A dz A

    dt dt (27)

    where G is the superficial gas velocity, A is the cross-sectional area, is the the volumetric

    liquid fraction of the foam. The first term on the right-hand side of Eq. (27) represents the gas

    in the foam, while the second term corresponds to the gas escaped from the collapsed foams.

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    33

    Figure 15. A pneumatic foam being generated by bubbling: no foam collapse (left panel) and

    collapse has occurred (right panel), reprinted from [1].

    Although it is widely accepted that the liquid fraction in the foam is spatially variant

    and decreases to an asymptote within a few cm up the foam column length [202, 224], Eq.

    (27) can be simplified to obtain the explicit expression for the foam growth rate

    /growth

    dH dt by introducing the average liquid fraction, , as:

    1

    g

    growth collapse

    jdH dH

    dt dt

    (28)

    where gj is the superficial gas velocity, H is the foam height, and / collapsedH dt is the

    foam collapse rate, which is dependent on the foam height (volume) [43]. The first term on

    the right-hand side of Eq. (28) represents the foam growth rate of stable foams (i.e., no

    bubble coalescence). The growth rate of transient foams, therefore, equals the growth rate of

    stable foam minus the foam collapse rate. In this section, the models of foam growth are

    categorized based on the treatment of the foam collapse rate: (i) the foam collapse rate is not

    considered; (ii) the foam collapse rate is expressed empirically; and (iii) the foam collapse

    rate is related to the critical film thickness or Plateau border size.

    3.2.1 Models without consideration of foam collapse

    An equation has been proposed to express the foam height as a function of the

    superficial gas velocity and average foam porosity (gas fraction), t [191]:

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    34

    gj

    H tt

    if t (29)

    where is the time for the foam height to reach its equilibrium value and

    0

    1,

    H

    t z t dzH

    (30)

    Since ( ) (1 )t , Eq. (29) would be the same of Eq. (28) as long as the foam collapse rate

    / 0collapse

    dH dt .

    The local foam porosity distribution has been assumed as a second-order polynomial:

    2

    0 1 2,z z

    z t a a aH t H t

    (31)

    where the coefficients 0a , 1a , and 2a are determined based on the boundary conditions at the

    top and bottom of the foam layer, as shown in Figure 16:

    10, ( )t t (32)

    0

    ,0

    z

    z t

    z

    (33)

    2,H t t (34)

    where 1 t is the porosity at the top of the foam as a function of time only, and 2 is the

    porosity at the bottom of the foam with a value of 0.74 for an ordered mono-disperse foam

    only [1, 225].

    Solving Eq. (31) with the boundary conditions, Eqs. (32)-(34) results in the following

    porosity profile of the foam:

    2

    1 2 1,z

    z t t tH t

    (35)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    35

    The average foam porosity t is then obtained by replacing ,z t with Eq. (35) in Eq.

    (30):

    1 22

    3

    tt

    (36)

    1 t has been treated as follows [191]: (i) constant porosity at the top, (ii) an exponential

    function of time, and (iii) obtained based on an approximate solution of the drainage equation.

    A similar model for the build-up and breakdown of foam in a glass melt has also been

    derived based on the hydrodynamics of the drainage of liquid lamellae [193].

    Figure 16. Schematic of a foam layer generated by bubbling and a coordinate system with

    notations where qpb is the volumetric velocity of liquid drainage, reprinted from [191].

    Although the authors obtained good agreement between the models predictions and

    experimental data for the case of low superficial gas velocity [191], they also pointed out that

    the model ignores some key physical processes taking place during foam formation, such as

    bubble coalescence on the top of the foam layer and Ostwald ripening between bubbles.

    Except for the missing physical processes, the authors did not provide any theoretical or

    experimental evidence for the assumption of the second-order polynomial for local foam

    porosity distribution in Eq. (31) as well as for the expression of average foam porosity in Eq.

    (36).

    3.2.2 Empirical equations for foam growth and collapse

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    36

    The half decay time ( 1/2t ) at which the foam column height is half of the original height,

    was found by expressing the foam collapse rate as [8]:

    /dH

    k tdt

    (37)

    Integrating Eq. (36), the expression for the change of foam height with time becomes:

    0 1/2/ ln / 0.5H H t t (38)

    where 0H is the original foam height, 1/2t is the half foam decay time, and is a constant.

    Because the relationship between H and t is known, 1/2t can be obtained by fitting the foam

    collapse data [223].

    An empirical equation for the growth of the froth in a non-overflowing column placed

    into a flotation cell has also been proposed [188-190]:

    /max 1 tH t H e (39)

    H is the fraction of air remaining in the froth at a given froth height H , which is a

    crucial parameter in froth modeling and plant operation involving froth flotation; it is

    expressed as:

    maxH H tdH t A A

    Hdt Q Q

    (40)

    where maxH is the equilibrium foam height; is the average bubble lifetime, which equals

    the unit of foaminess, , defined by Eq. (11); A is the column cross-sectional area; and Q is

    the gas flow rate to the flotation cell [189]. Eq. (40) assumes that the liquid fraction of the

    froth is zero and H is different from the gas fraction or foam porosity (when

    maxH t H , 0H but the gas fraction is not zero). The growth rate of the froth height

    can be deduced from Eq. (39) used with max gH j as follows:

    1g

    dHj H

    dt (41)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    37

    Suppose that the liquid content is low, i.e., 0 , by comparing Eq. (41) with Eq. (28), we

    can see that the foam collapse rate is proportional to the foam height. Both Eq. (38) for foam

    decay and Eq. (39) for foam growth are empirical equations based on experimental results

    without any fundamental basis.

    3.2.3 Foam growth model relating to critical Plateau border size

    Bubble burst rate on the top of the foam layer has been correlated with the critical

    Plateau border size (area), critA with assumption that there is no internal bubble coalescence,

    as follows [202]:

    1 crit

    Burst

    dHk A

    dt

    (42)

    where

    13 PB

    gk

    C

    (43)

    is the liquid density, g is the acceleration due to gravity, PBC is the Plateau border drag

    coefficient, which can be determined by forced drainage experiments and is the liquid

    viscosity.

    The foam initially grows at a rate of gj (superficial gas velocity) with zero liquid

    fraction assumption ( 0 ). When critical Plateau border size is reached, i.e., a sufficient

    force is exerted to overcome the disjoining pressure in liquid film, bubbles (liquid films) at

    the top of foam layer starts to burst and the foam either reaches an equilibrium height and

    stop growing, or it continues to grow with a reduced growth rate,

    1 crit g

    Final

    dHk A j

    dt

    (44)

    depending on the gas velocity. The critical Plateau border area can be obtained based on the

    foam growth curve if the gas flow rate is sufficiently high:

    1

    1crit

    Initial Final

    dH dHA

    k dt dt

    (45)

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    38

    The comparison between theoretical and experimental foam growth data for the foam created

    from a 4 ml/l Teepol solution is shown in Figure 17.

    Figure 17. Comparison between theoretical (lines) and experimental (symbols) foam growth

    data for foam created from a 4 ml/l Teepol solution, replotted from [202].

    Although main fundamental physics governing the foam collapse process has been

    considered, a significant discrepancy between the simulated and experimental results for

    growing foams remains (Figure 17). This discrepancy has been explained as follows: “In a

    real foam, the films are unlikely to burst at a single exerted pressure, but rather are likely to

    exhibit a distribution in the bursting probabilities….This distribution in the film stabilities

    probably accounts for the more gentle transition from not bursting to bursting seen in the

    experiments” [202]. Also, the assumption that the bubbles start to burst when the critical

    Plateau border size is reached requires further experimental validation.

    3.3 Summary

    The equations to describe the equilibrium foam height and foam growth reported in the

    literature are summarized in Table 1.

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    39

    Table 1. Summary of equations for the equilibrium foam height and growth and collapse of

    foam and froth

    Source Equation Note

    Equations for equilibrium foam height

    [192] max 2 / /c m c mH r j j j j j j -

    [200]

    5/4

    max

    1/4 7/4

    0.55 vdH

    d g

    Rigid interface assumption

    [194]

    1.80

    max 2905

    Re

    H Fr

    r Ca

    For highly viscous fluids

    [211]

    1.760.8

    max 1.8 0.961.64118

    m

    o

    j jH

    r g DS

    Applied for surfactant solutions

    Equations for foam or froth growth and collapse

    [191] gj

    H tt

    No bubble coalescence

    [8] 0 1/2/ ln / 0.5H H t t Empirical equation

    [188] /max 1 tH t H e Empirical equation

    [202] 1 crit gFinal

    dHk A j

    dt

    After onset of bursting

    4. CONCLUDING REMARKS

    This review tackles the issue of foamability and foam stability from the perspective of

    different length scales of the foam: the air-water interfaces (surface tension and surface

    viscoelasticity), the liquid films (intermolecular forces, antifoam actions and solid particles

    on its stability), the bubbles (coalescence) and the foam (foam drainage and foam column

    kinetics).

    As for the foamability, it is widely accepted that foaming solutions with a low

    equilibrium surface tension, fast adsorption kinetics, an optimal surface viscoelasticity to

    restore the air-water interface after disturbances via the Gibbs-Marangoni effect and a strong

    and long-range repulsion between interfaces of a liquid film, have a good foamability.

    However, the measurements of the properties of the air-water interface and liquid film are not

  • ACC

    EPTE

    D M

    ANU

    SCR

    IPT

    ACCEPTED MANUSCRIPT

    40

    directly relevant to the foam formation because the conditions that we studied are different

    from those that occur during foam formation [2]. Many questions remain unanswered,

    including: What is the adsorption reference time for the foam formation? What is the

    frequency for the surface viscoelasticity that is relevant to the foam formation?

    As for the foam stability, it has been found that drainage of liquid films and foams,

    coarsening and coalescence are three main processes that determine the foam lifetime.

    However, the collective effects and interactions of the three processes make it very difficult

    to study the foam stability. Furthermore, the studies on intermediate scales, such as air-water

    interface and foam films, are inadequate to explain the mechanism of foam stability due to

    the