You are on page 1of 47

Advances in Colloid and Interface Science, 40 (1992) 3’1-83 37

Elsevier Science Pub&here B.V., Amsterdam

99989A

WET FOFIMS: FORMATION. PROPERTIES AND MECHANISM OF STABILITY

Kazimierz MAtYSA

Institute of Catalysis and Surface Chemistry Polish Academy of

Sciences, ul. Niezapominajek, 30-239 Cracow, Poland.

Contents

Abstract 37

1. Introduction 39

2. Foam formation 41
2.1. Bubbles formation and motion 41
2.2. Role of the wall effects 46
3. Measurements 49
3.1. Foams 49
3.2. Single foam films 54
4. Composition of the wet foam 57
4.1. Water content within the foam height 57
4.2. Thickness of rupture of the foam films 61
4.3. The foam model 63
9
. . Forces stabilizing the foam 64
5.1. Surface elasticity forces 65
5.2. FI relation between the foam stability and surface

elasticity forces 73
6. Concluding remarks 77

ABSTRACT

Overall picture of phenomena occuring during formation and

existence of the wet foams is presented. Properties and mechanism

This paper is dedicated to Professar A. Scheludko on the occasion

of his 70th birthday.

Al-~/92/$15.~ 0 1992 - Elsevier Science Publishers B.V. AI1 righta reserved.


38

of stability are discussed on the example of the wet foams obtained


from solutions of two homologous series of surface active
substances; the fatty acids and n-alkanols. In general
three physical processes which contribute to foam stability can be
distinguished: drainage of liquid out of the foam, coalescence
and/or rupture of bubbles, and disproportionation (which may be
called Ostwald ripening or gas diffusion from one bubble to
another). Dynamic and non-equilibrium character of the wet foams 1s
stressed.

Motion of a bubble through the solution causes disequilibration


of the surface concentration alongside the bubble surface. The
surface concentration on the upstream part of the bubble is much
smaller than the equilibrium concentration. Thus, the bubbles
arrive at the solution surface with non-equilibrium surface

concentration, and these actual non-equilibrium surface covet-ages


determine possibility of formation and properties of the foams.

Solution content 8 in the volume of wet foam is high (of an


order 30%), while in top foam layer it 1s much smaller (pZ5%). It
shows that rupture of the wet foam takes place practically only in
the top layer of bubbles and durability of these toe foam films
determine stability and volume of the whole foam column. On the
basis of measurements of liquid content 8 and lifetimes of bubbles
in the top foam layer it was estimated that thicknesses of rupture
of these top films were of an order of a few micrometers. Fit such
thicknesses the force of disjoining pressure do not attain yet any
meaningful value.

Influence of kinetics of adsorption, frequency of external


disturbances, surface activity of the solute and lifetime of the
foam films on magnitude of the surface elasticity forces induced in
the systems studied is discussed. It is shown that stability of the
wet foams can be explained in terms of the effective elasticity
forces, i.e. the surface elasticity forces which are induced at an
actual non-equilibrium surface coverage. There is agreement between
the courses of the dependence5 of the foamability parameter
(retention time. rt) and the effective elasticity forces a5 a
function of the number n of carbon atoms in the fatty acid and
39

n-alkanol molecule. This shows that the effective elasticity forces


are decisive parameter in formation and stability of the wet foams.

It also explains why the foamability of a substance with a stronger


surface activity can be lower than that of a substance with a
weaker surface activity. The foamability, especially under dynamic
conditions, cannot simply be correlated with the surface activity.

1. XNTRCYDUCTION

Foams are colloidal systems comprising a gas being the

dispersed phase and a liquid (or solid in the.case of the solid


foams) being the continous phase. They are commonly met in every
day life (washing, cleaning, shampoos, etc.) and in many
industrial applications (froth flotation, foam fractionation,
fire-fighting, wastes treatment, etc.). Depending on application5
the foams with different properties are required. For example slow
draining and very stable foams are sought for fire-fighting
purposes. Such foams should resist destruction during contact with
the fuel and exposure to high temperature. The iow density of the
foam ensures that it floats on burning oil or gasoline and inhibits
oxygen supply. On the other side foams of very low stability are
required in flotation processes. According to Kitchener El3 froth

flotation can be claimed to be the largest application of surface


chemistry, as an estimated lo0 tonne of ground ore was treated by
flotation annually and foam is inseparably connected with these
processes. During flotation the grains of useful mineral are
concentrated in the foam and they are collected by cutting-off a
foam layer containing the concentrate. Therefore, the foam should
be stable enough to enable concentration processes to occur and

simultaneosly it should rupture quickly outside the flotation


machine. Too stable foam overflows the flotation machines and
causes serious technological problems.
The control, inhibition or destruction of foam is important
in many industrial processes. The best way of the foam control
would be to get to know all factors affecting foam formation and
stability, and their mutual interdependences. A dynamic foam is a
40

complicated physicochemical system and therefore there still does


not exist a satisfactory comprehensive theory of foam stability.
It is well established CZ-63 that pure liquids do not foam and a
presence of a second component is a necessary but not sufficient
condition for foam formation. Foam properties depend on many
factors from which composition of the solution, type of a surface
active substance, number of surfactants used, a presence of
contaminations, conditions of the foam formation and existence are
the most important. CI wide range of foam persistence is observed
with different solutions. Foam film stability ranges from fraction
of seconds to years (if they are isolated from external
disturbances). There is no sharp transition from a weakly-foaming
to a strongly-foaming or non-foaming solution, but rather a
continous transition of properties. Nevertheless terms: transient
or wet or unstable foams and dry or metastable foam5 are often
used to describe systems being near two opposite limits of a
spectrum of the foam stability.

Wet or transient foams are an assembley of spherical bubbles


separa'ted by thick liquid walls (lamellael. They can exist
practically only during an action of foam forming agent (force).
When the foam formation is stopped the wet foams collapse in a
time not exceeding a few tens of seconds. On the other end of the
spectrum there are dry or metastable foams in which bubbles (foam
cells) acquire a polyhedral shape. The bubbles are separated by
planar or only slightly curved liquid films of very small
thicknesses. Such foams can be stable for hours or even days after
formation. It is worth to point here that the metastable foams
must pass, during their formation, through a stage at which they
consist of spherical bubbles separated by thick liquid films c71.
Topic of the paper is focused on wet foams but they are discussed
as a part of a more general problem of stability of foams. Mutual
importance of various factors affecting formation and stability of
wet foams is discussed in a frame of some general regularities
commom for all foam systems.
41

2. FOAM FORMCITION

Foams, like others two-phase colloidal Systems, can be


generated by either dispersion or expansion (condensation) methods.
In dispersion method5 the future discontinous phase is initially
available a5 a large volume of gas. Foam generation in a dispersion
method results from mixing gas and solution together with an input
of energy, as in shaking, whipping, pouring, etc. Here, the most
common technique is bubbling gas through small orifice5 such as
capillaries, fritted glass disks or drilled plates what results in
dispersing gas into bubbles of sizes being mainly a function of
solution properties,orifices dimension5 and gas flow rate.

In expansion method5 the future dispersed phase is initially


present as a solute, that is, as molecules dissolved in the liquid.
Condensation results from generation of local gas pocket5 within
the solution to be foamed. Properties of the whole system have to
be changed in a such manner that the solution becomes
supersaturated with gas. Gas can be generated either chemically
(chemically generated fire-fighting foams) or by microbiological
processes (fermentation) or by lowering pressure in the system
(soft drinks).

2.1 Bubbles formation and motion

In studies of wet foam5 mainly dispersion method5 are used


and most commonly gas is injected into a solution through one or
many orifices. Problem of a bubble formation at a single orifice
is quite analogous to that of drop formation in stalagmometric
methods of surface tension measurements. When bubble is formed and
grow5 the buoyancy force tends to lift it while the surface
tension forces act in the opposite direction. When the bubble
grows slowly then its detachment starts a5 5oon as the buoyancy
force equilibrate5 or exceeds the surface tension forces, i.e.,
when:
v,@g 2 2nrd (11
where v b is the bubble volume, g is the acceleration due to
gravity, Ap =(pQ - ps) is the difference between the densities of
gas and liquid respectively, r is the radius of the orifice and d

is the solution surface tension. The equation (1) is a very crude


approximation but it shows the principal forces acting during
formation and detachment of the single bubble. There are several
respects in which the Eq.(ll is inadequate, perhaps the most
obvious being that in practice the whole bubble never detaches
from the orifice - similarly as the whole drops never fails from
the capillary tip. However, while in the case of falling drop an
empirical calibration of the drop-volume method has been given by
Harkins and Brown [S] and these correction factors have been
commonly used and improved [9], there do not exist such tabulated
corrections for the bubbles detachment. When bubbles formation
occurs quickly and/or in streams and/or from multiple orifices
then the problem complicates further because of mutual interaction
between growing and detaching bubbles, motion of the liquid near
the sparger, lack of adsorption equilibrium on expanding bubble
surface, etc. cls a result of it the dimensions of the bubbles
formed on a porous disk become also a function of the gas flow
rate [3, 10-121. Similarly a number of effective sites Cl21 for
the bubble formation is found to be a function of not only surface
tension, viscosity and density of the liquid, but also of the gas
flow rate through the disk.

A bubble detached from a orifice starts to flow and reaches a


terminal velocity which value depends on the bubble size and
solution properties. Hydrodynamics of the bubble motion has been a
subject of a number of studies C13-271. In a pure liquid, under
conditions of laminar flow, a rising bubble moves faster than
predicted by Stokes' law. Hadamard Cl31 and Rybczynski Cl41
extended Stokes' theory for a motion of gas bubble or liquid drop
through a infinite medium and the terminal velocity of the bubble
ub is given by:

2 'P1. - PQ) 9 r2
w (21
ub=
97-l
43

where at, the Hadamard-Rybczyhski correction factor to Stokes' law,


has the value:

3r) - 3 T)L
9
x = (3)
37-l +27? 1
9

and and nL are viscosities of the gas inside the bubble and
r) of
9
the liquid, respectively. For the gas bubble in water, where n <<
9
the value of 1( is equal to 312 and the bubble should move 50%
Y
faster than solid sphere of the same radius and density.
C\ccording to the theory [13,141 this is due to the mobility of
liquid/gas interface. The medium exerts a viscous drag on the
surface of the bubble and so sets up a circulation of the gas
contained inside. 6s a result of it the velocity gradients are
smaller than those in the case of a solid interface. Decreasing the
velocity gradients leads to a decrease in the energy dissipated in
the liquid and therefore bubbles or fluid drops move faster.

Presence of a surface active substance in solution reduces


the bubble terminal velocity C15-241, because adsorption of
surfactant molecules at gas/liquid interface lowers mobility of
the interface. Boussineq C281 offered an explanation of this
effect in term5 of an enhanced viscosity of the interface.
CIccording to him when a surface active substance is present in the
medium then viscosity of the interface is enhanced and it reduces
internal circulation on both sides of the interface. He derived an
alternative correction factor to Stokes' law in which x has the
value:
3T +37I
9
+ 2 e/r
b
X= (4)
2 +37-l + 2 e/r
? 9 b

where e is the surface dilational viscosity (N rnmi s) which


expresses the relation between surface tension and the rate at
which the area of the surface of the liquid changes. It is seen
that for big bubbles the influence of the surface viscosity can be
negligible small, while for small bubbles, i.e., when e/r b * (8, +
'09L the opposite can be true. Thus, in the Boussineq's theory the
44

phenomenological parameter e depends only on properties of the


surface layer and is not dependent on properties of the volume
phase [23].

Basing on eperimental findings that bubble or drop motion


depends strongly on a presence and concentration of surface active
substances in the solution Levich and Frumkin ci5, 291 have
developed so called adsorption theory of retardation. Ciccording to
it the adsorbed solute is not uniformly distributed on the surface
of a moving bubble. The surface concentration on the upstream part
of the bubble is less than equilibrium concentration, while that
one on the downstream part is greater than equilibrium. This
disequilibration of the concentration is caused by a viscous drag
of the medium acting on the interface and as a result of it there
arises a gradient of surface tension alongside of the bubble
surface. This gradient of the surface tension induces flow in the
direction opposite to that induced by the shear stress in the
outer liquid acting on the interface. It increases rigidity of the
interface and inhibits circulation of internal gas. When the
interface is completely rigid the terminal velocity of bubble
motion is reduced to that given by Stokes' law. According to
Levich Cl53 correction factor x to Stokes' law is:

3T * 3 7-/
9
+ y

at X
(5)
2 “L +3r)
9
+Y

where y is a coefficient of retardation and its value depends on


a character of transport of surfactant molecules to and alongside
of the interface.

Deryaguin et al. [16, 171 have analysed variations in surface


coverages between the top and rear part of the moving bubble. They
estimated that if << 10-'cm, where r is the equilibrium
ro'co 0
surface concentration and C is the equilibrium bulk
0
concentration, then the difference would be relatively small.
However, when >> 10w4cm then almost all adsorbed material
ro'co
is carried toward the rear of the bubble Cl61 and the surface
coverage at the top of the bubble can be considered as equal to
45

zero. This analysis of the concentration variation alongside

rising bubble was extended by Saville C211 who presented the

dependence of the dimensionless surface concentration a5 a


function of the angular position on the bubble surface. This
dependence for very weakly surface active substances is shown in

Fig.1 as eb = r/r0 = f(4) dependence, where cp is angular


position on the spherical bubble measured from its forward pole, r
is a surface coverage at a given position. It is seen there that

even for very weakly surface active substance the surface coverage
at the top of the rising bubble is a very small fraction of the
equilibrium coverage. Moreover, over much of the surface (up to +

@b
6

0 30 60 90 120 150 180


9

Fig.1. Surface concentration on the bubble surface as a function


of angular position.

t 1007 r is lower than r. ( eb<l 1. The surface concentration


rises sharply near the rear pole of the bubble.

Thus, we should be aware that as a result of phenomena


described above the bubble approaches solution/gas interface with
surface coverage at its top much smaller than the equilibrium one,
and this actual non-equilibrium surfactant concentration can be
decisive for foam formation and the foam films stability.
46

2.2. Role of the wall effects.

When a particle moves near any interface its motion is


influenced by additional particle-wall (interface) hydrodynamic
interactions C301. As a result of these interactions the particle
velocity near the interface can be slowed down. Magnitude of this
effect depends mainly on: i) the nature of the interface and the
particle, ii) the distance between the interface and the particle,
iii) size and shape of the interface, iv) the direction of the
particle motion relative to the interface orientation and the
direction of gravity.

All studies on foamability or bubble motion have been carried


out in containers of finite dimensions. In discussion of bubble
motion two wall effects should be considered: i) effect of
the container walls, parallel to the bubble motion, and ii) effect
of the solution/gas interface, perpendicular to the direction of
the bubble motion.

Effect of walls on particle motion is commonly expressed by a


resistance coefficient correction A, being a ratio of particle
velocity in infinite fluid (ubco) to its velocity in bounded
fluid:

%co
h = -
(6)
ub

The influence of walls of cylindrical tube on motion of bubbles


and drops was discussed in C31J. Clift et al. C313 presented a
dependence of the resistance coefficient correction h on the ratio
(S) of bubble diameter to the column diameter and showed that the
experimental results were well described by the following
relationship between h and S :

l/X = ( 1 - s2 j3’*

It is seen from Eq.7 that if s < 0.1 then h 9 1, i.e., effect


of the column walls on bubble velocity can be neglected. Moreover,
it was showed that these wall effects were diminishing with an
increase of Reynolds number (Re). They will have negligible
influence (less than 2%) on terminal velocities, under the

following conditions:
ReSO.1 , s 5 0.06
0.1 < Re 5 100 , S 5 0.00 + 0.02 loq Re
Re ? 100 , s 5 0.12
Usually in foamability measurements the ratio of bubbles diameters
to column diameter is smaller than 0.06 and therefore the influence
of the container walls on bubble velocity can be neglected. The
situation becomes more complicated when we have a large number of
flotatinq bubbles because some fraction of them will flow closer to
the container walls and therefore their motion can be affected in
some manner by the "wall effects". Nevertheless, one can conclude
that the container walls cannot slow down bubble in a degree
necessary to restore the equilibrium coverage over its whole
surface.
Motion of a spherical particle (solid or fluid) towards a flat
wall (interface), under creeping flow conditions (Re<O.l), is one
of a few cases for which there exists exact theoretical solutions
[30,32-331 confirmed experimentally [33-351. Under creeping flow
conditions the wall effect is significant and for example at the
distance H = 0.1 radius of the particle from a flat solid interface
the velocity of the solid particle is about 10 times lower and the
bubble velocity is about 5 times lower than their respective
velocities in unbounded liquid. Bartsch c331 have presented a
theoretically calculated relation between resistance coefficient
correction X and dimensionless distance l-lbetween a bubble and flat
free surface:

a
n (n+l) 2 sinh(2n+l)a + (Zn+l)sinh a
X = 2 sinh a '; -1
/_ (2n-1)(2n+3) I 2 cosh(2n+l)a - 2 cash 2a
n=i

(8)
where:
a = In (l-i+ 1 + ((l-i+ 1)' - 1):")
48

Equation (81 is valid only under creeping flow conditions and under
the assumption that the interface and/or bubble are not deformed,
i.e., their shape is not changed. It can be found from the relation
(8) that in the case of the liquid/gas interface the wall effect is
the smallest one, but nevertheless at the distance l-4 = 0.1 the
bubble velocity should be above 2 times smaller than in infinite

liquid.

During foam formation the bubbles float through the solution


with velocity strongly dependent on their dimensions, but as a
rule the Reynolds number characterizing their motion is much bigger
than 1. Unfortunately, at higher Reynolds numbers the amount of
information concerning effect of the interface on velocity of a
particle approaching it perpendicularly is very scarce c351.
Nevertheless, it was found c351 also in this case, that the
magnitude of the wall effects diminished with the increasing
Reynolds number. Fig.2 presents the dependence of the resistance
coefficient corrections A on Re for two selected distances l-l
between the sphere (solid lines) or the bubble (dashed lines) and

solid interface. It is seen there that the values of the resistance

w 4 30 Re
Fig.2. Changes of the resistance coefficient correction A with the
Reynolds number
49

coefficient corrections decrease very quickly with the increasing


Reynolds number. The changes are especially rapid in the range of
Reynolds numbers within 0.01 - 1. These changes are more pronounced
in the case of bubble than in the case of solid sphere approaching
the solid plate. It is seen in Fig.2 that at Reynolds number of an
order 3 the resistance coefficient corrections are approaching the
limiting value 1. There is a clear tendency to reach this limiting
value at higher Reynolds numbers. Thus, one can assume that at
higher Reynolds numbers, characteristic for floating bubbles, the
resistance coefficient correction is equal to one and we can
conclude that the bubble approaching the solution/foam interface is
not slowed down enough to re-establish the equilibrium coverage at
its top.

3. MEASUREMENTS

3.1. Foams

Majority of methods applied for measuring foam formation and


solution frothing properties consist in dispersing a gas in
liquid. Most commonly gas is injected into a solution tested
through a capillary, or capillaries, or a sintered glass [Jb-453.
However, other methods such as for example: it shaking up and down
a cylinder partly filled with the solution [46,49-511, ii) pouring
the solution from one into another vessel [3,521, iii) stirring,
whipping or beating the solution [3,53-553 are used. Methods
being a combination of these actions are also met C553. Details
of these methods are thoroughly described in C3, 563.

In dynamic, steady-state methods the foam height or the foam


volume is measured while the gas is injected with a conrjtant
velocity. Various experimental procedures have been proposed
[3,3&k-41, 431 in an attempt to obtain the most objective and
reproducible results. However, there is still a lot of unanswered
questions. Well-known and popular is a method proposed by Bikerman
t363. He aimed at establishing foaminess as a definite physical
property of a liquid, and defined E3, 363 a "unit of foaminess"
50

“f t “f
x =- =- (91
V u
9 9

where V f is the average foam volume, and V is the volume of the


g
gas forced through the column in time t. is a
ug = "glt
volumetric gas flow rate. It was stated C3, 361 that when volume
of the solution studied was so large that its magnitude ceased to
influence the ratio (Vf t/ Vg) and the gas velocity WZ45 not to

small and not to great, then 2X was independent of the gas flow
rate, of the shape and dimensions of the measuring tube and of the

average pore of the sintered glass used for gas dispersing. X is


equal to an average bubble lifetime in the foam if coalescence of
the bubbles inside the foam volume and volume of the liquid in the
foam can be neglected. The first of these assumptions seems to be
reasonable while the second one is usually not ful filled,
especially in the case of wet foams. For such systems the liquid
content in the foam column can be as high as 40 ~01% [57-591

Sun [601 introduced a frothability index (FI) recomended


later by Gaudin C613 as very useful for comparing foaminess of
different flotation frothers. The frothability index was defined
as a ratio of the foam volume produced from a solution of the
frother tested to the volume of foam produced from solution of
n-hexyl alcohol, which was chosen as a standard. However, the FI
values were determined for arbitrary chosen and identical (despite
differences in surface activities1 bulk concentrations and
therefore were not really connected with any physicochemical
properties of the system investigated. It can lead to false
conclusions.

A parameter called retention time rt was defined 1391 basing


on an observation that for every dependence of the total gas
volume V contained both in foam and solution, on the gas flow
9'
rate u there can be distinguished a linear part. Examples of such
9
dependences are presented in Figs. 3 - 5 for aqueous solutions of
n-octanol, n-heptanoic and n-nonanoic acid solutions,
respectively. The linear part of the plot can be easily noticed on
51

16C

20 40 60 U. bm’/hl

Fig.3. Dependence5 of the total gas volume Vg, contained both in


solution and foam, on the gas flow rate for n-octanol
rjolutions of various concentrations

r-

. 54-5 M
0 MO-( M
d 2.10-’ ”
n-heptcnols oni
c! 3.10-’ t.,
x 5.10+ M
. 1.10-3 M

121

IM

6(

Fig.4. Dependence of the total gas volume Vg, contained both in


solution and foam , on the gas flow rate for solutions of
n-heptanoic acid of various concentration
52

l-

l-

l-

I
3

Fig.5. Dependence of the total gas volume Vg, contained both in


solution and foam, on the gas flow rate for solution of
n-nonanoic acid of various concentration

each of these curves. Thus (c.f. Figs. 3 - 5):

AV
a
rt = (10)
Au
g

The rt values, similarly as Bikerman's unit of foaminess, are


independent of the gas flow rate, and of the dimensions of the
measuring column Morever, they have a clear physical meaning.
Physically, the rt is equal to the average time necessary for a
unit gas volume to pass through the system, that is, solution +
foam. Indeed, the total gas volume in the system is equal to the
product of the average values of the gas bubble volume (Vb)'
number of bubbles generated in a unit of time (nb), and an average
lifetime tb of a single bubble in the system. Since the gas flow
rate u = v we have:
g b nb'

AV = (n;; vi - n: vi) tb (11)


9
and
Au = n; v;: - ni v; (12)
g

Dividing Eq.11 by Eq.12 and comparing with relation 10 we can see


immediately that rt = t
b
53

A relation between Biker-man's unit of foaminess and the

parameter rt can also be easily shown. The total gas volume

contained in the system is a sum of its contents in foam V* and in


9
the solution VS. Thus,
9

tv;,” + (v;,*n - w;,9 - w;,


rt = (13)
u'I - u)
s s

Volume of the solution is constant but its distribution between


foam and liquid phases changes. Nevertheless at any time V = V>
g
VS, vt= us'+ v; and v? v: = const, where (and Vr are volumes
g
of the solution contained in foam and liquid phases, and V is
L
volume of the liquid phase (solution + bubbles floating in the
solution). Using these relations the Eq.13 can be written as
follows:

(V,)“- (Vf)’ (V
I
)“- (V,)’
rt = + (14)
u-l - u' u" _ u'
s g g s

According to Bikerman the Z parameter is independent on gas flow


rate ,i.e. (Vf)"/ u; = (V,)'/ u' = z , and we can obtain from Eq.14
s
that

1 A vt
u I’ U' A “1
g g
rt = --- E:+- = r+- (15)
Au
g
Au
9 I Au
g
where AU 1. is a variation in volume of the liquid phase (solution +
bubbles) with a change of the gas flow rate. Thus, the rt value is
a sum of the Bikerman's parameter X and a magnitude AVL/Au showing
g
changes in volume of the phase:(solution + floating bubbles) and it
is free from the unrealistic assumption that the solution content
in the foam is zero.

Recently Ross and Suzin C431 considered advantage of using a


conical foam vessel. However, finally they concluded that there is
54

no “right answer". According to them a foam test should produce


foam and measure its stability under conditions similar to those of
its ultimate application. The shape of the container is one of the
significant condition.

3.2. Single foam films

Single foam films have been an object of great interest not


only because they can be considered as a simplest model of a foam
unit but also because they are a convenient model for studies of
forces of interactions in the dispersed systems. Nevertheless One

should remember always that a foam system is not a simple sum of


the foam films. In majority cases the single foam films are
investigated under static conditions and they are carefully
isolated from any external disturbances. Such situation never
occurs in foam. Even in the case of the metastable dry foam, where
we can assume that syneresis is negligible, there are still some
processes going on such as for example: gas diffusion,rupture
of films, rearrangements in foam structure, etc. In the case of
wet foams, which can exist practically only under dynamic
conditions, we never have a situation in which the foam films are
isolated from external disturbances. There occur permanently flow
of bubbles upwards and the solution downwards, rupture of bubbles
in the top foam layer, droplets of solution from ruptured film5
fall on the top bubbles, pressure shocks, bubbles rearrangements,
etc. However, despite these complicating factors the properties of
single films are crucial for formation and stability of the foam
column. It should be stressed here, however, that we have in mind
properties of foam films under conditions identical to conditions
under which the foam is formed and exists. Quite often these
dynamic properties of the foam films can be quite different from
the equilibrium ones. Unfortunately number of information
concerning dynamic properties of single foam films under dynamic
conditions is rather limited [53,62,631.

Methods of investigations of properties of single foam film5


can be divided into 2 groups depending on lateral dimensions of
56

the films. There are methods for studying the microscopic [64-693
and the macroscopic [70-721 foam films. Both groups apply
interferrometric methods to measure the film thicknesses. P review
of these methods was given in C661 and recently in C561.
Investigations on macroscopic foam films can be carried out only
in the systems producing very stable films and foams. Methods
applying microscopic films are much more universal. Microscopic
circular foam films are formed when two bubbles touch C641 or a

part of solution is sucked off from a biconcave drop [65,661. The


method elaborated by Scheludko and Exerowa C651 is especially
useful and widely used. It is the only method which
enables determination properties of single foam films formed from
solutions of substances with relatively low surface activity.

Stability of the equilibrium foam films is described in terms


of DLVO theory c73,741. Deryaguin C751 postulated that the
properties of liquid in the thin film are different from those of
the bulk phase and introduced a concept of the disjoining pressure
as a measure of the corresponding change in the thermodynamic
properties. The increase of the chemical potential of the thin
film ph in relation to the chemical potential of the bulk phase ~1,
is equal to:

Ap (h) = /J,,
-p, = - n v (161

where n denotes disjoining pressure and v is a volume per


molecule. The sign of the disjoining pressure is chosen so that n
> 0 means acting towards disjoining of the film surfaces, i.e.
towards thickening of the film. Plccording to the classical DLVO
theory the stability of an equilibrium foam film is determined by
forces arising from deformation of diffuse electric double layer
and the disjoining pressure may be represented as a sum:

f?=?T yw(h) + noL(hl (17)

where nw(h) is the component representing van der Waals


intermolecular attraction forces, and n is the electric
01
component. In the generalized DLVO theory C76,771 there are
56

considered also other components of the disjoining prersure such


as structural force5 and steric forces. However, these components
are discussed rather in relation to stability of thin liquid
films between two solid surfaces.

The DLVO theory applies very well to normal thin liquid films
but in the case of common black and especially Newton black films
there is a lot of discrepances C783. It should be also mentioned
that forces of disjoining pressure (electrical component) start to
be of meaningful magnitude at film thickness of an order of 1oooa
and lower. In the case of thicker films other factors, such a5 for
example surface elasticity or viscosity which affect the kinetics
of film drainage, have to determine their lifetime and stability.

First experimental determination of the Gibbs elasticity in


foam films were performed by Mysels and co-workers [793. The Gibbs
elasticity of a thin film is defined as the ratio of the increase
in film tension (Zdd) resulting from an infinitesimal increase in

area, and the relative increment of the area (dlnA):

2 dd
Eo = (18)
dln A

It is assumed in the concept of the Gibbs elasticity that locally


an equilibrium between surface and interior of a film element i5
re-established quickly, i.e. in a time shorter than time scale of
the film disturbance. If the film i5 thin then the total amount of
surfactant present in the film element is insufficient to restore
the original concentration at the surface and in the film interior
liquid. This deficit give5 rise to an excess surface tension in the
extended region of the film. This excess surface tension is a
measure for the Gibbs elasticity.

Mysels and co-workers measured the Gibbs elasticity of a thin


liquid film extended on a glass frame and being in contact with
solution volume. The method involved simultaneous determination of
the change in the surface tension acting upon a film under
observation and of the motion of the interference fringes which
this produced. They did not observed the expected from theory
[80,81] dependence of the Gibbs elasticity on film thickness and
composition. Prins et al. C82,511 using an improved technique

showed that the measured Gibbs elasticity was dependent on the

film composition and decreased with increasing thickness of the


film element as predicted by the theory. Further quantitative
verification of the Gibbs elasticity was done in works of Krotov
and Rusanov [83-851 where a method developed for horizontal

isolated films was used. Works dealing with the Gibbs elasticity

of free liquid films have been reviewed by Lucassen C21 and by

Krotov and Rusanov C861.

We would like,however, to point out here that in all these


works there was an equilibrium coverage at the film surface at the
beginning of the experiment . Therefore, these results cannot be

related in a straightforward manner to properties of dynamic foams


where an existance of an equilibrium coverage at foam film surface
is not obvious.

4. COMPOSITION OF THE WET FOC\M

4.1 Water content within the foam height

Wet foams are an assembley of spherical bubbles separated by


thick lamellae.The bubbles move constantly upwards to the top foam
layer. Direct observations and results of measurements of solution
content indicate that rupture of the film takes place practically
only in the top layer of the wet foam and coalescence in volume of
the foam can be neglected. Solution content in the wet foams was
measured by a direct method described elsewhere C57,871. The method
consists in cutting-off under steady state condition a chosen
volume of the foam using the column of a special construction

[57,871. The column upper and lower parts are attached to two well
polished plates and fastened together by a special holder. The
holder squeezes the plates together to asssure the water-tightness.
The principle of action of the column is illustrated in Fig.6. The
upper part of the column can be shifted from position I to position
II (Fig.61 and the volume of solution from the cut-off foam layer
is collected and measured.
58

Fig.6. Principle of action of the column

Results of measurements of the liquid content (p in the foam


as a function of the gas flow rate u are presented in Fig.7. for
g
n-amyl alcohol solutions of concentration 1~10-~11.

p = - v: (19)
vY

where V is the volume of solution contained in the cut-off foam


0

Fig.7. Dependence of the liquid content cp in foam on the gas flow


rate. Line 1 - in the whole column, line 2 - in the top
layer of thickness 1421 mm.
59

&P_ AmOH. Zl6'IMl


L %I

LO-

30-
0 n
LI-
20. 0 -0 -0
o-
a ”
‘2 I\
IO -
h
Y 2 A
0 ,""...,'.""'*'.Q.~.",,A
IO 15 20 25 30 Ug Cl/h1

Fig.8. Dependence of the liquid content cp in foam on the gas flow


rate.Line 1 - in the whole column, line 2 - in the top
layer of thickness 1421 mm

layer and VI is the volume of the cut-off foam. Line 1 presents


the p = f(ugl dependence for the whole foam layer formed under
these conditions. It should be added that the term "whole foam
layer" used here,denotes the foam layer cut-off always 521 mm
above the solution/foam interface. Line 2 presents the Q = f (ug)

dependence for top foam layer of thickness 14+1 mm. It is seen in


Fig.7. that the liquid content in the top foam layer is
significantly lower then the average for the whole foam column.
Moreover, the average liquid content in foam increases with
increasing gas flow rate while for the top layer the cp values
initially decrease slowly and then reach a constant level of 5% . p
= f(ugl dependence5 for foams formed from n-amyl alcohol so lutions
of concentration 2x10-'M are shown in Fig.8. Here also the liquid
content in the top layer is much smaller then the average va lue for
the whole foam column.

Various heights of the foam column obtained under a chosen


steady state conditions can be cut-off in the experiments. In such
way we can obtain a distribution of the liquid as a function of

23
60

the height of the foam layer . The dependence of p on height of the


cut-off foam layer formed on n-amyl alcohol solutions of
concentration 1x10-=11 at us= 401/h is shown in Fig.9. It is
clearly seen there that the Q values decrease towards the top of
the foam column.

Fig.9.
The liquid content cp as
a function of height of
the cut-off foam layer.

The dependences of liquid content in foam on concentration of


n-amyl alcohol solutions are shown in Fig.10. for the gas flow
rate u = 40 l/h. It is seen that the average liquid content in the
s
whole foam initially increases quickly with increasing solution
concentration and then for solution concentrations c > 5~10-~M
reaches a plateau value of +z.r 32X . The liquid content in the

AmOH +,=&01/h
9
C%l -

IO-
' Qp, 2

0' 1 2 C.102CM7

Fig.10. Dependence of the liquid content cp on concentration of


n-amyl alcohol solutions. Line 1 - in the whole foam
column, line 2- in the top layer of thickness 14?1 mm.
61

top (14 t lmm) foam layer is 5% and is practically independent on

the solution concentration.

These results clearly show that the foam films are the
thinnest in the top foam layer. Thus, it indicates that rupture of
the wet dynamic foams takes place at the top layer of the bubbles

where the thicknesses of the thin liouid films are the smallest.

4.1. Thickness of rupture of the foam films

Rccordinq to de Vries [WI an average thickness hf of


lamellae in foam layer can be calculated as:

= 2 v: <r2>/(3 Vg<t->l
hf
where V is the volume of gas contained in the foam, and <r> is an
g
average bubbles radius. Taking into account that

v =Vf-Vf *
and p=Vf/V
D f
g

the Eq.20 can be rearranged into the following:

hf = 2 p < r2 >/C3 Cl-cpl < r >I (21)

On the basis of distribution of liquid content p within the


height of the foam column formed from n-amyl alcohol solutions and
assuming (on the basis of the results given in C401) that r = 0.6

mm we can find from Eq.21 that for the foam layer of thickness
6 mm, i.e. for about 5 layers of bubbles, the average thickness of
the foam lamella would be of an order of 15 pm. It indicates that
the thickness of rupture of thin liquid films in the top layer of
bubbles should be below 15pm, because the liquid content decreases

towards the top of the foam.

Second estimation of thicknesses of the films rupture can be


made on the basis of lifetime of bubbles in the top foam layer.

The averaqe lifetimes of bubbles in the top layer of foam formed

from solutions of n-amy 1, n-octy 1 and n-nony 1 alcohols were

measured using a movie camera C40,893 and the dependences of T on


62

the solutions concentration are shown in Fig.11. It is seen there

that initialy the average bubble lifetime increases rapidly with


concentration, but at higher concentrations reaches a constant

value. The r values were of the order of 0.11 s for n-pentanol,

0
I o.opo2 ago04 0.4606 -(Ml,
003 C&
0.01 a02

Fig.11. Average lifetime T of gas bubbles in the top foam


layer as a function of concentration of n-pentanol,
n-octanol and n-nonanol solutions. Note the big difference
in the concentration scales.

0.13 s for n-octanol, and 0.09 s for n-nonanol at the highest

concentrations of their solution (Fig.ll.1. Thus, the differences

between the values of the average lifetime T for various n-alkanol

solutions were not large. It is remarkable, however, that the r

values for n-nonanol were the lowest, and it coincides (as it will

be shown further) with the lower frothability of n-nonanol

solutions. It was shown in C401 that lifetime of single bubble was

of an order of a few seconds, while drainage time of single


microscopic film formed from n-amyl alcohol solutions w.as of an

order of 100 seconds. Simultaneously thickness of rupture of the

single microscopic film was below 100 nm, while in the case of
single bubbles it was estimated to be above 300 nm c401. The

average lifetimesof bubbles in the top foam layer were by orders


of magnitude lower than the observed for single bubbles and

microscopic films. It indicates that thin liquid films in the top


foam layer rupture at thicknesses at least of an order of 1 pm or
63

larger.

Thus, on the basis of the both estimations we can say that


the most probable thickness of rupture of the thin liquid films in
wet dynamic foams is within a range lpm-lOpm, i.e.,within the
thickness range where the forces of disjoining pressure do not

start to operate, yet.

4.3. The foam model

In the previous chapters various processes occuring during

bubble formation, motion and aproach to the solution/foam


interface and in the foam column were discussed. We tried to
underline the dynamic character of the processes occuring there
and to show that generally in the case of the wet foams there is
not enough time to estabilish equilibrium coverages at surfaces
of the foam films, and therefore an actual non-equilibrium surface
coverages should be considered in discussion of mechanism of the
foam stability.

Wet foams exist practically only during an action of the foam


forming force. When a column of the wet foam is under steady state
conditions then at any chosen level in the foam volume an average
number of bubbles is constant and amount of the draining solution
is equal to the amount of the solution brought up by bubbles
entering this level. Measurements of the solution content show that
in the foam volume the films separating the bubbles are very thick
and therefore probability of the bubbles coalescence in the foam
volume is negligibly small. Foam films formed over the top layer of
bubbles are under quite different conditions. They are in contact
with an ambient atmosphere and are not supplied by solution
syneresing from above. Therefore, the forces acting on these film5
are balanced to far lower degree than inside the foam volume. The
top foam films are subjected to various external disturbances such

as pressure shocks caused by rupture of the neighbouring films,

fall of aerosol drops from a “fountain” formed from rupturing

films, movements and rearrangements caused by bubbles entering the

layer, solution drainage, etc. Their time of survival depends on


64

magnitude of the stabilizing forces counteracting all these


external disturbances. The durability of these foam films, i.e.
their ability to induce forces counteracting the external
disturbances, determines stability of the whole foam system.

Thus, in our model of the wet foam there are three essential
points: il rupture of the foam takes place only in the layer of top
foams films, ii1 durability of these top foam films determines
stability of the whole foam column, and iii) it is a
non-equilibrium system and phenomena which occured in the solution
phase affect also properties of the foam.

5. FORCES STABILIZING THE FOAM

Specific forces of interaction (disjoining pressure),


viscosity and/or surface elasticity forces are commonly referred
to in a discussion of foam stability. Usually, there are attempts
to explain stability of various foam systems in terms of one of

these factors. However, one should be aware that for example


forces of the disjoining pressure attain a meaningful value only

at film thickness below 1000 8. This means that attempts to


explain the stability of wet spherical foams in terms of
disjoining pressure are doomed to failure because these foams
rupture at greater thicknesses. On the other hand, surface
elasticity forces are induced and operate only under dynamic
conditions. Thus, in the case of "static" foams they can be the
determining factor at the stage of the foam formation, but later
when one deals with well-drained quasi-static polyhedral foam then

the surface elasticity forces can become of only secondary


importance. In such systems the foam films reach thickness much
below 1000 8, common or even Newton black films are formed [73, and
the forces of disjoining pressure can be there the major
stabilizing force. Thus, the conditions of existence should be
always carefully specified in discussion of the foam stability.
65

5.1. Surface elasticity forces

Surface elasticity forces, arising from variations in surface

tension during deformation of the interface, can be induced not


only in foam film5 but also on surface of a semi-infinitive

solution and for single interface we have:

dd
E = (221
dln CI

In the most general case C90-931 the viscoelastic properties

of fluid interfaces can be characterized by a combination of


"compositional" and "structural" contributions. The former is

determined by the extent and rate of surface dilation and shows the

elastic and viscous resistance against changes in area. The surface

dilational modulus E is a measure of the resistance against the


creation of surface tension gradients, and of rate at which such

gradients disappear once system is again left to itself. Its


magnitude is a strong function of frequency of external
disturbances and at sufficiently high deformation frequencies, when

surface layer behaves as insoluble monolayer the surface elasticity


forces are at their highest and are called the Marangoni dilational
modulus Ey. Any relaxation process taking place in near interface
(for example diffusional exchange) causes surface bahaviour to be
viscoelastic and therefore at lower frequencies the modulus E has

both elastic and viscous components. The structural part arises


from treating interfaces as autonoumous phases characterized by
their own rheoloqical properties. It is linked with the extent and
rate of shear and shows the elastic and viscous resistance against
changes in shape of surf ace elements. Results of many studies
C90-921 show that the resistance of the surfaces against shearing
motion is much smaller than against compresion/dilation and the
dilational properties can be of several orders of magnitude higher
than the corresponding shear properties of the same surface. In
molecular terms it can be explained that it is easier to make

molecules at the surface to move past one another at a constant

value of their average distance than it is to move them apart and

increase their mutual distance.


66

The surface dilational modulus E is a complex number in the


general case of both elastic and viscous contributions:

E = E' + E"i = E' + i o T)~ (23)

Here E' is the real part of the elasticity modulus (storage


modulus), and E" = w r)d is the imaginary part (loss modulus) that
characterizes energy dissipation due to irreversible transport
processes and is proportional to the surface dilational viscosity
nb. Longitudinal wave technique [94-96,911 and the pulsating
bubble method C97-993 has been used to determine values of the
surface elasticity forces in various systems.

Below some results obtained by the method of pulsating bubble


are presented. In the method of pulsating bubble a small bubble
formed on a tip of capillary dipped into a solution is forced by a
special electromechanical system to radial harmonic pulsations of
definite amplitude. The force necessery for obtaining the required
pulsations is measured in function of pulsation frequency and
solution concentration. The magnitude of this force depends on the
elasticity forces induced on the bubble surface as a result5 of
its dilation.

1
T
2 ao-
22
IiF
60-

LO-

20-

0 I I 1
-6 -5 -1, -3
1% c 0 [mo'e"l

Fig.12. Marangoni dilational modulus E, of n-heptanoic and


n-nonanoic acids measured by the pulsating bubble method
(points) and calculated theoretically from Eq.25 (lines)
67

In Fig.12 the measured value5 of the Marangoni dilational

modulus EM for n-heptanoic and n-nonanoic acids are presented. The

point5 show value5 obtained from measurement5 Cl001 by the

pulsating bubble method:

1
dlnr C(flvport dH dlnr 4rrf " RTr dH
E,,,= - Aa-Ado + _ _
dlnA 2VoAH dr dlnA D
I [I
(241

where:A,H,r o denote the surface area, height of the bubble cap and
radius of its curvature at the pole, V. is a constant of the
pulsating bubble apparatus (internal gas volume), x = cP/c is the
V
adiabatic constant,po is the pressure inside the bubble, C(f) is an
apparatus constant depending on frequency f = o/Zn only,AH is the
amplitude of pulsation at the bubble pole, a is a constant of the
differential equation describing the pulsation of the bubble Aa =
as- a is the difference of constants a for solution and solvent,
W
is the difference between surface tensions of solvent and
A"O
solution, R, T, and D are the gas constant, absolute temperature
and diffusion coefficient, c 19 and T)~ are parameters (dimension of
concentration) depending on adsorption kinetics. The lines are the
values calculated theoretically by applying the Langmuir isotherm:

where r is an equilibrium surface coverage. Within the


0
concentration intervals applied there is a good agreement between
the elasticity isotherms resulting from the oscillating bubble
measurements and. those obtained from the equilibrium surface
tension-concentration isotherm.

The Marangoni dilational elasticity modulus EY represents the


highest possible value of the surface elasticity force5 for the
given system. This value is obtained under the conditions that the
equilibrium surface coverage is estabilished at the surface at the
moment when the dilation of the interface starts and that the
68

disturbance is so rapid that the adsorption layer behaves as an


insoluble monolayer (no diffusional exchange).

However, as it was discussed in the previous chapters, the


real dynamic foam is not a system in which the adsorption
equilibrium is fully estabilished everywhere and at everytime.
Therefore, when the foam lamella is under any external force
tending to destroy it then the forces opposing disruption, i.e.,
stabilizing the foam, are not connected with the equilibrium
surface coverage but with the actual non-equilibrium state of the
surface. This effective elasticity force Eeff induced in a foam

lamella cannot be larger than EY and usually is always lower than


the Marangoni dilational modulus because the surface cover-ages
are lower. Values of the Eeff are strongly dependent on adsorption
time and frequency of disturbances. They were calculated for
homologous series of n-alkanols Cl011 and fatty acids Cl001 under
the following assumptions: i) the mean lifetime of a bubble in the
top foam layer is the time available for the surfactant molecules
to be adsorbed at the surface of the foam film (tad]; ii) the
surface coverage r(tad) at the time tad controlles the elasticity
value and can be calculated using diffusion controlled adsorption
kinetics C1021; iii) owing to external disturbances oscillations of
the bubble surface arise and there is superposition of adsorption
and exchange of matter at the bubble surface; iv) the exchange of
matter in non-equilibrium state of adsorption r(tad) < r. can be
calculated in the same way as in the equilibrium adsorption state
following the adequate theory C991. To calculate the value of the
effective surface elasticity it is assumed that the dilational
elasticity, acting at the bubble surface at a surface coverage
lower than equilibrium adsorption, can be calculated in the same
way as in the equilibrium. After the time tad the surface coverage
has reached the value r(tad). The r(tad) values were calculated
using a new derived solution of the diffusion controlled
adsorption kinetics, in a form of power series developed on the
basis of an orthogonal collocation method C1021:
69

r(e) =ro;ai i=l


11 -
c1+
1

8 + &I2 co/a 3 l/2


(26)

with
8 = ( Dt / K2 1"' (27)

and

K = ro / co = roe / ko + aL) (28)

where a L is a concentration of half surface coverage (parameter of


the Langmuir isotherm). The coefficients ai are tabulated for the

concentration range 0 < co / aL < 10 for M = 3. This solution was

proved to be as accurate as all other procedures available

including the direct numeric solution of the diffusion problem.

With respect to the Langmuir isotherm we obtain the dilational

elasticity as

) = R T I-@)r(tad)/ roe - r(tad) (29)


E(tad c 3

This is a maximum value of the elasticity forces at a time tad


which can be induced only under the condition that the dilation of
the interface at the time tad is so rapid that there is no

diffusional exchange of matter. However, the disturbances are of a


definite frequency f and therefore the elasticity forces really

induced and operating are lower. Their values depend on the


frequency of the interfacial oscillations. Here, it is assumed that
the average period of external disturbances responsible for the

rupture of the foam lamellae can formally be correlated with the


frequency of oscillations in the pulsating bubble method. As a
result of a definite frequency of oscillation this maximum value
1 is reduced by a factor C991:
E(tad

F Cw> oJ2a1 + *1wI


= (30)
F co> a:+ (al + cIl~)2
70

with

a = (o /2Dl*"
t
and

cli = roe aL / CD(a, + ~~‘1 (311

This leads finally to the values of the effective elasticity forces


acting at the foam lamella after an adsorption time tad under a

disturbance of the frequency f

E = E(tad) p - F(U) / F (O,] (321


eff

Results of the calculations are presented in Figs.13-15 . The


calculations were performed using values of ra, = 0.3 x
-10 -6
10 mole/cm' and D = 5 x 10 cm’/s. In Fig.13 there is presented
an effect of the adsorption time on values of the effective
elasticity for n-heptanoic and n-nonanoic acid solutions.The Eeff

n-nononoico.
I a
30-
t 0

20- n-heptonoic 0.

0 1 I I I I I

0 1 2 3 4 t5
ads. [sl

Fig.13.Dependence of the effective elasticity values 0” the


adsorption time for n-heptanoic and nonanoic acid solution5
at concentrations co/a = 2.
L
71

= f(tad ) dependence5 shown in Fig.13 were calculated for identical

values of the reduced concentration c /a of the both acids. a is


0 l. L
the concentration of half surface coverage, i.e., concentration at
which the surface concentration of the appropriate substance
reaches half of its maximum value C The reduced concentration
ao'
enables to compare properties of substances with very different

surface activities on a unique scale. It is clearly seen in Fig.13

that longer adsorption time tad makes possible to induce larger


effective elasticity forces. Moreover, at identical values of the
reduced concentration i.e. at comparable conditions of the surfaces
in equilibrium, the E values are quite different for both acids.
eff
This results from differences in their adsorption times due to
differences in their surface activity. The more surface active the
substance is (n-nonanoic acid) the longer is the time necessary to
obtain a given surface coverage.

For homologous series of fatty acids (Cs- Cio) the influence


of adsorption time on the calculated values of the effective
elasticity forces is shown in Fig.14 for a constant frequency f =
2000 Hz and reduced concentration co/a = 2.0. fk one might expect
I.,

25.

20-

15-

10

Fig.14. Influence of the adsorption time on values of the effective


elasticity forces (E ) for various fatty acids at a
reduced concentration?'; /a = 2 and a frequency of external
disturbances f = 2000 H?. L
72

for the lower (n = 5 or 6) members of the homologeous series of


fatty acids, there is practically no influence of adsorption time
within the range tads = 0.02 - 0.64 s, i.e. within a range of
lifetime5 of bubbles in the top layer of wet foam, on the
calculated E values. Under these conditions the most significant
eff
effect of the adsorption time is observed for n-octanoic and
n-nonanoic acids solutions.

The influence of the frequency of external disturbances on the


induced values of the effective elasticity is shown in Fig.15 for
the frequency range 200 - 25 600 Hz. Although the frequency range
changes by two orders of magnitude a maximum is always seen for
every Eeff = f(n) dependence. The magnitude of the maximum depends
on the frequency.

+$=10 f=
1- 2OORz
t,,,:0155 2- 4OOHz
Gf 3- EOOHz
ImN/m.1. 4- 1600Hz
5- 3200%
G- 6200Hz
7- 12800Hz
o- 25GOOliZ

20

15

10

'1 5 6 7 a 9 10 n

Fiq.15. Chanqes in the effective elasticity induced at various


frequences of external disturbances as a function of the
number (n) of carbon atoms in the fatty acid molecule

Summarizing the influence of various parameters on the


effective elasticity force5, which potentially can be induced in
various systems, one can see that increasing the adsorption time
affects mainly the higher members of the fatty acids homofogous
73

series without having practically any influence on the lower ones.


Frequency of external disturbances has a quite reverse influence.
CIs a result of these competing effects the highest values of the
effective elasticity are observed for homologs with medium chain
length, i.e., for n = 7 or 8.

5.2. 64 relation between the foam stability and surface elasticity


forces.

The dependence5 of the retention time value5 on the


concentrations of n-heptanoic and n-nonanoic acids are shown in
Fig.16. As it was shown in Fig.4. and 5., the slopes of V = f (u )
g
dependences were lower in the case of n-nonanoic acid than in tghe
case of n-heptanoic acid. Therefore, the rt values for n-nonanoic
acid solutions are lower (Fig.16.) despite it5 higher surface
activity under equilibrium conditions. Note, please the different
concentration scales used in Fig.16. because of differences in the

surface activities of the two acids. The retention time values for

Fig.16. Dependence-s of the retention time values on the


concentration of n- heptanoic and n-nonanoic acid solutions.
homologous series of fatty acids f from n-pentanoic tc? n-decanoic

acid 1 are presented in Fig.17. as a function of reduced

concentration co/a L. Compariflg the foamability of Va!-ii?JU+ fatty

acids (Fig.171 it is seen that initially their foamability


increases with increasing chain length (up to Cal, but then it
rapidly falls for n-nonanoic and n-decanoic acids. In Fig.18 the rt
values for homologous series of n-alkanois (n-butanol to n-decanol)
are also presented as a function of their reduced concGtration
cofaL E1013. Here also one can see that initially the retention
time increases with increasing chain length, then it passes through
a maximum and subsequently rapidly decreases for n-nonanof. and
n-decanol solutions. Thus, for the both homologous series of
substances we observe a maximum on a dependence of foamability vs.
chain length despite that surface activity increases with chain
length in a monotonic way.

0 2 3 4
%L

Fig.17. Dependence of the retention time, rt, on the reduced


concentration of various fatty acids <Cs- C*o)
75

Fig.18. Dependence5 of the retention time, rt, on the reduced

concentration of various n-alkanols (Cd- CiO)

In Figs.19 and 20 the experimentally determined rt=f(n) and the


caiculated E ,ff=f(n) dependence5 are compared for solution5 of
n-alkanols and fatty acids, respectively. It is seen there that the
both dependence5 reveal maxima for n=4 - 8. The magnitudes and
positions of the maxima are a function of the concentration. The
agreement between the courses of the dependence5 of the foamability
parameter, rt, and the effective elasticity forces as a function of
the number n of carbon atoms in molecule of n-alkohol or fatty acid
shows that the effective elasticity forces, i.e. the elasticity
forces which are induced at an actual non-equilibrium surface
covet-ages, are the decisive factor in the formation and stability
of the wet foams under dynamic conditions. It also explains why the
lowest foam stability is observed- for n-nonenol and n-decanol in
the case of n-alkohols homologous series and for n-nonanoic and
n-decanoic acids in the case of fatty acids homologous series,
despite their highest surface activities. Foam stability,
I I I I f 8 I

bs' 0.16s
f -2048Hz
c=
r:i
40

30

20

,I0

I I I I 0
4 6 8 IO n

Fig.19. komparison of the dependences of rt and E on the number


n of carbon atoms in the n-alkanol mole?& i e for various
solution concentration.

rt
lS1

9 5 6 7 0 9nKl 5 6 7 8 9n

Fig.20. Comparison of the dependences of rt and Eof on the number


n of carbon atoms in the fatty acid molecu i e for various
solution concentration.
especially under dynamic conditions, cannot simply be correlated
with the surfactant surface activity. Surface activity is a
parameter characterizing properties of the solution/gas interface
being under equilibrium condition while the foam, and especially
the wet foam, is a dynamic system.

CONCLUDING REMPlRKS

Foam is very often treated as a completely independent phase.


While this approach can be justified in the case of dry metastable
foams, it becomes less and less accurate when the the foam becomes
less and less stable, and it is certainly far from reality in the
case of the wet foams. Properties of the wet foams should be
discussed in connection with their "history", i.e. in connection
with phenomena occurring in the solution from which the foam has
been obtained and conditions under which the foam has been formed
and exists.

Overall picture of phenomena occuring during formation and


existence of the wet dynamic foams is presented. Their importance
and mutual dependence5 and connections are discussed in the paper.
We have especially tried to stress dynamic non-equilibrium
character of the processes. Lack of the adsorption equilibrium has
its consequences in the magnitude of the stabilizing forces which
can operate during the foam formation and existence. Analysis of a
motion of a bubble through a solution shows that the adsorbed
solute is not uniformly distributed along the surface of the
moving bubble. This disequilibration of the surface concentration
is caused by a viscous drag of the medium acting on the interface.
The surface concentration on the upstream part of the bubble is
smaller than the equilibrium concentration, while that on the
downstream part is greater than the equilibrium. When the bubble
with such disequilibrated surface concentration arrives near the
solution/gas or solution/foam interface then, in principle, its
78

velocity could be lowered as a result of the bubble hydrodynamic


interactions with the interface ("wall effects"). Such lowering of
the bubble velocity could enable re-establishing the equilibrium
coverage at its top. However, it was shown that these wall effects,
which are very significant under creeping flow conditions, i.e. for
Reynolds number Re<O.Ol, disappear with increasing Re number and
can be neglected at Reynolds number characteristic for the floating
bubbles. Therefore, a bubble approaching solution/gas or
solution/foam interface is not slowed down and the equilibrium
coverage at its top cannot be re-established. In a dynamic wet foam
a pernament motion of the bubbles in the foam column takes place
and therefore the processes described above do not cease to occur

there. As a result of this, when the foam lamellae is under any


external force tending to destroy it, the forces stabilizing the
foam are not connected with the equilibrium surface CoveraQe but
with the actual non-equilibrium state of the surface.

It was shown that in the case of wet foams a rupture of the


foam films take place practically only at the top layer of the
foam. Thickness of rupture of these foam films is of an order of a
few micrometers. At such thicknesses forces of disjoining presusure
do not start to operate yet and it is one of the indications that

the surface elasticity forces are the main factor determining


stability of the wet foams. However, magnitude of these surface
elasticity forces is determined by the actual non-equilibrium
values of the surface coverages, not by the equilibrium quantities.
When a system is not equilibrated then its properties should not be
described in terms of equilibrium quantities - it can lead to false
conclusions. The effective elastcity forces EeffS broadly discussed
in th paper, are connected with the actual non-equilibrium state of
surface and are strongly dependent on adsorption kinetics and
frequency of external disturbances tending to destroy the foam
film.

Foamability and mechanism of stability of the wet foams is


discussed in details on the example of the foams formed from
solutions of two homologous series of surfactants; the fatty acids
and n-alkanols. It is shown that stability of these foams can be
79

explained in terms of the effective elasticity forces. There is a


correlation between changes of the foamability parameter (retention

time, rt) and the effective elasticity forces as a function of

number n of carbon atoms in molecule of the fatty acid and

n-alkanol. This shows that the effective elasticity forces, i.e.

the elasticity force5 which are induced at an actual

non-equilibrium surface coverage, are the decisive factor in

formation and stability of the wet foams under dynamic conditions.

It also explains why the lowest foam stability is observed for

n-nonanoic and n-decanoic acids in the case of the fatty acids

homologous series, and for n-nonanol and n-decanol in the case of


the n-alkanol homologous series, despite their highest surface

activities. Foam stability, especially under dynamic conditions,


cannot be correlated with the surfactant surface activity, which is

the parameter characterizing properties of the solution/gas


interface being under equilibrium conditions.

REFERENCES:

1. J.A.Kitchener, Chemistry Industry, 54 (1965).


2. J.Lucassen, Surfactant Science Series, Marcel Dekker, Basel,
1981, ~01.11, chap.6.
3. J.J.Bikerman, "Foams", Springer-Verlag, Berlin, Heidelberg, New
York, 1973.
4. S.Ross, I. D.Morrison, "Colloidal Systems and Interfaces", John
Wiley 81 Sons, New York, Chichester, Brisbane, Toronto,
Singapore, 1988.
5. M.J.Rosen, "Surfactants and Interfacial Phenomena", John Wiley
& Sons, New York, Chichester, Brisbane, Toronto, 1978.
6. J.A.Kitchener, C.F.Cooper, Quart. Revs., 13 (19591 71.
7. Khr.Khristov, K.Ma%ysa, D.Exerowa, Colloids Surfaces,
11 (19841 39.
8. W.D.Harkins and D.E.Brown, J.Am.Chem.Soc., 41 (1919) 499.
9. J.L.Lando, H.T.Oakley, J.Coll. Interface Sci.,25 (1967) 526.
10. R.T. Baxter, A.E.Wraith, Chem.Eng.Sci., 25 (1970) 1244.
11. A.K.Uhurana, R.umar, Chem.Eng.Sci., 24 (1969) 1711.
Eio

12. B.Bowonder, R.Kumar, Chem.Eng.Sci., 25 (1970) 25.


13. J.Hadamard, Compt.Rend, 152 (1911) 1753.
14. W.Rybczynski, Bull.Int.Acad.Polon.Sci.,Ser.& (1911) 40.
15. V.Levich, "Physicochemical Hydrodynamics", Prentice Hall,
Inc., New York, 1962.
16. B.V.Deryaguin, S.S.Dukhin, V.A.Lisichenko, Zh. Fiz. Khim., 33
(1959) 2280.
17. D.V. Deryaguin, S.S.Dukhin, V.A.Lisichenko, Zh. Fiz.
Khim., 34 (1960) 524.
18. D.W.Fuerstenau, C.H.Wayman, Trans. AIME, 211 (1958) 694.
19. R.E.Davis, CI.Acrivos, Chem. Eng. Sci., 21 (1966) 681.
20. J.F.Harper, Adv. Appl. Mech., 12 (1972) 59.
21. D.A.Saville, The Chem. Eng. J., 5 (1973) 251.
22. M.D.Levan, J.Newman, CIJChE, 22 (1976) 695.
23. N.N.Rulyov, Kholoidn. Zh., 42 (1980) 252.
24. M.D.Levan, J. Coll. Interface Sci., 83 (1981) 11.
25. Y.Suzin, S.Ross, J. Coll. Interface Sci., 103 (1985) 578.
26. S.S. Dukhin, N.N.Rulyov, D.S.Dimitrov, "Koagulacija i dinamika
tonkikh plenkov" , Naukova Dumka, Kiev 1986.
27. J.F.Harper, Q.J. Mech. appl. Math., 41 (1988) 201.
28. J.Boussinesq, Ann. Chim. Phys., 29 (1913) 349.
29. A.N.Frumkin, V.Levich, Zh. Fiz. Khim., 21 (1947) 1183.
30. J.Happel, H.Brenner, "Low Reynolds Number Hydrodynamics",
Prentice Hall, Inc.,,Englewood Cliffs, N. J. 19.
31. R. Cliff. J. R. Grace, M. E. Weber, "Bubbles, Drops, and
Particles" academic Press, New York, San Francisco,
London 1978.
32. A.D.Maude, Et-it. J. clppl. Phys., 12 (1961) 293.
33. E.Bart, Chem. Eng. Sci., 23 (19681 193.
34. M.fidamczyk, Z.Adamczyk, T.G.tl.van de Ven, J. Coil. Interface
Sci., 96 (1984) 204.
35. K.Ma%ys;a, Bull. Pol. &cad. Sci., Chemistry, 35 (1987) 441.
36. J.J.Bikermann, Trans.Faraday Sot., 34 (1938) 634.
37. S.Ch.Sun, Trans. AIME, 4 (1952) 65.
38. V.N.Prigorodov, Kholoidn.Zh., 24., 33 (1971) 459.
39. K.Ma&ysa, A.Pomianowski, Fizykochem. Problemy Przerdbki
Kopalin, 10 (1976) 119.
40. K.Ma&ysa, R.Cohen, D.Exerowa, A.Pomianowski, J.Coll.Interface
Sci., SO (1981) 1.
41. K.Ma*ysa, K.Lunkenheimer, R.Miller, C.Hartenstein, Colloids
Surfaces, 3 (1981) 329.
81

42. G.Nishioka, S.Ross, J.Coll.Interafce Sci., 81 (1981) 1


43. S.Ross, Y.Suzin, Langmuir, 1 (1985) 145.
44. Khr.Khristov, D.Exerowa, P.Krugliakov, Colloid Polymer sci.,
261 (1983) 265.
45. D.Exerowa, Khr.Khristov, J.Penev, in "Foams" R.J.Akers (Ed)
Academic Press 1976, p.109.
46. D.O.Shah, N.F.Djabbarah, D.T.Wasan, Colloid Polymer Sci., 256
(1978) 1002.
47. F.S.Shih, R.Lemlich, Ind. Eng. Chem. Fundam. 10 (1971) 254.
48. J.Cengel, R.Lemlich, J.Coll.Interface Sci., 128 (1989) 608.

49. O.Bartsch, Kolloidchem, Beidhefte, 20 (1924) 27.


50. A.Scheludko, Chemia Koloidc5w, WNT Warszawa 1969.
51. PI.Prins, M. van den Tempel, Proc. 4th Intern. Congress Surf ace
Activity, Brussells 1964, ~01.2, p.l.

52. J.Ross., G.D.Miles, Oil Soap, 18 (1941) 99.


53. A.Prins, in "Foams", R.J.Akers (Ed.) Academic Press 1976, p.51

54. G.Racz, E.Erdos, K.Koczo, Colloid Polymer Sci., 260 (1982) 720.
55. K.Koczo, B.Ludany, G.Racz, Periodyca Polytechnica Chem. Eng.,
31 (1987) 59.
56. P.M.Krugljakov, D.Exerowa, "Pena i pennyje plenki", Khimija,
Moscow 1990.
57. K.Ma%ysa, A.Pomianowski, In*. Aparatura Chem., 1 (1977) 24.
58. K.MaIysa, Fizykochem. Problemy Mineralurgii, 22 (1990) 101.
59. K.Ma&ysa, Proc. VII-th. Intern. Conf. "Liquid Properties in
Thin Layers", Kiev 1990 (in print).

60. S.Ch.Sun, Trans. PIME, 4 (1952) 65.


61. A.M.Gaudin, "Flotation", 2nd ed., New York-Toronto-London.
62. T.B.Thomas, J.T.Davies, J.Coll.Interface Sci., 48 (1974) 427.
63. K.Maaysa, Khr.Khristov, D.Exerowa, Colloid Polymer Sci.,
(acccepted).
64. B.V.Deryaguin, A.S.Titievskaya, Proc.11 Int.Congress Surface
Clctivity, 1 (1957) 11.
65. A.Scheludko, D.Exerowa, Kolloid Zh., 165 (1959) 148.
66. A.Scheludko. Fldv. Colloid Interface Sci., 1 (1967) 391.
67. D.Exerowa, A.Scheludko, Comm. Rend. Acad. Bul. Sci., 24 (1974)
47.
68. D.Exerowa, D.Ivanov, Compt. Rend. Bulg. Sci., 23 (1970) 547.
69. D.Exerowa, Khr.Khristov, I.Penev, in "Foams", R.J.FSkers (Ed)
&cademic Press 1976, p 109.

70. J.Lyklema, P.C.Scholten, K.J.Mysels, J.Phys.Chem., 69 (1965)


116.
82

71. J.Lyklema, K.Mysels, J.Amer. Chem. Sot., 87 (1965) 2539.


72. W.R.Entee, K.J.Mysels, J.Phys. Chem., 73 (1969) 3016.
73. B.V.Deryaguin, L.Landau, Zh. Eksper. Theor. Fiz., II (19411
802.
74. E.J.Verwey, J.T.S.Overbeck, "Theory of the Stability of
Lyophobic Colloids", Elsevier, Amsterdam 1948.
75. B.V.Deryaguin, M.M.Kusakov, Acta Physicochim. URSS, 10 (19391
25.
76. N.V.Churayev, B.V.Deryaguin, J.Coll. Interface Sci., 103 (19851
542.
77. D.U.Deryaguin, N.V.Churayev, Colloids Surfaces, 41 (3989) 223.
78. D.Exerowa, D.Kashchiev, Contemporary Physics, 27 (19861 429.
79. K.J.Mysels, M.C.Cox, J.D.Skewis, J.Phys. Chem. 65 (1961) 1107.
80. A.Scheludko, "Colloid Chemistry" Elsevier, Amsterdam 1966,
p.254 (from a 1963 Bulgarian text).
8i. M. van den Tempel, J.Lucassen, E.H.Lucassen-Reynders, J.Phys.
Chem. 69 (1965) 1798.
82. A.Prins, C.Arcuri, M. van den Tempel, J.Coll. Interface Sci.,
24 (19671 84.
83. V.V.Krotov, A.J.Rusanov, Kholoidn. Zh., 34 (19721 81 and 34
(1972) 297
84. V.V.Krotov, A.J.Rusanov, N.A.Ovrutskaja, Kholoidn. Zh., 34
(1972) 528.
85. V.V.Krotov, A.J.Rusanov. N.D.Rjasanova, Kholoidn. Zh., 34
(19721 534.
86. A.J.Rusanov, V.V.Krotov,Progress in Surface and Membrane
Science, Academic Press, New York, vol. 13, p 415.
87. K.MdfySd, Fizykochem. Problemy Mineralurgii, 22 (19901 101.
A. De Vries, in "Adsorptive Bubble Separation Techniques",
R.Lemlich (Ed) Academic Press, 1972, chap.2.
89. K.MaBysa, K.Lunkenheimer, R.Miller, C.Hempt, Colloids Surfaces,
16 (19851 9.
90. E. H .Lucassen-Reynolds, in "Anionic Surfactants”,
E.H.Lucassen-Reynolds tEd1, Marcel Dekker, New York 1979,
vol.ll.chap.5, p.173.
91. L.Ting, D.T.Wasan, K.Miyano, S-Q Xu, J.Coll. Interface Sci.,
102 (1984) 248.
92. E.H.Lucassen-Reynolds, J.Coll.Interface Sci,ll7 (19871 589.
93. Z.Adamczyk, J.Coll. Interface Sri., 133 (1989) 23.
94. J.Lucasseo, Trans.Faraday Sot., 4 (19681 2221; 64 (19681 22230
95. J.Lucassen, M.van den Tempel, J.Coll. Interface Sci., 41
(179721 491.
96. H.C.Maru, D.T.Wasan, Chem. Eng. Sci., 34 (1979) 1283.
83

97. S.Kretzschmar, K.Lunkenheimer, Bet-. Bunsenges. Phys. Chem., 74


(19701 1064.
98. K.Lunkenheimer, G.Kretzchmmar, Z.Phys. Chem. (Leipzig). 256.
(1975) 993.
99. K.D.Wantke, R.Miller, K.Lunkenheimer, Z.Phys. Chem. (Leipzig),
261 11980) 1177.
iOO.K.Ma$ysa, R.Miller, K.Lunkenheimer, Colloids Surface, 53 (15'911
47.
iOl.K.Mafysa, K.Lunkenheimer, R.Miller, C.Hempt, Colloids Surface,
16 (1985) 9.
lOZ.R.Miller, M.Ziller, Colloids Polym, Sci., 266 (1988) 532.

You might also like