Вы находитесь на странице: 1из 11

Current Opinion in Colloid & Interface Science 20 (2015) 81–91

Contents lists available at ScienceDirect

Current Opinion in Colloid & Interface Science

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

Defoaming: Antifoams and mechanical methods


Peter R. Garrett
School of Chemical Engineering and Analytical Science, University of Manchester, Oxford Road, Manchester M13 9PL, UK

a r t i c l e i n f o a b s t r a c t

Article history: Recent progress in both the mode of action of antifoams and mechanical defoaming is reviewed. New insights
Received 29 December 2014 concern the simulation of the orientation of particles in interfaces and films, the role of dynamic surface effects
Accepted 2 March 2015 in antifoam action, antifoam action under micro-gravity, deactivation of oil/particle antifoams and antifoam
Available online 10 March 2015
action in hydrocarbon media. Progress in mechanical defoaming is mainly confined to new insights concerning
the use of ultrasound.
Keywords:
Antifoam mechanism
© 2015 Elsevier Ltd. All rights reserved.
Particles
Deactivation of oil/particle antifoams
Dynamic effects
Ultrasound

1. Introduction in academic circles. A recent foam conference for example, included


only one paper explicitly concerning defoaming [1] out of a total of
Defoaming is a surprisingly ubiquitous requirement. In the oil over 100 oral and poster presentations.
industry, for example, processes such as gas–crude oil separation and Despite this seeming lack of recent interest much progress has been
desulphurization of natural gas by bubbling through alkanolamine made in understanding the mode of action of antifoams over the past
solutions usually require defoaming, which often involves the use of thirty years. A detailed account of that progress can be found in a recent
chemical antifoams. Other industrial processes such as the jet dying of monograph [2⁎⁎] which is concerned with defoaming by both anti-
textiles, radioactive waste treatment, the kraft pulp process and fer- foams and mechanical means. Briefer accounts of defoaming by anti-
mentation using oxygen bubble columns are also beset by unwanted foams have also been published recently by Denkov et al. [3⁎] and
foam. Again treatment with antifoams is often used but mechanical Karakashev and Grozdanova [4]. Another recent review by Owen [5] is
means are sometimes employed in the case of bubble columns. The con- distinct in that it ignores much of the progress made over the past
trol of flotation froth in mineral processing can be adversely affected if twenty years. Earlier noteworthy reviews include those by Miller [6]
hydrophobed mineral particles too readily destabilize the froth. Forma- and Denkov [7⁎⁎].
tion of undesirable foam can also be an aspect of the use of certain for- Use of antifoams always implies contamination of the system to be
mulated products such as waterborne latex paints, and detergents for defoamed. Such contamination is, however, sometimes unacceptable.
use with machine washing of textiles and dishes. This problem usually Examples include preparation of pharmaceuticals by fermentation and
necessitates the addition of antifoams to the relevant formulations. downstream processing in petrochemical plants where catalysts may
However some products, such as shampoos and hand dishwashing liq- be poisoned by antifoam residues. In these situations recourse is some-
uids, are designed to produce copious amounts of foam during use. times made to defoaming by mechanical means, utilizing for example
These products must be formulated to minimize the antifoam effects either ultrasonic or various rotational devices. Progress in understand-
of the triglyceride soils which are intrinsic to their application. There ing the mode of action of such devices has however been rather limited.
are even medical applications of defoaming. These include the use of an- A complete review is to be found in the recent monograph [2⁎⁎].
tifoams to treat gastrointestinal gas, to eliminate any foam obscuring Developments in the understanding of the mode of action of anti-
the view of colonoscopy cameras and in filters for removal of foam bub- foams, which have been published over the past 5 years or so, are the
bles from the aspirated blood derived from surgery in order to permit main concern of this review. Earlier work, however, is also cited in
recirculation. Defoaming is therefore clearly widely relevant in many order to provide relevant context. Antifoams are defined here as parti-
contexts and does in fact present many interesting challenges to funda- cles, oils or mixtures of oils and particles which reduce either the
mental scientific understanding. This is however not always recognized foamability or foam stability of the liquid in which they are dispersed.
Some account of defoaming by mechanical means is also included
where however a dearth of recent relevant publications means a review
E-mail addresses: p.garrett@manchester.ac.uk, pgarcgar@btinternet.com. largely confined to ultrasonic defoaming.

http://dx.doi.org/10.1016/j.cocis.2015.03.007
1359-0294/© 2015 Elsevier Ltd. All rights reserved.
82 P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91

2. Defoaming by particles years ago. Dippenaar [8] investigated the rupture of single aqueous
films by orthorhombic hydrophobed galena particles using high-speed
2.1. Hydrophobic particles; theory cinematography. He attributed his finding that aqueous film rupture
by such particles occurs with contact angles of only (80 ± 8)° to the par-
It is now well known that finely divided particles, which are suffi- ticle shape. He concluded that such particles could adopt two orienta-
ciently hydrophobic, can produce antifoam effects when dispersed in tions at the air–water surface as that surface is pinned to edges as
aqueous surfactant solution [2⁎⁎,3⁎,8]. Whether these effects are appar- shown in Fig. 1a and b. According to Dippenaar [8] only the diagonal
ent is dependent upon the contact angle and the shape of the particles. orientation can give rise to foam film rupture by a mechanism analo-
In the case of spherical particles antifoam effects are only present if the gous to that for a sphere but in this case, for a particle of a square
contact angle (measured through the aqueous phase), θAWN ~90° [2⁎⁎, cross-section, contact angles must lie in the range 45° b θAW b 135°.
3⁎,8]. As the film drains the particles bridge the film, and dewet (now A problem with this argument concerns the assumption that only
often called the “bridging–dewetting mechanism” after Denkov [7⁎⁎]) two orientations of an orthorhombic particle at a planar air–water sur-
leaving a hole in the foam film which subsequently expands to cause face are possible. That is strictly only true if the aspect ratio ≫ 1. In the
film rupture. The essentials of this mechanism have been verified exper- case of the diagonal orientation the air–water surface must satisfy the
imentally in model experiments for spheres using high-speed cinema- contact angle against the two particle end surfaces perpendicular to
tography [8]. the plane of the paper as shown in Fig. 1c and d. If for example the
Joshi et al. [9⁎] present interesting observations concerning the anti- air–water surface remote from the particle is planar then the capillary
foam mechanism of hydrophobic fatty alcohol particles in aqueous sur- pressure across that surface must be zero. For equilibrium the capillary
factant solution. The fatty alcohols consist of a blend of C14 –C30 chain pressure in the segment of air–water surface against the particle ends
lengths. The particles appear to have a spherical or ellipsoidal geometry must also be zero. That is only possible if it forms a catenoid element,
and should therefore require contact angles N 90° in any given aqueous which must also contact everywhere the perpendicular surface of the
medium for antifoam effect. However in the absence of surfactant the particle at the relevant contact angle. In the case of a particle with an as-
particles possessed advancing and receding contact angles of 93° and pect ratio close to unity the contribution of the surface energy of the two
95° respectively and in the relevant aqueous surfactant solution 85° particle end surfaces, including the catenoid elements, to the total work
and 71° respectively. The effect of these particles on the stability of the of adhesion of the particle to the air–water surface must be significant.
film formed between two colliding bubbles was observed by video- This factor therefore greatly complicates calculation of the relative
microscopy. In these experiments the fatty alcohol particles were dis- work of adhesion and probabilities of the two orientations depicted in
persed in an aqueous concentrated ethoxylated alcohol solution con- Fig. 1a and b in the case of aspect ratios close to unity. It also suggests
taining a thickener and preservative — the resultant concoction often the possibility that yet other orientations are possible.
being described commercially as an aqueous or surfactant antifoam. Calculating the relative work of adhesion for orthorhombic particles
This concoction was then added to an aqueous solution of a second sur- as a function of contact angles has been made tractable by use of an
factant in order to achieve an antifoam effect for the latter. In this ar- iterative surface energy minimization technique described by Morris
rangement one bubble was allowed to equilibrate with the antifoam et al. [11–13,14⁎⁎]. The technique makes use of the Surface Evolver soft-
dispersion so that the fatty alcohol particles could attach to the relevant ware developed by Brakke [15]. A more detailed description of the use
air–water surface. The second bubble was then grown towards the first of this technique is given in the accompanying review by Morris et al.
to form a film between the two bubbles. The subsequent flow of liquid [16]. It permits calculation of the (surface) energy profile of a particle
from the film produced a surface tension gradient and effectively re- in an air–water surface or foam film as a function of the orientation, con-
moved the particles. However when that flow ceased, as the second tact angle and shape. The profile reveals the presence of energy minima,
bubble closely approached the first, a reverse flow occurred because consistent with stable orientations, where the depth of the minima
the surface tension of the air–water surfaces in the film was now higher gives an indication of their relative probabilities. This analysis reveals,
than that on the remaining surface of the first bubble. The resulting for example, four possible orientations for an orthorhombic particle
Marangoni effect dragged the antifoam particles back into the film bridging a foam film [14⁎⁎]. These are designated vertical, horizontal,
which then rapidly ruptured by a bridging mechanism. Joshi et al. [9⁎] rotated and diagonal and are shown in Fig. 2. The relative stabilities of
attribute the origin of the reverse flow to desorption of surfactant these orientations are determined by both the contact angle and the
from the fatty alcohol particles. However spreading from the shorter aspect ratio of the particle. This is exemplified in Table 1 for the aspect
chain length fatty acids in the blend used here could also contribute ratios corresponding to the only direct experimental observations of
(the equilibrium spreading pressure of, for example octadecanol – a the rupture of aqueous films by orthorhombic particles.
component of the blend used – will lower the surface tension of water Careful examination of the cinematographic film frames of Dippenaar
at the relevant temperature to some 20 mN m−1 below that of the solu- [8] reveals that the aspect ratio of the orthorhombic hydrophobed galena
tions of the second surfactant considered here [10]). Clearly a dynamic particle is about 1.4. With a contact angle in the range of 72–88°
process is occurring which probably means that dynamic contact angles (i.e. 80 ± 8°) the results of Table 1 would suggest that either both hori-
are relevant which could explain the apparent discrepancy between the zontal and diagonal orientations are likely to co-exist or that the diagonal
measured receding angles and the requirement that θAWN ~ 90°. orientation alone is to be present. Dippenaar [8] suggests that his obser-
There are in fact two main difficulties with generalizing the condi- vations are in fact consistent with the former possibility. However the
tion θAWN ~ 90° to all foamability measurements. The first concerns relevant cinematographic frames do suggest a reconfiguration to a rotat-
measurement of the contact angle under conditions directly relevant ed orientation immediately prior to rupture of the aqueous film. Morris
for those actually existing during foam generation. As the work of and Cilliers [13] have recently repeated these observations of Dippenaar
Joshi et al. [9⁎] suggests contact angle hysteresis and dynamic effects [8] with a similar hydrophobed galena particle of contact angle in the
due to the rate of surfactant transport relative to the rate of air–water range of 70–90° (i.e. 80 ± 10°) but of aspect ratio unity. As expected
surface generation can conspire to produce antifoam effects even from the calculations summarized in Table 1 this particle adopted a
though the receding contact angle b 90° [2⁎⁎]. However arguably a rotated orientation although with no apparent observations of either
more significant factor concerns particle geometry. The presence of horizontal or vertical orientations.
sharp edges and rugosities appears to confer antifoam effects at contact Morris et al. [11–13,14⁎⁎] have extended these surface energy min-
angles significantly b 90° [2⁎⁎]. imization calculations to estimate the critical capillary pressure required
Orthorhombic particles have received particular recent interest in to rupture an aqueous film containing a bridging orthorhombic particle.
this context following the work of Dippenaar [8] more than thirty In particular Morris and Cilliers [13] have compared such calculations
P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91 83

air

water
film
thins

a horizontal orientation

air
water
film
thins

b diagonal orientation; aqueous phase peels


off particle to form a
hole

planar air-water
catenoid air- surface hinging on
water surface an edge

air

water

c d

Fig. 1. Orientation of an orthorhombic particle in a foam film. (a) and (b) Horizontal orientation stabilizes a draining foam film whereas a diagonal orientation leads to foam film rupture
provided θAW N θp = 45o according to Dippenaar [8] for the case of a particle with a square cross-section and an aspect ratio ≫ 1 (i.e. a 2D view). (c). Mechanical equilibrium requires that
the capillary pressure over the whole air–water surface be zero which must mean a catenoid profile over the end surfaces of a diagonally oriented particle which become progressively
more significant as the aspect ratio decreases. (d). Vertical slice through centre of particle where the contact angle is satisfied and the air–water surface is concave, the orthogonal radius
of curvature must then be convex if the surface is to form a catenoid as shown in (c).

with experimental observations which supposedly repeat those of stabilization aqueous films of solutions of surface active materials
Dippenaar [8] but, as we have seen, using a hydrophobed galena particle against rupture. Application of this technique to the actual process of
of aspect ratio unity rather than 1.4. The calculations indicate that the film rupture by a bridging particle for comparison with the experiments
particle, placed in a rotated orientation but bridging the film, maintains of Morris and Cilliers [13] and Dippenaar [8] is nevertheless justified be-
that orientation as capillary pressure is increased and the film thins until cause the aqueous films in that context consisted of pure water — no
it ruptures. These calculations suggest that film rupture does not occur surfactant at all was present and therefore a positive contribution to
as a result of the film liquid peeling off the particle surface as indicated the disjoining force was absent. However the presence of significant
in Fig. 1b and apparently revealed for an orthorhombic particle by the positive disjoining forces due to the presence of surfactant is likely to
cinematographic observations of Dippenaar [8]. Morris and Cilliers suppress the process of foam film rupture by the indirect influence of
[13] rather suggest the particles induce rupture directly in foam films a bridging particle on the neighbouring film. After all it is that property
where “its highly distorted shape around the particle is what forces of surfactants which justifies the addition of antifoams at all! It probably
the opposite sides (of the film) together”. Unfortunately the experimen- means that in those circumstances foam film rupture would have to
tal observations were not able to convincingly verify this conclusion. occur by the mechanism originally suggested by Dippenaar [8] where
Moreover strict comparison with Dippenaar [8] was not possible the film liquid peels off the particle until the two three phase contact
because of selection of an aspect ratio of unity rather than 1.4. lines become coincident as shown in Fig. 1b. However as described by
The surface energy minimization technique used in these calcula- Garrett [2⁎⁎] application of that mechanism to particles of complex ge-
tions of the critical capillary pressure is not able to account for the pos- ometry requires knowledge of their orientations at air–water surfaces
itive contribution to the disjoining forces, which will generally facilitate and in foam films. The surface energy minimization technique may
84 P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91

a b

c d

Fig. 2. Possible stable orientations of an orthorhombic particle bridging a thin liquid film (of thickness ≪ dimensions of the particle). (a). Horizontal (H), (b). Diagonal (D), (c), Rotated (R),
(d). Vertical (V). Orientations calculated by a surface energy minimisation technique using the Surface Evolver software [14⁎⁎,15]. The catenoid surface suggested by Fig. 1c is clearly
revealed in the diagonal orientation (b).
Reprinted from Morris, G., Neethling, S.J., Cilliers, J.J. J. Colloid Interface Sci. 361, p373; copyright 2011, with permission from Elsevier.

therefore potentially yield valuable insights in that context especially [2⁎⁎]. Similarly addition of excess counterion to solutions of sodium do-
for particles of complex crystal habit. Such considerations are important decyl sulphate lowers the Krafft temperature to produce a crystalline
not only for rupture by particles of air–water–air foam films but also for precipitate. That precipitate also exhibits a finite contact angle [17]
the rupture by particles of oil–water–air pseudoemulsion films which against the saturated solution and appears to function by a bridging–
represents an important aspect of the mode of action of commercial an- dewetting mechanism as an antifoam [2⁎⁎,18].
tifoams for application in aqueous surfactant solution [2**]. In contrast to single chain surfactant molecules, exemplified by
carboxylate soaps and SDS, addition of counterions to double chain
2.2. Effect of precipitation of crystalline and mesophase particles on foam molecules tends to precipitate mesophase dispersions – often as
multi-walled vesicles of lamellar phase. The precipitates formed upon
Addition of counterions to aqueous solutions of anionic surfactants addition of Ca2 + ions to aqueous solutions of sodium alkyl benzene
can result in partial substitution of monomers or micelles by precipi- sulphonate [19,20] represent an example of this behaviour. Decreases
tates of either crystalline or mesophase particles. This change in physical in foamability of the solution accompany formation of the precipitate.
state of the relevant solution usually produces a marked reduction in The residual foam is however stable.
foamability. For example calcium soaps exhibit extremely low solubil- The defoaming effect accompanying addition of Ca2+ ions to solu-
ities so that they may be precipitated from aqueous mixed solutions of tions of sodium alkyl benzene sulphonate can find practical application
sodium soaps and other surfactants. The relevant precipitates are crys- in detergency by the elimination of residual foam during rinsing with
talline and can act as antifoams for solutions of the latter. Finite contact hard water. Removal of the lamellar phase precipitate by filtration
angles suggest that they function by a bridging–dewetting mechanism does not however restore the foamability [19,20] which means that
the precipitate does not function as an antifoam. This is probably a con-
Table 1 sequence of the structure of the multi-walled vesicles which are formed
Stable orientations of an orthorhombic particle in a thin aqueous film as a function of contact upon precipitation of lamellar phase. The outer layer of the vesicle
angle and aspect ratio calculated by surface energy minimization. H = Horizontal, D = formed in an aqueous environment has to be covered with surfactant
Diagonal, R = Rotated, and V = Vertical (see Fig. 2). From Morris et al. [14⁎⁎]. head groups if it is composed of bilayers. That surface is therefore hydro-
Aspect ratio Contact angle, θAW/degrees philic and is extremely unlikely to posses the contact angle of 90° re-
quired for antifoam action by a sphere — indeed it is likely to be near-
45 60 65 70 75 80 90
zero.
1.0 H/V H/V H/V H/R/V H/R/V R R Measurements of the dynamic surface tensions indicate that the
1.36 H/V H/V H/D/V H/D H/D D D
reduction of foamability simply accompanies replacement of labile
P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91 85

micelles with non-labile mesophase particles which are larger (with the likely antifoam effects. We should however emphasize that these
smaller diffusion coefficients and slower breakdown kinetics) [2⁎⁎]. generalisations are drawn from relatively few examples. There is clearly
This is illustrated by a comparison of the effect of addition of Ca2 + a need to extend these observations to other systems and include more
ions on the foam and dynamic surface tension behaviour of a micellar detailed characterization of the nature of the precipitates.
solution of sodium dodecyl 4-phenyl sulphonate shown in Fig. 3. The
figure reveals a marked increase in dynamic surface tensions with the 3. Antifoam effects by neat oils
onset of precipitation of the mesophase. In turn this suggests slow trans-
port of surfactant to air–water surfaces, which will contribute to low ad- 3.1. Oil drops in aqueous foam films; definitions and theory
sorption levels, low positive contributions to disjoining forces and high
capillary pressures in Plateau borders. These factors will all conspire to There is some confusion, even in recent literature, about the exact
produce lower foam film stability during foam generation. However criteria for effective defoaming action by liquid antifoams. Here we
after foam generation has ceased surfactant can slowly adsorb to ame- give an abridged outline using an antifoam oil–water combination as
liorate these factors ensuring the stability of any foam which survives an example — with some important exceptions the same basic criteria
the process of generation. apply to other fluid–fluid systems. Full details are to be found elsewhere
All of this suggests the generalisation that precipitation of crystalline (see for example reference [2⁎⁎]).
or amorphous crystalline particles by addition of counterions to aque- The first requirement for effective antifoam action is that the solubility
ous solutions of anionic surfactants may produce antifoam effects limit of the antifoam liquid in the foaming liquid should be exceeded
whereas precipitation of mesophase is unlikely to do so. However pre- (there are however some exceptions to this generalisation — see for
cipitation in either case means that labile entities are replaced with example Section 3.3 and [2⁎⁎]). The second requirement is that the oil
non-labile entities. In turn this means that the rate of transport of sur- be dispersed as drops, sufficiently small and numerous to ensure
factant to the rapidly expanding surfaces present during foam formation presence in foam films. The third requirement is that the oil drops
will be reduced. This factor will always contribute to diminished spontaneously emerge into the air–water surface. This means that the
foamabilities. In the case of crystalline precipitates it will even enhance entry coefficient should be positive — a thermodynamic criterion.
However the aqueous “pseudoemulsion” film separating an oil drop
from the air is usually metastable and must first be ruptured by for
Clear micellar
example, particles of a suitable wettability, if emergence of the drop is
region
Volume air in to occur [2⁎⁎,3⁎].
foam/cm3 Once the oil has emerged into the air–water surface there are
essentially three different phenomena which can occur. In the simplest
100
Lamellar phase
case the oil can form lenses without in any way contaminating the air–
precipitate region water surface — so-called partial wetting (using the nomenclature of
80
Brochard-Wyatt et al. [21]). Alternatively the oil may spread to contami-
60 nate the air–water surface to form an unstable film which subsequently
breaks up to leave oil lenses in equilibrium with an air–water surface con-
40 taminated with oil — so-called pseudo-partial wetting [21]. Finally the oil
may spread to form a duplex film with no lenses. This film is so thick that
20
the oil–air and oil–water surface tensions of the film have the values of
0 the bulk phase — so-called complete wetting [21]. In each of these three
-4 -3 -2 cases the equilibrium spreading pressure, Se, must satisfy Eq. (1) where
2+ we can only have [22]
Log10 ([Ca ]/M)
a
e e
−2σ OW ≤ S ≤0 ð1Þ

Precipitation where σeOW is the oil–water surface tension. Eq. (1) means that (at equilib-
boundary rium) Se can only be either zero or negative. However in the case of both
pseudo-partial and complete wetting the oil must first spread over the
75 air–water surface before equilibrium is attained. In these cases a transient
“initial” positive spreading pressure prevails, defined by non-equilibrium
65 values of the relevant surface tensions.
An emerged oil drop, for which Eq. (1) is satisfied, may actually
55 bridge across both sides of an aqueous film. It was shown more than
thirty years ago [23] that if such an oil bridge forms in a plane parallel
45 foam film then it will be unstable if the angle θ* formed between the
air–water and oil–water surfaces is N π/2. That angle, θ*, is determined
35 by Neumann's triangle of surface forces so that we must have

25 
π=2 b θ ≤π ð2Þ
-4 -3.5 -3 -2.5 -2
Log10 ([Ca2+]/M)
b where θ* = π represents complete wetting of the aqueous film by a
duplex oil film. It follows from Neumann's triangle that we may rewrite
Fig. 3. Effect of precipitation of lamellar phase by addition of Ca2+ to a solution of 2 mM the Eq. (2) as
sodium 4-phenyl benzene sulphonate in 17 mM NaCl at 25 °C (a). Foamability by Bartsch
method (cylinder shaking by hand — see [2⁎⁎] for details). (b). Dynamic surface tensions e e
at surface ages; ▲, 0.1 s; ■, 1 s; ♦, 10 s. [19,20]. 0 b B ≤ 2σ OW σ AW ð3Þ
86 P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91

where B is the bridging coefficient defined by however be overcome in the case of drops trapped in Plateau borders
by the high capillary pressures attained as foam drains, producing
e 2 e 2 e 2
B ¼ σ AW þ σ OW − σ AO ð4Þ slow declines in the overall stability of foam [2⁎⁎]. Whether the resulting
foam film rupture, occurring in or at the edge of the Plateau border, is de-
where σeAW and σeAO are respectively the equilibrium air–water and air– termined by the bridging coefficient is however not yet established [2⁎⁎].
oil surface tensions. Here we should note that if the oil exhibits a posi- More commonly, however, rupture of pseudoemulsion films is achieved
tive initial spreading coefficient, then mechanical equilibrium does not in the case of commercial antifoams by adding hydrophobed particles
exist between the three surface tensions while that transient condition to the oil. Those particles supposedly rupture the pseudoemulsion
prevails and Neumann's triangle is violated. The bridging coefficient is films by a bridging mechanism similar to that by which hydrophobic
then meaningless with values higher than the upper limit of Eq. (3). particles rupture aqueous foam films (for full accounts of this see for
This restriction has not always been recognized [7⁎⁎,24–27]. example [2⁎⁎,3⁎,7⁎⁎]). Other possibilities are sometimes claimed to
There is much experimental evidence that the antifoam action of an exist however. Shu et al. [27] for example show that addition of acid to
oil in an aqueous surfactant solution is largely confined to situations an aqueous solution of sodium oleate precipitates oleic acid drops and
where the bridging coefficient is positive [2⁎⁎,3⁎]. It should be stressed causes reduction of foamability and foam stability. They claim that this
also that Eqs. (2) and (3) account for antifoam effects by oils exhibiting is due to a bridging antifoam effect by the oleic acid drops. The relevant
all types of possible wetting behaviour — complete wetting, pseudo- pseudoemulsion film is supposed intrinsically unstable at low pH due
partial wetting and partial wetting. Positive initial spreading coefficients to low levels of adsorption of charged species at the oleic acid–water in-
then simply imply a transient process of rapid supply of oil lenses or du- terface. The resulting absence of electrostatic effects would therefore
plex film to foam films so that unstable liquid bridges can be formed. imply that particles are unnecessary in this case. However Shu et al.
Denkov et al. [3⁎,7⁎⁎] have proposed that unstable bridging drops [27] have not shown that removal of the oleic acid drops causes an in-
will lead to foam film rupture by either a bridging–stretching mechanism crease in foamability and therefore actual antifoam behaviour in this
or by a bridging–dewetting mechanism. In the bridging–stretching case is not proven. The effect could be due entirely to transport effects
mechanism the drop is stretched until the two oil–air surfaces touch deriving from the change in concentration and nature of the remaining
whereupon a hole forms and the drop is to torn apart to rupture the dissolved surface active species after the addition of the acid. We should
film. By contrast the bridging–dewetting mechanism is supposed to note also that Shu et al. [27] present positive values of both spreading
occur if the aqueous film peels off the drop in a manner similar to that and bridging coefficients in violation of Neumann's triangle so that the
for a solid spherical hydrophobic particle [3⁎]. It would seem possible quoted bridging coefficients are meaningless.
that this mechanism could prevail if the viscosity of the oil is high It has recently been shown that the defoaming of crude oil by
enough. Recently Denkov et al. [3⁎] claim to have observed bridging– suitable neat liquid antifoams appears to involve a bridging mechanism
dewetting in the case of non-spreading mineral oils. The results are characterized by negative values of spreading coefficients and positive
however as yet unpublished. values of bridging coefficients [30⁎]. Thus the available evidence
Theoretical challenges here concern the exact nature of the process [30⁎] suggests that the criteria represented by Eqs. (1)–(3) are still
by which a stretching oil drop ruptures a film — what are the conse- relevant for this non-aqueous system (making the substitutions of
quences for the drop, does it disproportionate into several smaller σeGL for σeAW; σeDL for σeOW; σeGD for σeAO where σeGL, σeDL and σeGD are the
drops as the foam film ruptures? What is the exact role, not only of cap- gas–crude oil, antifoam–crude oil and gas–antifoam surface tensions re-
illarity, but oil viscosity in the process and is the bridging–dewetting spectively). We therefore have for the formation of unstable bridging
mechanism likely to prevail at sufficiently high viscosities? There is configurations by a neat antifoam liquid in this non-aqueous context
also some evidence that the magnitude of the bridging coefficient
(or θ*) can determine relative antifoam effectiveness even when the e e
0 b B ≤ 2σ DL σ GL : ð5Þ
Eq. (3) is satisfied [28] for which theory has no convincing explanation.
A recent study by Gao et al. [29] attempts to simulate the bridging–
stretching process in the case of an oil drop in a foam film formed from However despite this overall similarity the differences in physical
an aqueous surfactant solution by recourse to a “steered molecular properties between water and hydrocarbon liquids produce some im-
dynamics” approach. Essentially this approach supplements classical portant differences in the criteria for selection of antifoam liquids. The
molecular dynamics with force probes to simulate processes which dielectric constant of hydrocarbons is nearly two orders of magnitude
are therefore only partially accounted for at a molecular level. In reality less than that of water which means that electrostatic effects derived
the relevant forces are capillary and disjoining pressures which are from charge separation of any solutes are essentially absent as shown
“mimicked” by arbitrary forces in this simulation, the molecular origins by Shearer & Akers [31] in the case of lubrication oils. Moreover adsorp-
of which are not revealed. In consequence it is not possible to link this tion of solutes at the air–hydrocarbon surface is expected to be small
simulation to quantitative experimental observation. Also the simula- [2⁎⁎,32]. In turn all this means that pseudoemulsion film stability is
tion produces only a two-dimensional view of the process little better likely to be low in this context and that therefore formulation of anti-
than the schematic diagrams to be found in the literature [2⁎⁎,3⁎]. foams with particles is unnecessary, as revealed by normal commercial
None of the issues listed in the previous paragraph have therefore practice. On the other hand the solubility of typical antifoam liquids
been addressed (or even can be addressed by this kind of approach?). such as polydimethylsiloxanes and perfluoropolymers in hydrocarbons
Perhaps a more fruitful approach would be the use of the surface energy is usually several orders of magnitude higher than in water (with poly-
minimization as exemplified by that described in Section 2.1 for the case dimethylsiloxane solubilities in crude oil of b10−3 g dm−3 and in water
of solid particles at interfaces. of b10−6 g dm−3 [2⁎⁎]). Even polydimethylsiloxanes, when present at
concentrations below their solubility limit act as pro-foamers for
3.2. Comparison of the criteria for antifoam action by neat oils in aqueous hydrocarbons [30⁎,31–33]. Effective foam control is then largely deter-
and non-aqueous liquids mined by solubility of such compounds. The study by Mansur and
coworkers [30⁎,33] of the antifoam behaviour of various alkoxylated
Realization of antifoam effects in aqueous solutions of surface active polydimethylsiloxanes in crude oil is illustrative. Here the solubilities
compounds by liquids which satisfy that Eq. (3) is inhibited by the in the crude oil are reduced by incorporation of the alkoxy groups in
metastability of oil–water–air pseudoemulsion films. That metastability the polydimethylsiloxane chains. The most weight effective of these
largely derives from the electrostatic and steric contributions to the rel- molecules is the least soluble in the crude oil — differences in antifoam
evant disjoining pressure isotherms due to adsorbed surfactant. It can effectiveness at equivalent concentrations above the solubility limits are
P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91 87

not apparent with similar values of σeGD, positive values of bridging geometry, contact angle and even particle size. Here we should note
coefficients and negative spreading coefficients [30⁎]. that even the orientation of particles of simple geometry at fluid–fluid
surfaces has been shown by surface energy minimization to be surpris-
3.3. “Cloud point” antifoams ingly complex (see Section 2).

Ethoxylated and propoxylated nonionic surfactants and polymers 4.2. Transient dynamic effects
exhibit high solubilities in water at low temperatures and partial misci-
bility at higher temperatures. Above a critical temperature, micellar Another aspect of antifoam function concerns the likelihood that the
solutions undergo a phase change to form two conjugate micellar rate of surfactant transport to the rapidly expanding air–water surfaces
solutions — one dilute and the other relatively concentrated. This pro- formed during foam generation is not fast enough to maintain equilibri-
cess is accompanied by the onset of turbidity where the cloud point is um. This factor can give rise to high dynamic surface tensions during
the temperature at which droplets of the concentrated conjugate foam generation and can therefore produce positive transient bridging
phase separate at a given overall surfactant concentration. Those drop- coefficients which become negative as surface tensions decline and
lets readily act as antifoams for solutions of the dilute phase by a bridg- trend towards equilibrium values after foam generation ceases. Similar-
ing mechanism apparently characterized by intrinsically unstable ly high dynamic surface tensions imply low dynamic adsorption levels
pseudoemulsion films (see [2⁎⁎] for a review of this phenomenon). and therefore diminished contributions to the magnitude of disjoining
Addition of ethoxylated and propoxylated nonionic compounds to forces. In turn this could reduce entry barriers and facilitate emergence
aqueous solutions of other surface active compounds can also produce of antifoam drops into the air–water surface. Approach to equilibrium
antifoam effects at temperatures above the cloud point. However the may subsequently lead to enhanced positive disjoining forces, failure
cloud point may be increased or even eliminated in such systems as a of drops to emerge into air–water surfaces and ineffective antifoam
result of the formation of mixed micelles. Nevertheless compounds action. Examples are known where the dynamic effect on the bridging
such as polyethoxy–polypropoxy block copolymers are often described coefficient appears to dominate. Li Ran [20] for example has recently
as “cloud point” antifoams. Despite this nomenclature Nemeth et al. [34] reported negative bridging coefficients using equilibrium surface
have shown that such compounds can exhibit antifoam effects when tensions for a tristearin/triolein antifoam in sodium dodecyl 4-phenyl
added to aqueous solutions of a protein–bovine serum albumin at low sulphonate solutions – antifoam effects are only apparent under
concentrations so that the cloud phase is absent. In such solutions the dynamic conditions – the foam becomes stable after foam generation
polymers simply displace the protein from the air–water surface to pro- ceases.
duce low foamability much as would be the case in the absence of the An obvious difficulty with these dynamic phenomenon concerns the
protein. Marinova et al. [35] recently report similar observations with actual state of the air–water surfaces during foam generation. Calcula-
this type of polymer but with a different protein–sodium caseinate. tion of bridging coefficients, for example, requires knowledge of the
However these observations include some curious features. Firstly in- air–water surface tension under dynamic conditions. Although dynamic
creasing the concentration of two of the polymers studied, up to con- surface tensions can be readily measured as a function of surface age, in
centrations where cloud phase is present, is reported to actually general, we have no clear knowledge of the actual effective surface age
increase the foam height. Since the activity of the polymer will be at a of the air–water surfaces during foam generation. Estimates have how-
maximum at phase separation it is difficult to see how that could de- ever been made, Garrett and Moore [28] have for example set the sur-
crease any antifoam effect concerning displacement of protein from face age during foam generation by Ross Miles technique as 0.5–1 s.
the air–water surface. Another interesting feature concerns the effect Denkov et al. [3⁎] estimate that surface ages during typical foam gener-
of addition of Ca2+ to the chosen system but with the polymer concen- ation as 0.1–1 s with a characteristic time for Ross Miles of 0.1–0.2 s.
tration high enough for cloud phase drops to be present. This apparently Differences in characteristic time with different foam generation
produces a precipitate of CaCO3, particles of which seem to be techniques mean that antifoam effectiveness can even vary depending
hydrophobed so that they adhere to the cloud phase drops. The pres- upon the technique as shown by Marinova et al. [1]. It would seem
ence of such particles appears to be essential for the cloud point anti- that enhanced dynamic antifoam effects are to be expected in decreas-
foam effect in this case because in the absence of Ca2 + there is ing order; hand-shaking cylinders (the Bartsch method) N Ross-
apparently little such effect. This then appears to represent an analogue Miles N bubble columns (see [2⁎⁎] for a description of these methodol-
of the well-known oil-particle antifoam synergy. In turn this suggests ogies). The key challenge here is of course direct measurement of the
that the pseudoemulsion films present in these systems are intrinsically surface tension of the upper air–water surface of foam during genera-
relatively stable in contrast to those formed in the case of aqueous dis- tion for each of these techniques.
persions of cloud phase in the absence of any additional surface active Transient antifoam effects accompanying foam generation have also
species. A detailed study of these issues would appear to be justified been shown to be apparent when foam is generated under micro-
in order to provide a firm basis for the optimisation of these cloud gravity. This behaviour has been observed during a comparison of
point antifoams which find wide application in dish washing formula- foam behaviour under micro-gravity in the International Space Station
tion, sugar beet processing, and fermentation. with that under earth gravity [36⁎]. This study included the effect of a
polydimethylsiloxane/silica antifoam on foam generation from aqueous
4. Mixed Oil-particle antifoams for aqueous solutions sodium dodecyl sulphate solutions using the Bartsch method [2⁎⁎].
Addition of this antifoam produced foam which totally collapsed in 2 s
4.1. Mode of action under Earth gravity using the selected conditions. That any foam gener-
ated rapidly collapsed under gravity implies that this antifoam is effec-
It is now well-known that oils which exhibit positive bridging coef- tive under both dynamic and the near equilibrium conditions prevailing
ficients in aqueous surfactant solutions are relatively ineffective as anti- after foam generation has ceased. In contrast significant amounts of
foams unless mixed with suitable hydrophobic particulate promoters stable foam were generated under micro-gravity. Not surprisingly
[2⁎⁎]. The role of the latter concerns reduction of the critical capillary Caps et al. [36⁎] attribute the presence of stable foam under micro-
pressure required to overcome the disjoining forces in the oil–water– gravity to the absence of gravity-driven drainage. Gas volume fractions
air pseudoemulsion films to permit film rupture by bridging oil drops of this residual foam were so low that foam films would not exist
[2⁎⁎,3⁎,7⁎⁎]. A detailed review of the present level of understanding is (i.e. with volume fractions of air ≤ ~ 0.72 [2⁎⁎]) — inter-bubble dis-
given in reference [2⁎⁎]. However there remain serious challenges to tances would be so thick that antifoam drops would not be able to
both experiment and theory, particularly concerning the role of particle form bridging configurations leading to foam collapse. However the
88 P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91

total volume of foam generated in the presence of the antifoam was sig- concentration may be used. Thus Garrett [2⁎⁎] has shown that a simple
nificantly less than that generated under gravity in the absence of anti- exponential can represent the antifoam concentration dependence of
foam. Clearly some foam films must have been formed during shaking foam behaviour in the case of sparging in the presence of polydimethyl-
of the relevant vessels to permit transient antifoam action to occur siloxane/silica antifoam at short times where deactivation is negligible
under micro-gravity despite the absence of gravity driven drainage. and concentration is therefore known. We can therefore write

Z
4.3. Antifoam deactivation t
V F ðt Þ ¼ V G expðm  cðt ÞÞdt ð7Þ
0
That polydimethylsiloxane (PDMS)/hydrophobed silica antifoams
gradually deactivate during the processes of foam generation (and where m is an empirical constant, VG is the volumetric gas flow rate and
even during emulsification) is well-known [2⁎⁎]. It has been shown VF(t) is the foam volume at time t in the presence of deactivating anti-
by Denkov and co-workers [3⁎] that this deactivation appears to be foam of residual antifoam concentration c(t). Integrating Eq. (6) and
driven by disproportionation of antifoam oil drops containing parti- substituting for c(t) in Eq. (7) produces an integral equation for the
cles into a population of inactive particle free drops and particle- time dependence of VF(t) containing three unknown constants — m, k1
enriched drops and agglomerates, where the latter can assume and k2. This equation can be fitted to a plot of VF(t) against time for a
sizes of the order of mms. A detailed review of this issue is to be given initial antifoam concentration, c(t = 0), to obtain values of those
found in [2⁎⁎]. constants. The fit is shown in Fig. 4. A test of theory is then the predic-
It has been shown that the process of disproportionation cannot tion of the time dependence of VF(t) for different initial antifoam con-
simply concern a random distribution of particles across drops [2⁎⁎] – centrations which are also presented in the figure. Semi-quantitative
that could in any case not account for the formation of large silica rich agreement is found — that agreement however represents a significant
agglomerates. Some process of coalescence is clearly implicated – in- improvement over earlier attempts [38] to interpret these results.
deed the presence of hydrophobed silica could facilitate that process
[2⁎⁎]. It has been suggested [2⁎⁎] that continuous drop splitting and co- 5. Mechanical defoaming
alescence could produce the disproportionation where the higher
viscosity of particle-rich drops inhibits splitting and facilitates coales- 5.1. Ultrasonics
cence. That drop splitting is also important is implied by overall de-
creasing rates of deactivation with increasing viscosity of the PDMS That ultrasound may represent a method of defoaming has been
[37]. Perhaps a rule-based simulation could be based on these observa- known for more than sixty years. Initial demonstrations of defoaming
tions to provide a clear explanation for the deactivation process. utilized sirens which proved impractical. Successful application re-
An additional complication however concerns the behaviour of quired the development of piezoelectric ultrasonic generators in the
other oil/particle antifoams. There is some evidence for example that late 1970s. A review of the history of defoaming by ultrasound is to be
hydrocarbon/hydrophobed silica antifoams also show deactivation found in [2⁎⁎].
when dispersions are subject to continuous aeration as do mixtures of Some context is provided if we consider the general features of the
PDMS with ethylene distearamide [2⁎⁎]. On the other hand there is interaction of sound with foam in the usual case where the ultrasonic
that with (organic liquid/organic particle)-based antifoams deactivation generator is placed above the foam, separated by an air gap, as shown
can be extremely slow (or even absent) [2⁎⁎]. For example antifoam in Fig. 5. Firstly we note that the velocity of sound in foam is significantly
mixtures of triolein with tristearin particles reveal little or no deactiva- less than that in either air or water [39] where we also remember that
tion after 2h of continuous aeration in solutions of sodium alkyl benzene the velocity of sound in water is more than three times faster than in
sulphonate [19]. air. This means that the wavelength of the sound pressure wave de-
If then all these various oil/particle antifoams are subject to the same creases as it passes into foam from the air as shown in the figure.
splitting and coalescence processes then we can have radically different Sound is also attenuated when passing through foam. Thus the rate of
outcomes. It seems likely that those differences derive from the ease
with which intractable particle-rich entities can be formed. This in
turn may concern the viscosity of the oil, the rheology of the oil/particle 3
Foam volume, V(t)/ cm
mixtures as a function of particle volume fraction and concentration, de- 1000
gree of aggregation and even particle geometry. Systematic experimen- 900
tal studies of the behaviour of different antifoam types with respect to calculated
800
rates of deactivation and characterized with respect to this list of vari- 700
ables would seem to have merit especially if combined with attempts 600
to develop a rule-based simulation of the deactivation process. 500
Experimental investigations of the effect of antifoam concentration 400 fit
on the rate of deactivation are extremely rare. One such investigation 300
is reported by Pelton [38] for a PDMS/silica antifoam with foam gener- 200
ated by sparging. A theoretical treatment of deactivation which affords 100
comparison with these experiments has been published recently 0
[2⁎⁎]. Somewhat simplistically we can represent the deactivation pro- 0 500 1000 1500 2000 2500
cess as the sum of a first order splitting phenomenon and a second
Time/s
order coalescing phenomenon so that the rate of loss, dc(t)/dt, of effec-
tive antifoam entities is given by
Fig. 4. Deactivation of a hydrophobed silica/polydimethylsiloxane antifoam emulsion
dispersed in 5 g dm−3 of a dishwashing liquid [38]. Foam generated by passing nitrogen
dcÞt 2 through a porous frit at 7 cm3 s−1. Antifoam concentrations, ■, 0.1 g dm−3; △,
− ¼ k1 cðt Þ þ k2 cðt Þ ð6Þ 0.3 g dm−3; and ♦, 0.5 g dm−3. Fit to results at 0.5 g dm−3 yielded m = −9.65 g−1 dm3;
dt
k1 = 1.3 × 10−6 s−1; k2 = 1.36 × 10−3 g−1 dm3 s−1 in Eqs. (6) and (7). These values
were used to calculate the plots shown for 0.1 and 0.3 g dm−3.
where k1 and k2 are rate constants and c(t) is the concentration of active ([2⁎⁎], © 2013 from The Science of Defoaming: Theory, Experiment and Applications by
antifoam at time t. Integration of Eq. (6) yields c(t) as a function of time. P.R.Garrett. Reproduced by permission of Taylor and Francis Group, LLC, a division of
An empirical relationship between the known foam volume and Informa plc).
P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91 89

source top of the foam column [2⁎⁎], only about 10% of the acoustic energy is
transmitted into the foam. The imbalance is even more marked at the
water–foam interface. The high velocity of sound in water means that
essentially no acoustic energy will be transmitted across that interface
regardless of whether the transmission is from water to foam or foam
to water as shown in Fig. 5. Winterburn and Martin [42] have in fact
shown that placing the ultrasound generator in the aqueous phase is in-
effective in causing defoaming.
Although defoaming by ultrasound can be achieved there have ap-
air parently been few attempts to understand or even optimize the process.
However some early insight is provided by the work of Komarov et al.
[43]. This study in part concerned the effect of low frequency sound
(using a loud speaker and therefore not in the ultrasound range) on
the steady state foam height obtained by sparging a solution of a surfac-
tant in aqueous mixtures of different proportions of glycerin and water.
For a given frequency defoaming became apparent only at a threshold
acoustic intensity (as measured by the acoustic pressure). Increase in
frequency however increased the threshold value of the acoustic inten-
sity. Increasing the viscosity of the solution decreased the threshold
foam slightly. All of this tends to suggest an optimal frequency where the
acoustic energy required for a given extent of defoaming is minimal.
Rodriguez et al. [44] have more recently made a study of the effect of
acoustic intensity on defoaming a fixed volume of foam sparged from an
aqueous solution of “soap”. This utilized only one ultrasonic frequency
of 25.8 kHz which was generated from a piezoelectric device. As found
by Komarov [43] there appears to be a threshold acoustic intensity
below which defoaming does not occur. Beyond the threshold
defoaming increases with increase in the total amount of acoustic ener-
gy — i.e. the higher the acoustic intensity and the longer it is applied
liquid above the threshold the greater the volume of foam destroyed. Usefully
Rodriquez et al. [44] include gas volume fractions and estimates of bub-
ble sizes in their study. The wavelength of the ultrasound can be calcu-
lated from the frequency and knowledge of the velocity of sound as a
function of gas volume fraction in a foam (see for example Kann [39]).
Fig. 5. Schematic showing propagation of sound waves through foam in a bubble column. We find then that in the case of foam with bubble diameters of
Continuous lines represent progressive waves and broken lines represent reflected waves. 0.2–2 mm the acoustic wavelength was ~4 mm and in the case of bubble
(after Komarov, S.V. et al. [43]. Originally published in ISIJ International).
diameters of 0.2–10 mm the wavelength was ~ 6 mm. These findings
hint at the possibility that defoaming by ultrasound concerns a
delivery of acoustic energy per unit area, the acoustic intensity IH, after resonance with the mesostructure of the foam. The nature of any quan-
passing through a foam layer of height H is given by [40] titative relationship between the bubble diameters and acoustic wave-
length would however require detailed knowledge of the bubble size
IH distributions so that at least mean bubble diameters can be calculated.
¼ expð−2αH Þ ð8Þ
IA A recent paper by Ben Salem et al. [45⁎⁎] represents a significant de-
velopment in the understanding of the propagation of ultrasound in
where IA is the acoustic intensity transmitted through the air–foam in- aqueous foams. In this work the ultrasound frequency was again main-
terface and α is the attenuation coefficient. The attenuation coefficient tained constant but at acoustic intensities below the threshold for
increases with increasing frequency [40]. The acoustic intensity is pro- defoaming. The bubble size of the foam was measured by image analysis
portional to the square of the amplitude of the pressure fluctuations and allowed to vary by the process of diffusional disproportionation.
caused by the sound wave. The amplitude of that wave is therefore The change in amplitude of the ultrasound signal of constant incident
shown in Fig. 5 to decline as it passes through the foam. However the acoustic intensity across a sample of foam of known thickness was
acoustic intensity arriving at a given interface is partially transmitted used as a measure of the attenuation of acoustic intensity. The phase dif-
and partially reflected at that interface. In general then [41] ference yielded the transit time across the foam and therefore the speed
of sound in that context. The relative amplitude varied as a function of
IA 4Z 1 Z 2 time as the bubble sizes increased due to diffusional disproportionation.
¼ ð9Þ
Ii ðZ 1 þ Z 2 Þ2 A plot of relative amplitude against time yielded a pronounced mini-
mum at a “transition time”, implying a maximum rate of energy absorp-
where Ii is the incident intensity at the interface. Z1 and Z2 are the spe- tion by the foam. A separate plot of the evolution of bubble size against
cific acoustic impendances of mediums 1 and 2 respectively where the time meant that this transition time could be related to a “critical” bub-
specific impedance of a medium is the product of the density and the ble size. Ben Salem et al. [45⁎⁎] used these techniques to compare the
velocity of sound in that medium. Here we note that if Z1 = Z2 the trans- transition times and critical bubble sizes of two radically different
mission is total but if there is a marked imbalance between the acoustic foams — a shaving foam in air and a dilute SDS solution in C6F6. Despite
impendances so that either Z1 ≫ Z2 or Z1 ≪ Z2 then transmission is min- radically different extents of acoustic attenuation and rates of dispro-
imal. In the case of foam of gas volume fraction 0.999 at the top of the portionation the critical bubble sizes of each of these foams were in
foam column, for example, the acoustic transmission is almost 94% at the same range, i.e. 50 ± 20 μm. Ben Salem et al. [45⁎⁎] then argue
the air–foam interface [2⁎⁎]. However in the case of a very wet foam, that at the critical radius the “bubbles become resonant, hence they
with a gas volume fraction at the Kugelschaum limit of ~ 0.72 at the pump a maximum of energy from the incident signal”. We should
90 P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91

however note that the critical bubble radius is about an order of magni- 6. Summarizing remarks
tude less than the relevant acoustic wavelengths — this leaves open the
question of what energy absorption process could scale in such a man- We conclude by summarizing the significant recent developments
ner. Perhaps it concerns the number and size of foam films rather than and the remaining challenges in the study of defoaming. Use of surface
that of the bubbles [46]. Nevertheless it would appear surprising if the energy minimization clearly represents an example of such a develop-
critical bubble radius is not also some function of the acoustic frequency. ment, representing a useful technique for predicting the orientations
These findings of Ben Salem et al. [45⁎⁎] may be relevant for of particles at surfaces. It may also represent a means of simulating
defoaming by ultrasound where the latter may also involve resonance other phenomena of importance in the study of defoaming such as the
but at higher incident acoustic intensities above the threshold leading bridging–stretching process and the effect of bridging drops in Plateau
to foam film destruction of bubbles of the critical size. This could suggest borders. Limitations include exclusion of disjoining forces in thin films
a possible speculative model for defoaming by ultrasound. Thus and, of necessity, viscous dissipation.
defoaming becomes maximal if the bubble sizes are at the critical size Foam generation is a dynamic phenomenon where air–water sur-
range for the resonant frequency. Increasing the intensity at that fre- faces may depart significantly from equilibrium if the rate of surfac-
quency will produce progressively more effective defoaming. However tant transport to surfaces is slow relative to the rate of formation of
increasing the frequency above the resonant frequency for a foam of a those surfaces. More evidence has been obtained to show that this
given bubble size will mean less efficient energy transfer into the can lead to transient antifoam effects which only occur during foam
foam, needing higher threshold acoustic intensities for defoaming as generation. In turn this can mean that the effectiveness of antifoams
shown by Komarov [43]. If realistic all this could imply that optimal can vary according the method of foam generation. A key and very
defoaming requires multiple frequency generators tuned to the bubble difficult challenge here is the characterization of the air–water
sizes in the polydisperse foams of practical interest. It does not however surface during actual foam generation. Such transient effects
address the nature of the actual process of foam collapse during the concerning antifoam action are also apparent under micro-gravity
application of ultrasound. but are then due to low gas volume fractions and the absence of liq-
uid drainage after foam generation has ceased and do not concern
5.2. Use of rotary devices transport to air–water surfaces.
Precipitation of anionic surfactant from aqueous solution as
Various rotary devices have been suggested for the control of foam, mesophase particles can be induced by the addition of, for example,
especially in sparged unstirred and stirred bubble columns [2⁎⁎]. excess counterions. Such precipitation is accompanied by declines
These devices are exemplified by spinning blades, rotor-stators, in foamability due to the slow transport of the surfactant present in
turbines, and discs, placed at the top of these bubble columns. There the mesophase particles to the air–water surface. Similar behaviour
appears to be little consensus about the mode of action of such devices, is apparent in the case of surfactants at temperatures below the rel-
being variously attributed to either centrifugal or shear forces [2⁎⁎]. evant Krafft temperature where the precipitate is then crystalline.
However the absence of any significant recent published literature on However the hydrophobic nature of such precipitates of surfactant
this matter places the subject outside the scope of this review — an means that antifoam effects are likely to be superimposed upon the
account of earlier work is to be found in reference [2⁎⁎]. transient effects due to slow surfactant transport to air–water sur-
faces. These generalisations are, however, based upon a relatively
5.3. Orifice defoamer small number of examples.
Extension of modern insights into the mode of action of antifoams
Liu et al. [47] describe a defoaming device consisting of a collection in aqueous systems to the study of non-aqueous systems has been
of perforated plates. Defoaming is induced by simply passing foam essentially absent until recently. In the case of a hydrocarbon based
through the orifices in the plates. However the orifices must have small- liquid like crude oil it has recently been shown that antifoam action
er diameters than those of the bubbles in the foam. This arrangement by liquid alkoxylated polydimethylsiloxane antifoams occurs with
means that large deformations of bubbles occur as they pass through positive bridging coefficients and negative spreading coefficients.
the orifices. Foam films are thereby stretched until they rupture. This The main differences with respect to aqueous systems concern
in turn causes bubble coalescence and collapse, both of which contrib- (1) the absence of factors which stabilize pseudoemulsion films so
ute to defoaming. Observations of defoaming with this device were that particles are not a necessary components of the antifoams and
however confined to aqueous submicellar sodium dodecyl sulphate (2) the relatively high solubility of these PDMS derivatives in the
(SDS) solutions of concentrations ≤~3.5 × 10−3 M. foaming liquid (i.e. crude oil).
A measure of the resistance to stretching of a foam film is provided Mixtures of PDMS and hydrophobed silica are known to be effective
by the Gibbs elasticity. Lucassen [48] has presented plots of Gibbs antifoams for aqueous systems. However these antifoams deactivate in
elasticity, εG, for submicellar solutions of SDS which show a maximum use. The process involves disproportionation of the dispersed antifoam
at ~ 2.5 × 10− 3 M for film thicknesses in the region of 1–2 μm. At into inactive oil drops without particles and particle rich drops and ag-
concentrations of SDS lower than that maximum the gradient of Gibbs glomerates. A simple theory based upon an assumption that this
elasticity with respect to concentration, dεG/dc, is positive. This means process involves both first order drop splitting and second order coales-
that if a foam film is stretched so that both the thickness and the cence is in semi-quantitative agreement with experimental observa-
intra-lamellar surfactant concentration are reduced then the Gibbs elas- tions of deactivation. However there is evidence that deactivation is
ticity also decreases [2**,48]. Lucassen [48] argues that the film will then not a common feature of all types of oil/particle antifoams. There is a
be dynamically unstable. This conclusion appears to be consistent with need therefore to study this issue in order to improve overall under-
the observations of Liu et al. [47] who report decreasing defoaming ef- standing of deactivation so that more effective antifoams may be
fects with increasing SDS concentration up to ~3.5 × 10−3 M which is developed.
close to the maximum experimental value quoted by Lucassen [48]. Recent accounts of defoaming by mechanical means are largely con-
Any discrepancy could be due to the role of dynamic factors in the func- fined to the use of ultrasonics. There are indications that defoaming by
tion of the device used by Liu et al. [47]. However if the apparent consis- application of ultrasound is a resonant phenomenon where an optimum
tency with the calculations of Lucassen [48] represents the reason for frequency is tuned to a critical foam dimension which unsurprisingly
the defoaming effect then increasing the surfactant concentration so appears to be some function of bubble size. More effective application
that dεG/dc b 0 (under the relevant dynamic conditions) will tend to of ultrasound in this context would appear to require detailed study of
eliminate that effect. this issue.
P.R. Garrett / Current Opinion in Colloid & Interface Science 20 (2015) 81–91 91

Acknowledgement [23] Garrett PR. Preliminary considerations concerning the stability of a liquid hetero-
geneity in a plane-parallel liquid film. J Colloid Interface Sci 1980;76(2):587–90.
[24] Arnaudov L, Denkov ND, Surcheva I, Durbut P, Broze G, Mehreteab A. Effect of oily
The author is grateful to Dr. Gareth Morris (Imperial College) for additives on foamability and foam stability. I. Role of interfacial properties. Lang-
supplying the images shown in Fig. 2. muir 2001;17:6999–7010.
[25] Basheva ES, Stoyanov S, Denkov ND, Kasuga K, Satoh N, Tsujii K. Foam boosting by
amphiphilic molecules in the presence of silicone oil. Langmuir 2001;17:969–79.
References [26] Zhang H, Miller CA, Garrett PR, Raney KH. Mechanism of defoaming by oils and
calcium soap in aqueous systems. J Colloid Interface Sci 2003;263:633–44.
[1] Marinova K, Tcholakova S, Denkov N. Role of foam dynamics for antifoam activity. [27] Shu X, Meng Y, Wan L, Li G, Yang M, Jin W. pH-Responsive aqueous foams of oleic
EUFOAM; 2014[to be published]. acid/oleate solution. J Dispers Sci Technol 2014;35:293–300.
[2⁎⁎] Garrett PR. The science of defoaming, theory, experiment and applications. Surfac- [28] Garrett PR, Moore PR. Foam and dynamic surface properties of micellar alkyl
tant Sci. Series, vol. 155. New York: Taylor and Francis; 2013.This is the only com- benzene sulphonates. J Colloid Interface Sci 1993;159:214–25.
prehensive monograph dealing with the subject of defoaming including an [29] Gao F, Yan H, Wang Q, Yuan S. Mechanism of foam destruction by antifoams: a
account of the mode of action of antifoams and featuring some material not pub- molecular dynamics study. Phys Chem Chem Phys 2014;16:17231–7.
lished elsewhere. The book also includes a review of mechanical defoaming to- [30⁎] Fraga AK, Santos RF, Mansur CRE. Evaluation of the efficiency of silicone polyether
gether with detailed accounts of applications ranging from detergent products additives as antifoams in crude oil. J Appl Polym Sci 2012;124:4149–56.For argu-
to gas–oil separation in crude oil production. ably the first time the modern concept of the bridging coefficient is successfully
[3⁎] Denkov N, Marinova K, Tcholakova S. Mechanistic understanding of the modes of applied to the interpretation of antifoam effects in a non-aqueous system.
action of foam control agents. Adv Colloid Interface Sci 2014;206:57–67.A good [31] Shearer LT, Akers WW. Foaming in lube oils. J Phys Chem 1958;62:1269–70.
review of the essentials of antifoam action. [32] Binks BP, Davies CA, Fletcher PDI, Sharp EL. Non-aqueous foams in lubricating oil
[4] Karakashev SI, Grozdanova MV. Foams and antifoams. Adv Colloid Interface Sci systems. Colloids Surf A Physicochem Eng Asp 2010;360:198–204.
2012;176–177:1–17. [33] Rezende DA, Bittencourt RR, Mansur CRE. Evaluation of the efficiency of
[5] Owen MJ. Foam control. Advances in Silicones and Silicone-Modified Materials, polyether-based antifoams for crude oil. J Petrol Sci Eng 2011;76:172–7.
vol. 1051. ACS Symposium series; 2010. p. 269–86. [34] Nemeth Z, Racz G, Koczo K. Antifoaming action of polyoxyethylene–
[6] Miller CA. Antifoaming in aqueous foams. Curr Opin Colloid Interface Sci 2008;13: polyoxypropylene–poloxyethylene-type triblock copolymers on BSA foams. Colloids
177–82. Surf A Physicochem Eng Asp 1997;127:151–62.
[7⁎⁎] Denkov ND. Mechanisms of foam destruction by oil-based antifoams. Langmuir [35] Marinova KG, Dimitrova LM, Marinov RY, Denkov N, Kingma A. Impact of the sur-
2004;20:9463–505.A thorough review of the work of Denkov and co-workers factant structure on the foaming/defoaming performance of nonionic block copol-
concerning the mode of action of oil-based antifoams in aqueous solutions. ymers in Na caseinate solutions. Bulg J Phys 2012;39:53–64.
[8] Dippenaar A. The destabilization of froth by solids. 1. The mechanism of film rup- [36⁎] Caps H, Vandewalle N, Saint-Jalmes A, Saulnier L, Yazhgur P, Rio E, et al. How
ture. Int J Miner Process 1982;9:1–14. foams unstable on Earth behave in microgravity? Colloids Surf A Physicochem
[9⁎] Joshi K, Baumann A, Jeelani SAK, Blickenstorfer C, Naegeli I, Windhab EJ. Mecha- Eng Asp 2014;457:392–6.Reveals something of the potential for study of antifoam
nism of bubble coalescence induced by surfactant covered antifoam particles. J action under conditions where gravity driven drainage is absent permitting vari-
Colloid Interface Sci 2009;339:446–53.Presents an interesting new approach to ation of gas volume fractions without that complication.
study of antifoam phenomena. [37] Koczo K, Koczone J, Wasan DT. Mechanisms for antifoaming action in aqueous
[10] Brooks JH, Alexander AE. The spreading behaviour and crystalline phases of fatty systems by hydrophobic particles and insoluble liquids. J Colloid Interface Sci
alcohols. Part IV. Equilibrium spreading pressure. In: La Mer VK, editor. Retarda- 1994;166:225–38.
tion of Evaporation by Monolayers: Transport Processes. Academic Press; 1962. [38] Pelton R. A model of foam growth in the presence of antifoam emulsion. Chem Eng
p. 259–69. Sci 1996;51(19):4437–42.
[11] Morris GDM, Neethling S, Cilliers JJ. The effects of hydrophobicity and orientation [39] Kann KB. Sound waves in foams. Colloids Surf A Physicochem Eng Asp 2005;263:
of cubic particles on the stability of thin films. Miner Eng 2010;23:979–84. 315–9.
[12] Morris GDM, Neethling S, Cilliers JJ. A model for investigating the behaviour of [40] Komarov SV, Kuwabara M, Sano M. Control of foam height using sound waves. ISIJ
non-spherical particles at interfaces. J Colloid Interface Sci 2011;354:380–5. Int 1999;39(12):1207–16.
[13] Morris GDM, Cilliers JJ. Behaviour of a galena particle in a thin film, revisiting [41] Rayleigh JWS. The Theory of Sound, vol. II. New York: Dover Publications; 1945. p.
Dippenaar. Int J Miner Process 2014;131:1–6. 69 [Chp. 13].
[14⁎⁎] Morris GDM, Neethling S, Cilliers JJ. An investigation of the stable orientations of [42] Winterburn JB, Martin PJ. Mechanisms of ultrasound foam interactions. Asia Pac
orthorhombic particles in a thin film and their effect on its critical failure pressure. J Chem Eng 2009;4:184–90.
J Colloid Interface Sci 2011;361:370–80.Illustrates the use of energy minimisation [43] Komarov SV, Kuabara M, Sano M. Suppression of slag foaming by a sound wave.
simulation to reveal the orientations of particles in surfaces and foam films to dis- Ultrason Sonochem 2000;7:193–9.
cover orientations otherwise unrealized. This illustrates a technique which has [44] Rodriguez G, Riera E, Gallego-Juarez JA, Acosta VM, Pinto P, Martinez I, et al. Exper-
some promise in predicting the potential of particles of various geometries as imental study of defoaming by air-borne power ultrasonic technology. Phys
foam and pseudoemulsion film breakers. Procedia 2010;3:135–9.
[15] Brakke K. The surface evolver. Exp Math 1992;1:141–65. [45⁎⁎] Ben Salem I, Guillermic R, Sample C, Leroy V, Saint-Jalmes A, Dollet B. Propagation
[16] Morris GDM, Hadler K, Cilliers JJ. Particles in thin liquid films and at interfaces. of ultrasound in aqueous foams: bubble size dependence and resonance effects.
Curr Opin Colloid Int Sci 2015;20:98–104. Soft Matter 2013;9:1194–202.Clear evidence of resonance between maximum
[17] Luangpiron N, Dechabumphen N, Saiwan C, Scamehorn JF. Contact angle of surfac- sonic energy absorption at a given frequency and a critical bubble size is present-
tant solutions on precipitated surfaces. J Surfactant Deterg 2001;4(4):367–73. ed. This may be applicable to defoaming and point the way to a more effective ap-
[18] Peck TG. The Mechanisms of Foam Breakdown by Oils and Particles. (PhD Thesis) proach than hitherto.
University of Hull; 1994. [46] Pierre J, Dollet B, Leroy V. Resonant acoustic propagation and negative density in
[19] Ran L, Jones SA, Embley B, Tong MM, Garrett PR, Cox SJ, et al. Characterisation, liquid foams. Phys Rev Lett 2014;112:148307.
modification and mathematical modeling of sudsing. Colloids Surf A Physicochem [47] Liu Y, Wu Z, Zhao B, Li L, Li R. Enhancing defoaming using the foam breaker with
Eng Asp 2011;382:50–7. perforated plates for promoting the application of foam fractionation. Sep Purif
[20] Ran L. Foaming of Anionic Surfactant Solutions in the Presence of Calcium Ions and Technol 2013;120:12–9.
Triglyceride-Based Antifoams. (PhD Thesis) University of Manchester; 2011. [48] Lucassen J. Dynamic properties of free liquid films and foams. In: Lucassen-
[21] Brochard-Wyatt F, di Maeglio J, Quere D, de Gennes P. Spreading on nonvolatile Reynders EH, editor. Anionic Surfactants, Physical Chemistry of Surfactant Action.
liquids in a continuous picture. Langmuir 1991;7:335–8. Surfactant Sci. SeriesNew York: Marcel Dekker; 1981. p. 217 [Chp 6].
[22] Rowlinson JS, Widom B. Molecular Theory of Capillarity. Oxford: Oxford University
Press; 1982.

Вам также может понравиться