Anda di halaman 1dari 40

Mode of action

of emulsiers
Manufacture of emulsions

Introduction to emulsions

1.1

Definitions

1.2

Emulsion types

1.3
1.3.1
1.3.2
1.3.3
1.3.4
1.3.5

Emulsifiers
Structure and mode of action of emulsifiers
Bancroft rule
HLB value
Kinetics
Emulsifier/coemulsifier principle

3
3
4
5
5
6

1.4
1.4.1
1.4.1.1
1.4.1.2
1.4.1.3
1.4.1.4
1.4.2
1.4.3
1.4.4

Properties of emulsions
Stability
Creaming and sedimentation
Aggregation and flocculation
Ostwald ripening
Coalescence
Droplet size distribution
Rheology
Electrical conductivity

8
8
8
9
10
11
12
13
13

Manufacture of emulsions

14

2.1

Microemulsions

14

2.2

Spontaneously emulsifying systems

14

2.3

Self-emulsifying systems

15

2.4

Mechanical emulsification

15

2.5

Emulsification using other phase boundaries

16

2.6

Phase inversion processes

17

Appendix

18

3.1

Marker method

18

3.2

Continuous emulsification in orifice systems

20

3.3

High-throughput screening: automated testing


and optimizing system

22

Stabilization of oil/water emulsions


with alcohol ethoxylates

23

Poly dimethyl siloxane emulsions in water


made with nonionic surfactants from BASF

30

Amino modified silicon microemulsions


made with nonionic surfactants from BASF,
e.g. for textile softening

32

Microemulsions

36

3.4

3.5

3.6

2
3.7

1 Introduction to emulsions
1.1 Definition

1.3 Emulsifiers

An emulsion is a dispersion of two


incompletely miscible liquids in one
another.

Emulsions* are thermodynamically


unstable. The droplets in the dispersed phase tend to coalesce into
larger droplets, thus reducing the
interfacial area between the two
phases and leading to a thermodynamically more favorable, i.e. lower,
energy state.

1.2 Emulsion types


Simple emulsions consist of a
hydrophilic (aqueous) and a lipophilic
(oily) phase. In the simplest case the
two phases are water and oil. The
internal or dispersed phase is dispersed in the external, continuous
phase in the form of fine droplets.
Depending on the nature of the
droplet-forming phase, a distinction
is made between oil-in-water (O/W)
and (W/O) emulsions.
In addition to simple emulsions,
there are also multiple emulsions. An
example of this type is the W/O/W
double emulsion, which has an
external phase of water and an internal phase consisting of a water-in-oil
emulsion.

oil-in-water
(O/W)

The interfacial energy of an


emulsion is given by:

U=A
U

interfacial energy
interfacial tension between
the two phases
interfacial area

This means that if the droplet size


decreases and the total volume of
the dispersed phase remains the
same the interfacial energy of the
emulsion will increase because the
total interfacial area increases.
But higher energy states generally
have lower thermodynamic stability,
so that the driving force for coalescence also increases (see also
section 1.4.1).

water-in-oil
(W/ O )

1.3.1 Structure and


mode of action of
emulsifiers
Emulsifiers are surface-active substances whose molecules consist of
a hydrophilic and a lipophilic part.
Because of their amphiphilic properties, free emulsifier molecules accumulate at the interface between
internal and external phases. A competing process also occurs in which
emulsifier molecules aggregate into
micelles. Above a certain concentration, known as the critical micellar
concentration (CMC), the proportion
of monomer emulsifier molecules
remains constant. In practice, the
best results are obtained when the
emulsifier is applied in concentrations well above the CMC.
The stabilizing properties of
emulsifiers are based on various
mechanisms, depending on the
type of emulsifier:
a) Electrostatic repulsion
b) Steric repulsion

water-in-oilin-water
(W / O / W )

Electrostatic repulsion
Different types of emulsions

Emulsifiers are used to reduce


the tendency to coalescence and
stabilize the droplets.
The interfacial tension of an O/W
interface is approx. 25 mN/m without emulsifier. Adding emulsifier
lowers the interfacial tension to
values typically around 3 5 mN/m.

* To distinguish them from the thermodynamically


stable microemulsions (cf. 2.1), emulsions are
sometimes also called macroemulsions.

By choosing suitable emulsifier systems, even lower interfacial tensions


of below 1 mN/m are possible.

3
Steric repulsion

In addition to repulsive forces,


there are also attractive forces
the London-van-der-Waals forces.

Potential energy
Steric repulsion
Electrostatic repulsion
London-van
der-Waals attraction
Sum of attractive and
repulsive forces

According to Derjaguin,
Landau, Verwey and Overbeek
(DLVO theory):

U total = U el + U st U vdW
U total
U el
U st
U vdW

Interdroplet distance

total potential energy


electrostatic repulsion
steric repulsion
London-van-der-Waals
forces

The resulting potential energy


is the sum of the electrostatic repulsion, steric repulsion and attraction
by London-van-der-Waals forces.
The figure below shows that below a
certain distance, after the repulsive
forces have been overcome, only
attractive forces operate. At this
point the droplets coalesce.

Dependence of attractive and repulsive forces on interdroplet distance

1.3.2 Bancroft rule


Whether an emulsifier is better able
to stabilize an O/W or a W/O emulsion depends on which is larger, the
hydrophilic or the lipophilic portion of
the molecule.
O
O
O

O
O
O

O
O
O

O
O
O

O
O
O

O
O

O
O

O
O
O

HO HO HO HO HO HO HO HO

O
O
O
O

O
O
O
O

O
O
O
O

O
O
O
O

O
O
O
O

O
O
O
O

O
O
O
O

O
O
O
O

HO HO HO HO HO HO HO HO

O
O
O
O
O
O
O
O
O
O
O

O
O
O
O
O
O
O
O
O
O
O
O

O
O
O
O
O
O
O
O
O
O
O
O

O
O
O
O
O
O
O
O
O
O
O
O

O
O
O
O
O
O
O
O
O
O
O
O

O
O
O
O
O
O
O
O
O
O
O
O

O
O

O
O

O
O

O
O

O
O

O
O

O
O

O
O

O
O

O
O

O
O

O
O

HO HO HO HO HO HO HO HO

Lipophilic portion larger

Hydrophilic and lipophilic


portions equal in size

Hydrophilic portion larger

Emulsifiers with a larger hydrophilic


portion are good O/W emulsifiers,
whereas those with a larger lipophilic
portion are better able to stabilize
W/O emulsions.
Too large a hydrophilic or lipophilic
part, on the other hand, leads to
both low interfacial affinity of the
emulsifier and poor packing at the
interface.
There is therefore an optimum size
ratio between the hydrophobic and
hydrophilic parts.

1.3.3 HLB value

For other emulsifiers, especially


ionic ones, a method was developed
by Davis in which the sum of experimentally determined increments is
calculated.

In many cases, the HLB value


(hydrophilic lipophilic balance) gives
an indication of the type of emulsion
for which emulsifiers are suitable.
Strictly speaking, the Griffin formula
applies only to nonionic ethoxylates,
which are classified on a scale of
0 to 20. Emulsifiers with low HLB
values tend to be lipophilic molecules, dissolving mainly in the oil
phase of an emulsion and better
able to stabilize W/O emulsions at
room temperature. Emulsifiers with
medium or large HLB values are
hydrophilic. They are more soluble
in water and are used preferentially
for stabilizing O/W emulsions at
room temperature.

In practice, the procedure is as


follows:
1) Choose an emulsion type (O/W
or W/O)
2) Find the HLB value of the oil from
tables 1
3) Select the HLB value of the
emulsifier or emulsifier mixture
to be the same as the HLB value
of the oil
4) Corrections may be necessary
if the temperature differs significantly from 25C or there are salts
in the aqueous phase. In the case
of ethoxylates, for example, the
HLB is chosen to be 0.5 1
higher for each increase of 10 K in
temperature or 5 wt% NaCl.

According to Griffins formula, the HLB value is calculated from the mass
of the lipophilic portion as a fraction of the total mass of the molecule:

HLB = 20 (1 M lipophilic/M total)


M lipophilic
M total

mass of lipophilic portion


total mass of molecule

lipophilic

hydrophilic

10

W/O emulsifiers

15
O/W emulsifiers

Mechanical energy

Coalescence

Deformation
and breakup

In addition to thermodynamics, the


kinetics of emulsions is very significant, especially in their manufacture.
The critical step is the fragmentation
of the internal phase. Large droplets
are divided into smaller ones by
introducing energy, e.g. by shearing.
The newly created surface must be
occupied by emulsifier molecules as
rapidly as possible to protect the
droplets and prevent them from
coalescing.
The rate at which an emulsifier
molecule occupies or vacates a
newly created interface depends
on a number of factors:
a) The rate at which emulsifier
molecules are transferred
from the continuous or the
dispersed phase to the vicinity
of the interface
b) Penetration of the interface
by emulsifier molecules
c) The orientation of emulsifier
molecules at the interface
d) Distribution of emulsifier
molecules over the interface
(Marangoni flow)
e) Removal of emulsifier molecules
from the interface by thermal
agitation
f) Removal of emulsifier molecules
from the interface by currents and
eddies

20

HLB scale according to Griffin

Continuous phase
+ emulsifier
+ phase to be dispersed

1.3.4 Kinetics

Under turbulent flow conditions,


transfer of emulsifier through
the continuous phase is faster than
droplet breakup, irrespective of the
emulsifiers diffusion coefficient.
Similarly, Marangoni flow at the
interface is usually more rapid than
the creation of new interfaces.
These two effects are therefore
seldom rate-determining.

Slow
emulsifier
Fast
emulsifier
Stable
droplets

5
1
e.g. Ullmanns Encyclopedia of Industrial
Chemistry, 6 th edition, Wiley-VCH, Weinheim, 2000.

Effect of adsorption rate on stabilization

This is not true of transfer through


the dispersed phase. In the case
of O/W emulsions, hydrophobic,
readily oil-soluble emulsifiers have
the advantage here. They become
concentrated in the oil droplets,
where they diffuse rapidly to the
phase interface.
Little is known about the kinetics of
transfer of emulsifier molecules to
the interface from micelles in the two
phases. Removal of small emulsifier
molecules from the interface is
dominated by thermal motion. In
the case of polymers with molecular
weights above approx. 100,000
daltons, eddies and currents play
an increasingly important role.
It has been found empirically that
small emulsifier molecules stabilize
newly generated interfaces more
rapidly than large ones.

1.3.5 Emulsifier/
coemulsifier principle
As described in the previous section,
small emulsifier molecules often
have an advantage when it comes
to rapid stabilization of the internal
phase during droplet fragmentation.
However, small emulsifier molecules
also have a disadvantage. Since
they are generally less tightly
adsorbed on to the interface than
larger emulsifier molecules, especially polymers, they are more readily
removed from the interface again by
Brownian motion. Furthermore, their
smaller molar mass often means
they have smaller repulsive groups
and therefore, according to the
DLVO theory, do not stabilize
droplets so well against coalescence.

Mechanical energy

Slow
emulsifier

Coalescence occurs
before interface is
occupied

Good long-term
stability

Fast
emulsifier

Coalescence can occur


because emulsifier
molecules leave
interface again

Poor long-term
stability

Dispersed phase

Time scale

Different long-term stabilities of rapid and slow emulsifiers

It has proved advantageous to


combine small and large emulsifier
molecules. The rapid adsorption of
the small molecules means they
immediately occupy a newly created
surface and then gradually make
room for the slower but more tightly
adsorbing large emulsifier molecules. In such combinations of two
or more emulsifiers, the emulsifier(s)
present in smaller quantities is (are)
known as the coemulsifier(s).

dN
Nd log M

Intensity

Emulsifier

Coemulsifier

log M
Molar masses (M) of emulsifier and coemulsifier

O
O
O
O
OH

O
O
O
O
OH

O
O
O
O
OH

O
O
O
O
OH

O
O
O
O
OH

Without coemulsifier

O OH O OH O OH O OH O
O
O
O
OH

O
O
O
OH

O
O
O
OH

O
O
O
OH

A further advantage of emulsifier/


coemulsifier mixtures is that higher
packing densities can be achieved at
the interface. Higher packing densities have the effect of increasing the
rigidity and thickness of the emulsifier film.

O
O
O
OH

Higher packing density


with surfactant alcohol as coemulsifier

It has been found in practice that


emulsifiers consisting of a mixture of
smaller and larger molecules as a
result of the synthesis process, for
example ethoxylates with a broad
EO distribution, do not need
coemulsifiers to be able to stabilize
emulsions well.

GPC of an ethoxylate with a broad molar mass distribution

1.4 Properties of emulsions

1.4.1.1 Creaming and


sedimentation

1.4.1 Stability
Emulsions have a much larger interfacial area between the two liquids
than the corresponding unemulsified
mixtures. Most emulsions are therefore thermodynamically unstable
even in the presence of emulsifiers.
Only in emulsions where the interfacial tension is extremely small
can the thermal energy exceed the
interfacial energy. Such emulsions
are known as (thermodynamically
stable) microemulsions (see 2.1).

The thermodynamic stability of the emulsion can be derived from the


Gibbs equation:

In addition to thermodynamic instability, there are other emulsion aging


mechanisms of significance to the
user, for example creaming and sedimentation. Creaming is reduced in
emulsions containing small droplets,
small density differences and a highviscosity continuous phase. It can
be quantified by optical techniques
or ultrasonic scattering. Very small
droplets (< 100 nm) are in favorable
cases also stabilized thermodynamically against creaming by Brownian
motion.

Stokes law gives the creaming


rate for a droplet dispersed in
a very dilute emulsion:

G = G emulsified G unemulsified = A T S
G

A
S
T

free energy
interfacial tension between the two phases
interfacial area
entropy
temperature

Normally, G > 0, i.e. the emulsion is thermodynamically unstable.


However, if the interfacial tension is very small, the greater disorder
(entropy) of the emulsified state leads to a thermodynamically stable
microemulsion (G < 0).

Density differences in emulsions lead to creaming (see figure) or sedimentation of droplets

v=
d
g

gd 2
18

droplet diameter
acceleration of gravity
difference in density
between dispersed and
continuous phases
viscosity of continuous
phase

Analogous but more complex


equations are used to describe
creaming for dispersed droplets
with finite surface viscosity,
for distributions of droplet diameters and for more concentrated
emulsions. But the parameters
derived from the simple Stokes
law apply here, too.

The degree of thermodynamic


stabilization is obtained from the
Boltzmann distribution law:

Height of
liquid
column h

d 3 g h 6kT
h
k
T
d

Emulsion with droplets


> 100 nm

height of liquid column


Boltzmann constant
temperature
droplet diameter
difference in density between
dispersed and continuous
phases
acceleration of gravity

Emulsion with droplets


<<100nm

Droplet density N/V

Emulsions with very fine droplets and small density differences can be entropy-stabilized against
creaming

1.4.1.2 Aggregation and


flocculation
Flocculation (or aggregation) refers
to the formation of clusters of two or
more emulsion droplets that behave
like distinct particles but in which the
identity of the individual droplets is
retained. Such clusters can even be
in dynamic equilibrium with single
droplets, individual droplets leaving
the cluster while new ones join it.
Flocculation is therefore a process
that is easily reversible, in contrast
to coalescence, though flocculation
often results in coalescence. Thermodynamically, flocculation is based
on a secondary energy minimum, as
described for example by the DLVO
theory (see 1.3.1 and 1.4.1.4).

The kinetics of irreversible flocculation in monodisperse, unstirred


emulsions can be described by the Smoluchowski equation:

( )

()

N
N
= W 4d D Tr
t V
V
W
d
D Tr

N/V
kT

()

kT N
2 V

probability of a successful collision


droplet diameter
diffusion constant of droplet
viscosity of continuous phase
droplet density
measure of thermal energy

It should be noted that to a first approximation flocculation kinetics


is independent of droplet size.

Attractive forces between droplets lead to aggregation or flocculation

1.4.1.3 Ostwald ripening


Another phenomenon of emulsion
aging is the growth of large droplets
at the expense of smaller ones:
Ostwald ripening. It arises from the
fact that small droplets dissolve
more readily in the continuous phase
than large ones. The phenomenon
was already described by Lord
Kelvin in 1871.

Kelvins equation describes the effect of droplet size on the


solubility of the dispersed phase:

RTln c(d) =

4 V Mol
+ RTln c(d = )
d

c(d) saturation concentration of the dispersed phase in the


continuous phase for a droplet of diameter d

interfacial tension
RT measure of thermal energy
V Mol molar volume of dispersed phase

Ostwald ripening causes large droplets to grow at the expense of small ones

In very dilute emulsions, the rate at which large droplets grow at the
expense of small ones is given by the Lifshitz-Slezov-Wagner (LSW)
equation.

64 D c (d = ) V Mol

(d)3 =
9 RT
t
D=diffusion coefficient of the dispersed in the continuous phase

10

The third power of the mean diameter d increases in proportion to


the interfacial tension and the solubility c of the dispersed in the
continuous phase. This growth is accelerated by the Brownian motion
of the droplets, the dispersed phase fraction and the presence of
micelles in the continuous phase.

The main way of reducing Ostwald


ripening is to ensure that the solubility of the dispersed in the continuous
phase is as low as possible. To
stabilize an emulsion in which the
dispersed phase is too soluble, an
inert, very poorly soluble auxiliary
can be added in the case of water,
for example, an extremely hydrophobic substance such as a hydrocarbon. Adding sufficient auxiliary
allows a stable equilibrium to be
established between the droplets.
The emulsion can then be technically
considered an emulsion of the auxiliary, the droplets of which contain
the original internal phase in a
dissolved state.

The Higuchi-Misra relation describes the effect of using an auxiliary:

4 V Mol
+ RTln a i = const
di
Adding an auxiliary that is very poorly soluble in the continuous phase
stabilizes an emulsion against Ostwald ripening if for every droplet i
the effects of its diameter d i and the activity a i of the more soluble
component in the auxiliary counterbalance one another until the above
equation is fulfilled.

V Mol
RT
di , ai

interfacial tension
molar volume of dispersed phase
measure of thermal energy
diameter of droplet i and associated activity of dispersed phase
in auxiliary

1.4.1.4 Koaleszenz
Coalescence is the merging of two
or more droplets into a single large
droplet, driven by the reduction in
interfacial area and hence interfacial
energy. Below a distance of approx.
100 nm, two oil droplets in an aqueous solution experience a perceptible attractive force that increases
as the droplets approach one
another. It is suspected that even at
these distances the water between
the droplets adopts structures of
lower entropy. The addition of a stabilizer such as an emulsifier super imposes on this attractive force a
repulsive force that derives from the
spatial requirement or charge of the
hydrophilic groups on the emulsifier.
The balance of these forces is
described by the theory of
Derjaguin, Landau, Verwey and
Overbeek (DLVO, see 1.3.1).
An energy barrier keeps the aggre-

Stabil

Adding a very poorly soluble auxiliary can stop Ostwald ripening

gated droplets at a distance and the


height of this barrier is critical for the
stability of the emulsion.

Apart from stabilizers, emulsions can


also be stabilized by restricting the
mobility of the droplets by a continuous phase of very high viscosity.
Such a high viscosity can, for example, be obtained by adding thickeners. Emulsions can also be stabilized
hydrodynamically. In this case,
strong turbulence reduces the contact time between two colliding
droplets to the point where drainage
of the liquid film of continuous phase
separating the droplets is impeded
and coalescence prevented.

Koaleszenz

11

The merging of separate droplets into a single larger droplet is called coalescence

1.4.2 Particle size


distribution
Typical emulsions have droplet
diameters ranging from 0.5 to
50 m. The size and size distribution
of the particles to a large extent
determine the properties of emulsions, for example their stability and
appearance. Thus emulsions tend to
appear white or milky because of
light scattering, unless the refractive
indices of the dispersed and continuous phases happen to be thesame.
Emulsions whose particles are much
smaller than the wavelength of
visible light appear opaque to clear.

9
8

Volume density
% distribution
q 3(x) (m -1)

As a rule, emulsion droplets are


present not in monodisperse form
but as a relatively broad size distribution. Only by applying special
techniques, such as membrane
processes (see below) or micromixers, can a virtually monodisperse
distribution be obtained. Various
techniques are available for measuring the size distribution, for example
laser diffraction methods.

Cumulative volume distribution Q3 (x) (%)

100
80

7
6

60

5
4

40

3
2

20

1
0

0
0.1

0.5 1

Volume density
% distribution
q 3(x) (m -1)

10

50 100

1000
Particle size (m)

Cumulative volume distribution Q3 (x) (%)

100
80

7
6

60

5
4

40

3
2

20

1
0

0
0.1

12

0.5 1

10

50 100

1000
Particle size (m)

Typical examples of broad and narrow particle size distributions (measured by laser diffraction)

1.4.3 Rheology

1.4.4 Electrical
conductivity

The rheological behavior of emulsions is determined mainly by the


fraction of the dispersed phase.
Dilute emulsions are characterized
by Newtonian flow.

The specific conductance of emulsions is determined by the conductivity of the continuous phase (or
phases in the case of bicontinuous
emulsions). It is therefore a simple
method for distinguishing W/O from
O/W emulsions. Moreover, the
transition from O/W to W/O, or to a
bicontinuous emulsion, is generally
accompanied by an abrupt change
in specific conductance, which is
therefore an indicator of the formation of new phases or of emulsion
inversion.

The viscosity of an emulsion depends on the viscosity of the


continuous phase. This relationship is described by Einsteins
equation:

emulsion = cont. phase (1 + 2.5)

viscosity
phase fraction of dispersed phase

The equation takes into account the contribution of momentum


transfer due to the droplets, which are regarded as rigid spheres.
The equation does not take into account further increases in
viscosity due to deformed, deformable or aggregated droplets, droplet
charges, polymers and thickener additions, or surfactant layers at
the interface. In concentrated emulsions the viscosity increases more
rapidly as a function of the phase fraction of the dispersed phase
than described by Einstein, and site exchange similar to that
encountered in diffusion phenomena in solids becomes the dominant
factor.

As the fraction of the internal


phase increases, the emulsion
starts to show non-Newtonian flow
behavior and the droplets are
packed more and more closely
together. At even higher internalphase concentrations, the droplets
are deformed into polyhedra and
rheologically the emulsion behaves
like a foam.
It exhibits a yield point and pronounced shear thinning caused by
deformation and relaxation
of the droplets. Such emulsions
are known as gel or high-internalphase emulsions.

Specific conductance (S/cm)

60
50

O/W

40
30
20
Bicontinuous

10

W/O
0

20

40

60

80
100
Phase fraction of oil (%)

Discontinuities in specific conductance indicate fundamental changes in emulsion type.


O/W emulsions generally have good conductivity, bicontinuous emulsions are much poorer
conductors, and W/O emulsions are practically nonconducting.

13

2 Manufacture of emulsions
2.1 Microemulsions
Microemulsions are a special case
because they are thermo -dynamically stable. By cleverly combining
the incompletely miscible phase
components and the emulsifier system or using very high surfactant
concentrations, extremely low interfacial tensions are obtained, so that
the emulsion is entropy-stabilized.

Microemulsions are thermodynamically stable. The interfacial tension


becomes so small approx. 1 to 100 nN/m that the increase in entropy
on emulsification exceeds the surface energy and the emulsion forms
spontaneously.

G = G emulsified G unemulsified = A T S < 0


Since the total energy G determined by the Gibbs equation is
< 0, the microemulsion is thermodynamically stable.

Water
Oil
0.1 m

In addition, the emulsifiers must not


be very soluble in either of the two
liquid phases at the application temperature, so that they remain almost
entirely at the interface and an emulsion forms spontaneously, creating
new interface in proportion to the
amount of emulsifier present.
In practice, the three-phase region
(Winsor III region) in the phase
diagram is sought, in which water,
oil and microemulsion coexist. The
amount of emulsifier, via its inter facial area requirement, determines
the volume fraction of the micro emulsion.

2.2 Spontaneously
emulsifying systems
Even systems that do not ultimately
form stable microemulsions can
emulsify spontaneously. The necessary interfacial energy is supplied by
the entropy of mixing. For example,
a mixture of oil, surfactant and
ethanol emulsifies spontaneously in
water because the ethanol diffuses
into the water. Another mechanism
involves intermediate microemulsions that form as a result of high
surfactant concentrations at the
interface and break down in the
continuous phase. Turbulence at the
interface can in both cases lead to
the formation of a (metastable) emulsion instead of two macroscopically
separate water and oil phases.

Electron micrograph of a frozen microemulsion


Source: Science, Vol. 240, Microemulsions, Manfred Kahlweit. Copyright 1988, AAAS

Emulsifiable concentrate
Mole fraction
of water
increases

Temperature

2 phases

3 phases
(microemulsion,
oil, (water)

1 phase
(Microemulsion)

Entropy of
mixing leads
to formation
of interface

2 phases

14
(wt. % surfactant)
Typical phase diagram of a microemulsion

When emulsifiable concentrates are added to water, chaotic


turbulence at the interface leads to the formation of an emulsion
instead of two macroscopic phases.

2.3 Self-emulsifying
systems
In the case of microemulsions and
spontaneously emulsifying systems,
emulsions are formed without energy
being supplied from the outside,
but with self-emulsifying systems
this is not strictly true. However, only
a small power input is required to
turn self-emulsifying systems into
emulsions, for example from a slowly
turning stirrer or simply by shaking.
Important for this process are small
dynamic interfacial tensions, so that
even a small power input will result
in a sufficiently high critical Weber
number to produce small droplets.
Generally speaking, oil-soluble
surfactants with a low HLB value,
or mixtures containing such
surfactants, tend to be used for
spontaneously or self-emulsifying
systems.

2.4 Mechanical
emulsification
In mechanical emulsification
processes involving high power
densities, very low interfacial
tension is helpful but not essential.
High-turbulence zones, laminar
shear flows and cavitation induce
emulsion formation even without an
emulsifier. Much more important in
such processes is good stabilization
of the resulting emulsion, since the
outlet of dispersion machines is
often characterized by turbulent
flow conditions and a high droplet
collision rate.

The Weber number is defined as the quotient of the external and


internal forces acting on a droplet. It describes the emulsion process
in terms of laminar forces.

We =
vl
c
d

v /l c d

laminar shearing field


viscosity of continuous phase
droplet diameter
interfacial tension between phases

The external forces are transmitted by the viscosity of the continuous


phase, while the internal restoring forces are caused by the interfacial
tension. The critical Weber number We cr is obtained by substituting
into the above equation the diameter of the smallest droplet that can
just still be (or can no longer be) broken up.
Critical Weber numbers can also be calculated from the type of laminar
flow (extensional or shear flow) and from the viscosities of the continuous
and dispersed phases. The diameter of the smallest attainable droplet is
then derived from the above equation.

The Reynolds number relates inertial and viscous forces to


one another:

Re =
v
c
l
c

v l c
c

flow rate
viscosity of continuous phase
characteristic dimension (see below)
density of continuous phase

If the dimensions of the stirrer or other emulsifying apparatus are sub stituted for l, then there is laminar flow up to Re 10 3 and turbulent flow
above Re 10 4. Between these values is a tran -sition range. Substituting
the droplet diameter for l allows the Reynolds number in the vicinity of a
droplet to be calculated. Below Re Tr 1 droplet breakup (Re Tr ) is caused
mainly by laminar forces; above Re Tr 1 mainly by inertial forces.

15

A large number of emulsifying


machine types are available on
the market. Typical are rotorstator
systems such as toothed disc dispersing machines, reaction mixing
pumps and colloid mills. Other
machines consist of pumps combined with static mixers, nozzles or
orifices, for example the very efficient high-pressure homogenizers.
Others again use ultrasonic
sonotrodes.

Significant criteria for selecting an


emulsifier for such a mechanical
emulsification process are that it
rapidly occupies the newly formed
interface, is not carried away again
by eddies, and protects droplets
from coalescence in the turbulent
discharge zone of the emulsifying
apparatus. Low-molecular fast
emulsifiers are generally superior in
this regard to polymeric emulsifiers
(see 1.3.4).

FineFine
emulsion
emulsion

Crude
emulsion
Crude
emulsion

High-pressure
High-pressure
homogenizer
homogenizer

Rotor-Stator-System
Rotor-Stator-System

High-pressure homogenizers and rotor-stator systems are commonly used to manufacture emulsions

2.5 Emulsification using


other phase boundaries

Droplet formation at membrane pores


Mechanical energy

Continuous phase
+ emulsifier

Membrane
Phase to be dispersed

Fast
emulsifier

Very slow
emulsifier
Mechanical energy

Membranes are very energy-efficient emulsifying apparatuses. However, they are mostly too
expensive for wide-scale industrial application.

16

Another method of producing emulsions is to use other phase boundaries, for example the solid-liquid
interfaces of membranes and micro mixers. Here the large interface of
what is to be the dispersed phase is
generated by appropriate differential
pressure across a membrane that
is not wettable by emulsifier.
The process is very energy-efficient.
Nature makes use of it, for example,
in producing milk.
A gas-liquid interface, such as in
aerosol and condensation processes,
can also be used to produce emulsions.

2.6 Phase inversion


processes
A simple way to obtain very fine
emulsions with very little energy
expenditure is the phase inversion
method. It is particularly simple to
apply with ethylene oxide based
surfactants. Three basic cases are
distinguished:
1) In the first, the surfactant is mixed
with, for example, oil and water
and heated to above, or just
below, the phase inversion temperature (PIT). The mixture is
then emulsified and the emulsion
is stabilized by cooling. This
approach takes advantage of the
low interfacial tension at the PIT,
and only a small amount of energy
is required to produce a fine emulsion. Emulsification above the PIT
and subsequent cooling is the
more efficient method, but care
must be taken to ensure that the
emulsion does not break during
inversion. The risk of this occurring is smaller if the mixture is
heated to just below the PIT*,
which assumes, however, that
the PIT is known exactly.

Temperature

W/O
Case 1

Case 3

PIT
O/W
Case 2
Oil fraction of the emulsion
The phase inversion method produces fine emulsions with little energy expenditure.
It is illustrated here for ethylene oxide based emulsifiers.
Case 1 involves the inversion of a W/O emulsion into an O/W emulsion by cooling.
Case 2 is isothermic phase inversion, and Case 3 is a combination of the other two cases.

2) In the second case, the surfactant


is mixed with, for example, oil and
very little water and emulsified.
Then, without altering the temperature, water is added to induce
phase inversion. (Such inversion
of an emulsion containing a very
large fraction of the future internal
phase shows hysteretic behavior
when the emulsion is rediluted
with oil and can be described by
catastrophe theory; thus it is
called catastrophic phase inversion. Inversion of small regions
of the emulsion ultimately leads
to inversion of the emulsion as a
whole.) Fine emulsions can be
produced very simply by this
method, but it requires an emulsifier that tends to be soluble in
oil rather than water. Such emulsifiers usually have small hydrophilic
groups and are therefore, according to the DLVO theory, poorer
stabilizers.

3) The third case is a combination


of the first two. A water-in-oil
emulsion or microemulsion, for
example, is produced at high
temperature and high surfactant
concentration and then inverted
by adding cold water. This is a
simple way to produce very fine,
stable emulsions.
These phase-inversion techniques
can also be used with ionic surfactants, but it should be remembered
that, because of the expan sion of
their Debye ion clouds, ionic surfactants become more hydrophobic
at low temperatures and more
hydrophilic at high temperatures
the exact opposite of ethylene oxide
based surfactants.

17
* It is usual to emulsify approx
10 15K below the PIT.

3 Appendix
3.1 Marker method
A technique used at BASF to determine the stabilizing properties of
emulsifiers is the marker method*,
based on the Danner dyeing
method**. In this technique, the dispersed phase is marked with marker
substances in such a way that after
emulsion coalesced and non-coalesced droplets can be distinguished
and determined quantitatively.
Here we describe how the method
is used to analyze O/W emulsions
with oil-soluble dyes as marker
substances.
First, two coarsely dispersed raw
emulsions are prepared with identical
formulations but a different-colored
oil phase in each (one blue and one
yellow). The two emulsions are mixed
by slowly stirring them together.

green) are formed by coalescence,


i.e. the droplets combine because
they are not sufficiently stabilized
by the emulsifier. The higher the proportion of droplets of mixed color,
the more poorly the emulsifier has
stabilized them against coalescence.
If only a few green droplets or none
at all are found in the sample, it
can be concluded that the emulsion
is well stabilized by the emulsifier.
The emulsion is evaluated quantitatively by digital image analysis,
which allows the area fraction of
each droplet color to be determined.
Quantitative determination of the
stabilizing properties under actual
emulsification conditions is therefore
possible.

The emulsion mixture is then subjected to a coalescence experiment.


This may, for example, be storage,
temperature change, shearing,
centrifuging or addition of chemicals.
The emulsion is subsequently
examined under a microscope and
the proportion of droplets of mixed
color determined. Droplets containing the mixed color (in this case

* DE 10247086
** Dr. Thomas Danner, Tropfenkoaleszenz in
Emulsionen, dissertation at the University
of Karlsruhe, GCA-Verlag, Herdecke 2001

Fine emulsification

Crude emulsion A

1:1 mixture
Micrograph showing
coalesced droplts
Crude emulsion B
Principle of the marker method

18

Advantages of the marker


method:
Coalescence can be isolated from
other effects such as Ostwald
ripening and flocculation.
Coalescence can be accurately
measured at an early stage,
reducing development time.
Coalescence can be measured
under shearing/stirring conditions.

Original image

 Droplet separation

Source: GCT, Dr. Thomas Danner

 Color evaluation

Evaluation of the micrographs


Left image sequence: no coalescence, two signals, one for each of the yellow and blue color areas
Right image sequence: 100% coalescence, only one signal is obtained for the green color area

19

3.2 Continuous
emulsification in orifice
systems

An emulsification technique whose


results can be readily evaluated by
the marker method (see separate
appendix) is emulsification in orifice
systems, as used for example in
high-pressure homogenizers. In a
laboratory test, an autoclave is filled
with two differently colored crude
emulsions with a particle size of
20 30 m. While being slowly

stirred, the crude emulsion with


different colored droplets is forced
through an orifice by gas pressure.
Before entering the orifice, the crude
emulsion passes through a region
of laminar extensional flow, which
deforms the emulsion droplets.
Within the orifice is a small zone
where laminar shear flow predominates. Here the droplets are further
deformed and some of them break
up. Most of the droplets, however,
break up in the turbulent outflow of
the orifice, where, depending on the
power input, cavitation may occur

Orifice

and produce shock waves. During


the emulsification process, the
droplets collide with one another.
Where droplets are created by division of larger droplets, the emulsifier
must rapidly occupy the newly
formed interface or they will coalesce. Details of these processes are
discussed in numerous monographs
on emulsification.

Turbulence

Laminar
extensional flow

High pressure,
low flow rate
Autoclave with orifice in outlet

Laminar
Extensional
flow

Emulsification zones in flow through an orifice

Shearing,
turbulence
Cavitation

Extension of droplets and occupation of newly formed surface by emulsifier

20

Low pressure,
high flow rate

To obtain a measure of the perfor mance of the emulsifier in terms


of coalescence, the coalescence
probability of the droplets is determined from the measured proportion
of green droplets. The lower the
coalescence probability, the better
the emulsifier stabilizes the emulsion
against coalescence in turbulent
regions and other shearing zones.

We recommend the emulsifiers listed


below for preparing oil-in-water
emulsions in laminar flow under
high shear with a high power input
(> 10 5 10 6 W/m 3). These conditions
are typically present in screen mills,
rotor-stator colloid mills and highpressure homogenisers. The
droplets that are formed in this type
of apparatus are usually small
enough, but the emulsifier molecules
have to be able to migrate quickly to
the newly formed interfaces and they
have to be capable of preventing the
oil droplets from coalescing in the
zones of turbulent flow in the outlet
ports of the machinery.
Of course, the size of the droplets
also depends on the type of emulsifi-

cation process and the on the power


that is introduced into the system.
Depending on the type of process
that is involved, much smaller or
larger droplets can be obtained if
required, but the ranking of the
emulsifiers in the diagrams still
applies.
Because of the similarity of these
processes, these recommendations
also apply without restriction to
membrane processes, and they
also apply to a certain extent to
ultrasonic processes and turbulent
emulsification processes with a high
power input, such as in high-speed
stirrers, toothed-ring dissolvers
or in large rotor-stator-type homo genisers.

Good emulsifier
Stable emulsion
No coalescence

Poor emulsifier
Unstable emulsion
Coalescence

Difference between emulsifiers with good and poor stabilizing properties

Emulsifiers recommended for process with a high power input


Small droplets with a low tendency to coalesce can be prepared with the
following emulsifiers:
Paraffin oil

Emulan P, AT 9, NP 3070

Naphthenic mineral oil

Emulan TO range, NP 3070, EL

Aromatic mineral oil

Emulan TO range, OU, OC, OG


Emulan NP 3070
Emulphor FAS 30

Triglycerides

Emulan EL

Silicone oil

Emulan TO range, EL
Emulan NP 3070

21

3.3 High-throughput screening: automated testing


and optimizing system
An emulsifier selection process that
is rational and above all reproducible
is the high-throughput screening
method (HTS).
This fully automated process, consisting essentially of a robot with
metering, emulsifying and analysis
units, makes it possible to vary a
large number of different parameters
(concentration, temperature, emulsifier type, composition of formulation,
etc.) in very extensive screening
tests.

The emulsification step can, for


example, be carried out using an
ultrasonic probe or a rotor-stator
system (both laminar and turbulent
flow). The stabilizing effect of
emulsifiers can be evaluated by
transmittance measurements or
by automatic measurement of the
particle size distribution using laser
diffraction. Other significant variables,
such as the viscosity and creaming
properties of the emulsion obtained,
can also be determined fully automatically.

Pipetting station for emulsifiers

HTS apparatus with robot

22

3.4 Stabilization of oil/water emulsions


with alcohol ethoxylates
Introduction
The HLB concept gives some
indication of which hydrophile of
the emulsifier is suitable for an existing emulsification task. On the other
hand, which emulsifier class is best
suited in the case of optimization
of the HLB value cannot be given
by the concept, since it oversimplifies
the relevant physical-chemical effects.
Ultimately, this is only possible
empirically.

Simple, industrial-grade alcohol


ethoxylates are especially suited for
a systematic study of the stability of
oil-water emulsions. They are simple
to produce from a great variety of H
active compounds. Their HLB value
is easily variable for any hydrophobic
part and they react only slightly to
impurities in both the oil and aqueous
phases, such as ions, for example.

Why can one surfactant class stabilize while another cannot?

OH

Ethoxylates of linear alcohols

C 10 Guerbet alcohol ethoxylate


(Lutensol XP)

n-Decanol ethoxylate

Nonylphenol ethoxylate (Lutensol


AP)
Macroemulsions were prepared with
a low power input.
Quantity ratios:
30% oil, 1% surfactant, 69% water
Emulsification procedure:
Propeller agitator, P/V=10 4 W/m 3,
23C, 15 minutes
Then the emulsions were stored at
23C, the stability index S determined by the marker method and
plotted for each ethoxylate class
as a function of the HLB value.

OH

Other alcohol ethoxylates

C 10 Guerbet alcohol ethoxylate


containing small quantities of
higher alkylene oxides (Lutensol XL)

C 12C 14 coconut fatty alcohol


ethoxylate (Lutensol A ..N)
C 13C 15 Oxo alcohol ethoxylate
(Lutensol AO)
C 16C 18 Tallow fatty alcohol ethoxylate (Lutensol AT)

Ethoxylates of oxo alcohols from


the oligomerization of higher
olefins
iso C 10 Oxo alcohol ethoxylate
(Lutensol ON)

Emulsion preparation
and methodology
The stability against coalescence at
room temperature of O/W emulsions
was measured with two model oils:
Thin fluid paraffin oil with a visco sity of 30 mPas (23C) and an
Mn of about 300 as a model of
hydrophobic hydrocarbons
Sunflower seed oil with a viscosity
of 56 mPas (23C) and an acid
number of maximum 0.15 mg
KOH/g as a model of native
triglycerides

iso C 13 Oxo alcohol ethoxylate,


medium branched (Lutensol TO)

Demineralized water was used as


the aqueous phase

iso C 13 Oxo alcohol ethoxylate,


highly branched (Lutensol TDA)

Eleven classes of alcohol ethoxylates


served as emulsifiers. For each class,
a set of ethoxylates was prepared
in steps of 1 2 HLB units and
measured.

iso C 17 Oxo alcohol ethoxylate,


highly branched

23

Visualized process for calculation of the stability index S

From repetitions it follows that s(S) 0.25.


A 2s bar is found in each diagram.

From time to time the average green content is


determined with 30 different microscopic images

S = - log (r x month) is repre-sented as a function


of the HLB
200 % / (100 % green %)
2.6
2.5
2.4
2.3
2.2
2.1
2
1.9
0
1
2
3

S = - log (r * month)
2
1
0
-1
r = 0.079 1/month

-2
-3

10 11 12

t in month

13

14

15 16

17 18

HLB value (= 20* wt.% EO)

A hyperbola was adapted to the measured data


by using the least squares method

Values of S = 1.5 correspond to roughly 10% coalesced droplets within a


half-year, values of S = -1.5 to about 10% coalesced droplets in five hours.

-1

C 12C 14 to C 16C 18 alcohol ethoxylates: HLB= approx. 9 13

-2

less suitable

Lutensol A..N, AO

In each case by about one order of


magnitude.

Lutensol AT

n-Decanol ethoxylate

-3
6

10

11

well suited

C 10  C 12C 14 = C 13C 15  C 16C 18

Weeks Months

C 20C 22 x 7.3 EO

C 10 Alcohol ethoxylates:
HLB = approx. 13

On the other hand, the resistance to


coalescence increase in the order

24

A comparison with C 20C 22 x 7.3 EO


(HLB = 10) shows that an additional
increase in resistance to coalescence by chains longer than C 18 is

Days

In the case of emulsions of paraffin


oil in water stabilized by ethoxylates
of linear alcohols, one recognizes,
on the one hand, that the suitable
HLB range depends on the chain
length of the alcohol:

10 % of the drops
have coalesced after

Stability index
2
2s

Minutes Hours

Emulsions of paraffin oil


in water ethoxylates of
linear alcohols

12

13

14

15

16

17 18
HLB value

not possible; therefore, among the


ethoxylates of linear alcohols C 16C 18
ethoxylates such as Emulan AT 9
or Lutensol AT 11 are optimal.

10 % of the drops
have coalesced after

Stability index
2
2s

Among the ethoxylates of branched


oxo alcohols one finds HLB ranges
similar to those among the linear
alcohol ethoxylates

Weeks Months

Emulsions of paraffin oil in


water ethoxylates of
branched alcohols

iso C 17 Ethoxylate

iso C 13-Alcohol ethoxylates,


medium branched, Lutensol TO
HLB = approx. 10 15

-1

Days

Lutensol TO

iso C 10-Alcohol ethoxylates,


Lutensol ON: HLB = approx. 14

Minutes Hours

Lutensol TDA
Lutensol ON
-2

iso C 13-Alcohol ethoxylates, highly


branched, Lutensol TDA
HLB = approx. 12 14

-3
6

10

11

iso C 17 Alcohol ethoxylates, highly


branched HLB = approx. 10 15
branched Lutensol TO types offer a
broader HLB window than the highly
branched Lutensol TDA brands,
which is an advantage, since with
Lutensol TO one has to make fewer
compromises with regard to other
target values such as foam or formulability.

One can also recognize a somewhat


better stabilization than with the
ethoxylates of linear alcohols. Therefore, Lutensol TO, for example,
stabilizes just as well as ethoxylates
of branched or linear alcohols with
16 to 18 C atoms in the hydrophobic
part. Surprisingly, the medium-

12

13

14

15

16

17 18
HLB value

As opposed to the linear alcohol


ethoxylates, among which C 16C 18
ethoxylates clearly stabilize better
than C 12C 14 ethoxylates, in the case
of the branched alcohols, the longer
chains, such as in the analogous
iso C 17 alcohol ethoxylates, do not
provide any significant extra advantage.

Emulsions of paraffin oil in water


direct comparison of C 10 ethoxylates
In the direct comparison of the C 10 alcohol ethoxylates linear, Guerbet and
iso C 10 oxo alcohol one finds ideal HLB values around 13 14, but above
all, the following surprising order in the coalescence stabili zation:

less suitable

well suited

Lutensol ON  Lutensol XP = linear decanol ethoxylate  Lutensol XL

10 % of the drops
have coalesced after

1
Lutensol XL
0
Decanol ethoxylate

Lutensol XP

Days

Weeks Months

Stability index
2
2s

Minutes Hours

-1
Lutensol ON
-2
-3
6

10

11

12

13

14

15

16

17 18
HLB value

25

Emulsions of paraffin oil in water


direct comparison of C 12-C 15 ethoxylates
In the direct comparison of the C 12 to C 15 alcohol ethoxylates one recognizes
once more the above mentioned order with slight advantages on the part of
Lutensol TO due to the larger HLB window:

less suitable

well suited

Lutensol A..N = Lutensol AO  Lutensol TDA = Lutensol TO

Weeks Months

10 % of the drops
have coalesced after

Stability index
2
2s

Lutensol TO

Days

Lutensol A..N, AO

Minutes Hours

-1
Lutensol TDA
-2
-3
7

10

11

Emulsions of paraffin oil


in water substitution of
nonylphenol ethoxylate
In the direct comparison of fatty
alcohol ethoxylate vs. nonylphenol
ethoxylate, C 16 to C 18 ethoxylates
can be excluded because of their
totally different behavior in terms of
surfactant properties and formulation
behavior, and Lutensol TDA because
of its poor biodegradability. Therefore one can consider
Lutensol TO and
Lutensol XL
to be optimal substitutes for nonylphenol ethoxylate for stabilizing
paraffin oil emulsions.

12

13

14

15

16

17 18
HLB value

Summary: Paraffin oil in water


In conclusion, we can say that
among the ethoxylates tested, the
types
Lutensol TO,
Emulan/Lutensol AT and
Lutensol XL
are the best suited for stabilizing
emulsions of paraffin oil in water.

10 % of the drops
have coalesced after

Stability index
2
2s

Weeks Months

1
Lutensol TO

Days

0
Lutensol XL
Nonylphenol ethoxylate
-2
-3
6

26

10

11

12

13

14

15

16

17 18
HLB value

Minutes Hours

-1

C 16C 18-Alcohol ethoxylates,


Emulan/Lutensol AT:
HLB = approx. 10 12

Weeks Months

1
Lutensol AT
0
C 20C 22 x 7.3 EO

Days

In the case of emulsions of sunflower


oil in water stabilized by ethoxylates
of linear alcohols, one finds a totally
different picture than in the case of
paraffin oil. Ethoxylates of linear
decanols cannot stabilize the emulsion at all, only from C12 on does one
recognize a slight stabilization around
HLB = 11. But good stabilization is
only achieved by the ethoxylates of
the long-chained C16C18 fatty alcohols.

10 % of the drops
have coalesced after

Stability index
2
2s

-1
Lutensol A..N, AO
-2

n-Decanol ethoxylate

-3
6

10

11

12

A comparison with C 20C 22 x 7.3 EO


(HLB = 10) shows that the stabilization of triglyceride in water with
ethoxylates of linear alcohols clearly

13

14

15

16

Minutes Hours

Emulsions of sunflower oil in


water ethoxylates of linear
alcohols

17 18
HLB value

decreases again beyond C 18. This is


due to the incompatibility between
the waxlike C 20C 22 alcohol and the
triglyceride.

Emulsions of sunflower oil in water


ethoxylates of branched oxo alcohols
less suitable

well suited

Lutensol ON  Lutensol TDA  iso C 17 ethoxylate  Lutensol TO

Months

10 % of the drops
have coalesced after
Lutensol TO

iso C 17 Ethoxylate

Days

Weeks

Stability index
2
2s

Lutensol TDA

Hours

-1
Lutensol ON
-2

Minutes

Now one finds a very surprising


picture with sunflower oil and
ethoxylates of branched alcohols.
The order C 10  C 13  C 17, anticipated from the ethoxylates of linear
alcohols is found only at a significantly higher level, although the
medium branched iso C 13 ethoxylates Lutensol TO display highly
reproducibly a very sharp maximum
in the stabilization around HLB = 11,
which corresponds approxmately to
Lutensol TO 6. Here the ethoxylates
exceed highly branched iso C 13
ethoxylates by more than an order
of magnitude in stabilization, so that
the following order now exists:

-3
6

10

11

12

13

14

15

16

17 18
HLB value

27

Emulsions of sunflower oil in water


direct comparison of C10 ethoxylates
In the direct comparison of the C 10 ethoxylates, one recognizes an
unambiguous order in the stabilizing effect

less suitable

well suited

Linear C 10 ethoxylate  Lutensol ON  Lutensol XL = Lutensol XP


where only the surfactants based on C 10 Guerbet alcohol act well against
coalescence.

10 % of the drops
have coalesced after
Weeks Months

Stability index
2
2s
1

Lutensol XP

Lutensol XL

Days

Minutes Hours

-1
Lutensol ON
-2

n-Decanol ethoxylate

-3
6

10

11

12

13

14

15

16

17 18
HLB value

Emulsions of sunflower oil in water


direct comparison of C 12-C 15 ethoxylates
In the direct comparison of the C 12-C 15 alcohol ethoxylates the great
differences in performance are again especially striking.

less suitable

well suited

Lutensol A..N, Lutensol AO  Lutensol TDA  Lutensol TO

Weeks Months

10 % of the drops
have coalesced after

Stability index
2
2s

Lutensol TO

Days

0
Lutensol TDA
Lutensol A..N, AO
-2
-3
6

28

10

11

12

13

14

15

16

17 18
HLB value

Minutes Hours

-1

Emulsions of sunflower oil in water


nonylphenol substitution
Compared with nonylphenol ethoxylate, one will note that the latter like
Lutensol TO, but with a slightly higher HLB displays a surprisingly good
stabilization and is exceeded only by Lutensol TO or long-chained ethoxylates such as Emulan AT or Lutensol AT.

Weeks Months

10 % of the drops
have coalesced after

Stability index
2
2s

Lutensol TO

Days

0
Lutensol AT

Minutes Hours

-1
Nonylphenol ethoxylate
-2
-3
6

10

11

Effect of the degree of


branching on the emulsion
stabilization
If the maximally attainable stability
index with optimal HLB value for a
group of commercial and experimental ethoxylates is plotted against the
number of branches per C atom in
the hydrophobic part, then one finds
a surprising correlation for the group
of C 10 and C 12 to C 15 alcohol ethoxylates*.
The optimal window contains
ethoxylates of alcohols with about
0.15 branches per C atom in the
hydrophobic part, such as the commercial ethoxylates Lutensol TO
or Lutensol XP. The reason for this
window is not clear; obviously an
optimum is achieved here in terms
of packing and flexibility in the interfacial film.

12

13

14

15

16

17 18
HLB value

It is further favorable that the ethoxylates of these medium-branched


alcohols are usually readily biode gradable and lowly aquatoxic. These
products are therefore the best
choice if one is searching for short
or medium chained non-ionic surfactants that are to stabilize emulsions
well.

2
Lutensol TO
C 12 C 15
Ethoxylates
of paraffin oil

C 12 C 15
Ethoxylates
of sunflower oil

C 10 Ethoxylates
of paraffin oil

-1
Lutensol XP

C 10 Ethoxylates
of sunflower oil

-2

-3
0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Number of branches per C atom in the hydrophobic part


* Ethoxylates of long-chained alcohols with
16 18 C atoms in the hydrophobic part
stabilize against coalescence so well that
no correlation with the degree of branching
is recognizable.

29

3.5 Poly dimethyl siloxane emulsions in water


made with nonionic surfactants from BASF
Three principal methods
are used to produce poly
dimethyl siloxane (PDMS)
emulsions:
Microemulsions can only be
made if the silicon oil is of very low
viscosity e.g. short chains or cyclic
oligomers or by using siliconbased emulsifiers. Here, simple mixing both the oil and water phases
together with a well-tuned emulsifier
package leads to the microemulsions; a special emulsification
machine is not required.
High-energy emulsification is
used to emulsify normal (viscous)
PDMS oils. For oils with viscosities
up to ca. 500 mPas rotor-stator
machines can be used, whilst highpressure homogenizers can emulsify
oils with viscosities up to ca.
100,000 mPas. Emulsification of
viscous oils at elevated temperatures as done with e.g. hydrocarbons is usually not helpful,
because of the low temperature
dependence of the PDMS viscosities.
Emulsion polymerization is
another common process to manufacture PDMS emulsions. For target
oil viscosities beyond 100,000
mPas emulsion polymerization is
the only feasible process.

The following charts illustrate the


stability of aqueous PDMS emulsions against coalescence for various alcohol ethoxilates as a function
of emulsifier chemistry and HLB
value as determined by the Marker
method.

Lutensol AT, TO and XL offer best


coalescence stabilization. Lutensol XL
is preferred, as it forms small droplets
at rather low power input and stabilizes across a wide HLB range.

Upon stirring add 10 parts of


emulsifier
Then add 50 parts of the PDMS
oil, mix thoroughly
Subject the mixture to the
emulsification procedure
Fine-tune the emulsifier content
and HLB value for optimum
stability.

1% of the droplets coalesced after

Charge your vessel with 40 parts


of water

1% of the droplets coalesced after

Performance of alcohol ethoxilates


Standard process to emulsify Performance
of alcohol ethoxilates
Coalescence
stability at room temperature
PDMS
Coalescence stability at room temperature
month

month

weeks

weeks

days

days

hours

hours

minutes

minutes
6

10

11

12

13

10

14

15

16

value 13
11HLB12

15 16
HLB value
30% oil, 1% emulsifier, 69% water, storage at 23C

30% oil, 1% emulsifier, 69% water, storage at 23C

14

Lutensol /Emulan AT
Lutensol XL

30

Lutensol XP

Lutensol /Em

Lutensol ON

Lutensol XL

Lutensol TO

Lutensol XP

APEO (for comparison)

Lutensol O

Lutensol TO

APEO (for co

Performance of alcohol ethoxilates


1% of the droplets coalesced after

1% of the droplets coalesced after

Performance
of alcohol ethoxilates
Coalescence
stability at elevated temperature
Coalescence stability at room temperature
month

month

weeks

weeks

days

days

hours

hours

minutes

minutes

10

11

12

13

14

10

11 HLB12value13

15

16

15 16
HLB value
30% oil, 1% emulsifier, 69% water, storage at 70C

30% oil, 1% emulsifier, 69% water, storage at 23C

14

Lutensol /Emulan AT

At elevated temperatures the longer


chain ethoxilates Lutensol AT and TO
offer significant better stability
than the medium chain nonionics
Lutensol XL and XP.

Lutensol XL
Lutensol XP

Lutensol /Emulan AT

Lutensol ON

Lutensol XL

Lutensol TO

Lutensol XP

APEO (for comparison)

Lutensol ON
Lutensol TO

APEO (for comparison)


Conclusion
For emulsions that are not sub For maximum emulsion stability
jected to elevated temperatures
we recommend the emulsifier
Lutensol XL 50 works perfectly.
Emulan AF. Lutensol TO 5 to 7 is
Also, Lutensol XP 50 suits nicely
the second best robust alter native.
here.
Performance of nonionic emulsifiers in emulsification of PDMS
Low
coalescence

Low
creaming

Small
droplets

Low
odor

Ecology

Emulan AF

++

++

++

Lutensol TO 5-7

++

++

Lutensol XL 50

+*

++

Lutensol XP 50

++

Emulan AT 9

++

++

++

Lutensol AT 11

++

++

++

APEO (for comparison)

++

++

* for extremely low odor choose Lutensol XA, the narrow range versions of Lutensol XL

In addition to perfect coalescence stabilization, the special nonionic emulsifier


Emulan AF offers good resistance against creaming plus a low odor.

31

3.6 Amino modified silicon microemulsions


made with nonionic surfactants from BASF,
e.g. for textile softening
Amino modified silicones
are perfectly suited for textile
softening
The ammonium group ...
... exhibits good affinity to the fabric,
thus the wash permanence is
good.
... offers the opportunity to make
silicon oil microemulsion concentrates with perfect shelf stability.
... offers the opportunity to make
silicon oil microemulsion concentrates which upon dilution form
fine nanoemulsions. Those are
able to penetrate the fiber bundle
very uniformly.

Making the microemulsion


concentrate
The literature cites a large number of
methods for making amino modified
silicon oil microemulsions. Those
include high- and low-energy input,
cold, hot and sometimes even
multi-step processes, with cofeed
of water, emulsifier or acids in
between.
All these processes are optimized
to overcome gel formation and to
achieve stable emulsions with small
droplet sizes. However, the simplest
processes can be sketched roughly
as follows:

Method 1
dissolve the nonionic surfactant
with water.
add the amino silicon oil and
emulsify as finely as possible
upon vigorous stirring add acid
to charge the amino oil. The
microemulsion will form automatically.

Method 2
mix the non-aqueous nonionic
surfactant with the silicon oil
and stir
upon stirring add acetic acid
to charge the amino oil
upon vigorous stirring add water
and emulsify
In both cases elevated temperatures
help to overcome the formation of
gels or gel particles.
Typical start values for the final
microemulsion are: 25% silicon oil,
12% nonionic emulsifier
(HLB = ca. 10 12), water to 100%,
adjust pH to 4.5.
Subsequently, optimize oil and
emulsifier content plus HLB.

32

Ranking of the nonionics


microemulsion
Definitions:
Efficiency is the required minimum
amount of nonionic emulsifier to
manufacture a microemulsion with
phase stability between 5 and 50C.
Tolerance window is the width of
the HLB window of the nonionic
emulsifier leading to a microemulsion
with phase stability between 5 and
50C.
Gelling tendency is the degree of
formation of gel lumps during
microemulsion manufacture, which
are difficult to dissolve.

Examples:
Microemulsions with different types
of amino oils are checked in a temperature interval of 5 to 50C to
determine both the efficiency and
the tolerance window of the nonionic
emulsifiers.

Clear, nonviscous microemulsion


concentrate

Tolerance window
Efficiency

18
17
16

Recipe:
25%
Silicon Oil 1.0 Pas/
0.6 mmol N/g
x%
Lutensol TO
75-x% Water
pH =
4.2

15

Please note that the microemulsion


phase diagram is highly dependend
on the oil type and needs to be
optimized for every formulation.

10

NPEO

Gel formation due to choice of


less suitable emulsifier

Lutensol TO
% Surfactant

High oil content suitability is the


possibility to form microemulsions
with both high oil content and low
viscosity.

Microemulsion

14
50C

13
12

o/w Emulsion

5 23C

11
9

Silicon oil A:
1.0 Pas, 0.3 mmol N/g

Silicon oil B:
1.0 Pas, 0.6 mmol N/g

solid/gel

solid/gel

Lutensol TO ca. 40,000 mPas

ca. 9,000 mPas

Lutensol XL

ca. 6,000 mPas

ca. 5,000 mPas

Lutensol XP ca. 2,000 mPas

ca. 3.000 mPas

10

11

12

13

14
HLB

High oil content suitability


Microemulsions with high oil content
have been optimized with respect to
both emulsifier content and HLB
value and their viscosity have been
determined. Systems with low viscosity offer the possibility to process
viscous amino oils or manufacture
micro emulsions of high oil concentration.
Experimental conditions:
40% oil, ca. 20% surfactant,
HLB ca. 10. Brookfield Spindle
4 6, 30 rpm, 23C.

33

Low gelling
tendency*

High oil content


suitability

Efficiency &
tolerance window

Lutensol TO

++

Lutensol XL

++

Lutensol XP

++

++

Lutensol ON

++

++

Conclusion:
ranking of the nonionic
surfactants with respect to
microemulsion formation
Please note that the table exhibits
the average trend. From oil to oil
slight variations may occur.

For comparison:
Nonyl phenol ethoxilate

Fatty alcohol ethoxilate

* Microemulsions are checked for the degree of


intermediate formation of gel particles and their
dissolution kinetics.

Ranking of the nonionics


diluted microemulsion under application conditions

Application
The microemulsion concentrate is
diluted to ca. 5 g/l silicon oil to be
applied on the fabric. Upon heating,
addition of salts, bases or shear
stress this diluted emulsion must
be stable, otherwise staining of the
fabric or silicon residues in the
machinery are possible. Emulsifiers
with additional high-speed wetting
power help to deaerate the fabric
during the impregnation step.

Stability against high


temperatures
The diluted microemulsion is
subjected to high temperatures
and stored or sheared. Instable
emulsions lead to fogging,
creaming or coalescence.

Silicon Oil 1.0 Pas/0.6 mmol N/g


with APEO
with Lutensol TO
storage at pH 10

Stability against high pH


The diluted microemulsion is
subjected to high pH and stored
or sheared. Instable emulsions
lead to fogging, creaming or
coalescence.

Performance under application conditions

Lutensol TO

++

++

++

Lutensol XL

Lutensol ON

++

Lutensol XP

++

Nonyl phenol ethoxilate

+ to ++

+ to ++

Fatty alcohol ethoxilate

++

++

++

For comparison:

34
Please note that the table exhibits
the average trend. From oil to oil
slight variations may occur.

**see next page

= Registered trademark of BASF group

Stability of the diluted emulsion


high temperature
MgCI2**
high pH

Stability against MgCl 2


The diluted microemulsion is
subjected to MgCl 2 and stored
or sheared. Instable emulsions
lead to fogging, creaming or
coalescence.

Silicon Oil 1.0 Pas/0.6 mmol N/g


with APEO
with Lutensol TO
storage in 1 w% MgCI 2

Conclusion amino modified


silicon oil microemulsions
For the manufacture of amino
modified silicon oil microemulsions
for textile softening we recommend
Lutensol TO and XL types. Both
offer easy formation of the micro emulsion. Lutensol TO additionally
offers optimum robustness in the
application plus good wetting and

Microemulsion

penetration. Lutensol XL offers highspeed wetting and thus very uniform


penetration, the formulation of
microemulsions of low viscosity plus
good robustness under application
conditions.

Robustness
during application

Wetting,
deaeration

Lutensol TO

++

++

Lutensol XL

++

++

Lutensol ON

++

Lutensol XP

++

++

Nonyl phenol ethoxilate

+ to ++

Fatty alcohol ethoxilate

++

For comparison:
Please note that the table exhibits
the average trend. From oil to oil
slight variations may occur.

35

3.7 Microemulsions

When an oil phase, an aqueous


phase and one or more surfactants
are mixed, then microemulsions can
be formed under certain conditions:
The hydrophobic parts of the
surfactants are miscible with the
oil phase
The hydrophilic parts of the
surfactants are miscible with the
aqueous phase
The surfactants are scarcely
miscible with the oil or water
phases on the molecular level
i.e. when the cmc* in oil and
water phase are exceeded,
the surfactants are behaving
amphiphobically.
Hydrophilic and hydrophobic part
of the surfactant or surfactant
mixture are approximately of equal
size; the surfactant film therefore
prefers planar geometries

In the microemulsion, oil and


water phases are still molecularly
separated by the surfactant film;
frequently bicontinuous structures
are formed, but lamellar and other
structures are also known.
Macroscopically one observes a
swelling of the surfactant phase by
oil and water phase; one therefore
says that the surfactant has solubilized the oil and water, and one
reports the solubilization parameters
S O and S W which describe the ratio
of the volumes of oil and water
phase, respectively, to the pure
surfactant in the microemulsion.
These values in efficient systems
may amount to more than S = 10
and correlate with the interfacial
tensions in the system. For high
solubilization parameters S micro emulsions appear opaque due to
the Tyndall effect; they are never theless thermodynamically stable.

The Krafft temperature of the


surfactants is exceeded
*cmc = critical micelle concentration

Water
Oil
0.1 m

Electron microphotograph of a frozen microemulsion,


Source: Science vol. 240, Microemulsions, Manfred Kahlweit. Copyright 1988, AAAS
[caption poorly legible]

36

Substantial differences
between microemulsions and
conventional (macro-)
emulsions
Microemulsions are thermo dynamically stable phases that
form spontaneously after the
mixing of all components without
an input of energy.
In microemulsions ultra-low interfacial tensions occur between the
oil and water phases.
Microemulsions react most
strongly to a change in intensive
magnitudes in the system, e.g.,
the temperature, the composition
of the oil and water phases as well
as the surfactant mixture. When
the range of existence of micro emulsions is exceeded, they often
transform into macroemulsions,
and this is an effective means of
producing the latter.
Microemulsions, as opposed to
conventional emulsions, separate
excess oil and water phases out
relatively quickly.
Microemulsions have maximal
stability at the phase inversion
point of the system.
Microemulsions are very dynamic;
boundary surfaces form and
disappear, usually within micro seconds.

Examples of applications
of microemulsions
Stable formulation of two nonmiscible liquid phases.
Formulation for production of
macroemulsions via spontaneous
emulsification.
Creation of a large contact area
between two non-miscible
phases, e.g. to facilitate material
transfers during dissolving and
cleaning processes, reactions or
extractions.
Generation of ultra-low interfacial
tension between two phases, e.g.,
to mobilize the one in the other.
Simultaneous transporting of
two non-miscible phases through
a porous medium, a fabric or
through narrow capillaries.

For use in the following areas:


Analytics
Soil sanitation
Fuels and propellants
Chemical reactions
Extraction
Corrosion protection
Cosmetics
Varnish
Leather industry
Metal working
Foods
Nanotechnology
Oil field
Plant protection
Pharmaceuticals
Cleaning
Textile industry
Environmental protection

37

Practical rules for


the preparation of
micro emulsions
The microemulsion is a stable phase
in the oil-phase, water-phase and
surfactant phase space. The practical
exploitation of this phase space,
however, is frequently made difficult
by kinetic inhibition of coalescence
by other, surfactant-richer phases,
often of higher viscosity, and precipitations of particularly ionic surfactants.
It has been shown empirically that
the simplest way to produce micro emulsions is with ethoxylates of
medium-chain branched alcohols.
The Lutensol ON brands are parti cularly useful for this, but also the
Lutensol XA, Lutensol XL or
Lutensol XP brands. It is best to
start with a 45% oil phase, 45%
water phase and 10% Lutensol;
as the starting point for water and
alkanes an HLB around 9 11 is
suitable. An addition of co-surfactants is also helpful; here branched,
short-chain alcohols with 4 6
carbon atoms are suitable.
Then the temperature is varied until
a single-phase region the micro emulsion or a three phase region
the microemulsion with additional oil
or water phase is established. Here
it is essential to distinguish between
the real three-phase region and
creamed, partially coalesced emulsion, which may have a similar
appearance.
One has now reached the phase
inversion. By varying the HLB value
of the surfactant or surfactant mixture, now the range of existence of
the microemulsion can be adjusted
to the desired temperature. A 10
Kelvin shift downward in the case of
ethoxylates corresponds to an HLB
value lower by about 0.5 1.0 units
and vice versa. The volume of the
microemulsion phase can finally be
expanded by increasing the surfactant content in the formulation until
the one-phase region is reached.

If the optimal HLB value of the


surfactant mixture is determined
with the medium-chain ethoxylates
Lutensol XA, Lutensol XL, Lutensol XP
or Lutensol ON, these may be
successively replaced with longerchain surfactants, such as, e.g.,
Lutensol TO, Emulan A,
Emulan AT 9 or Lutensol AT in order
to optimize the properties of the
micro emulsion.

Example of a microemulsion
In a 1 liter upright cylinder at 23C
one adds:

436.4g
112.8g
50g
5g
354.4 g

H 2O
NaCl
Lutensol TO 6 (HLB = 11)
sec-butanol
n-dodecane

Then the components are thoroughly


mixed and allowed to stand over night. Three phases are formed:
A dodecane phase on top, a salt
solution at the bottom and an
opaque microemulsion in the middle.
By varying the temperature, one can
observe how the ratios of dodecane
and salt solution in the microemulsion vary. For example, if one
touches the upright cylinder with
the hand in the region of the
microemulsion phase, the latter
will break up suddenly into a w/o
emulsion due to the warming effect.
The microemulsion can also be
removed and a stable, finely divided
O/W emulsion produced by injection
into the tenfold quantity of water.

Isotherm 23 C

38

Berhrung mit der Hand

Bereiche mit w/o Emulsion

Temperature-insensitive
microemulsions
To formulate microemulsions with a
broad thermal existence range, one
requires a surfactant mixture that
has the same space requirement
for its hydrophobic and hydrophilic
parts irrespective of the temperature. Here one will exploit the fact
that when ethylene oxide chains are
heated they successively lose their
hydrate shell and therefore occupy
less space, while in the case of ionic
groups, the Debye radius increases
and with it the space requirement.
A mixture of ethoxylate and ionic
surfactant is therefore recommended
for the formulation of such micro emulsions.
Empirically, Lutensit A-BO has
been shown to be of value as
an ionic partner. By combining
Lutensit A-BO with ethoxylates,
microemulsions with existence
ranges over 60 Kelvin, and more,
can be produced.

Example of a three phase


system expanded from
10 to 50C
19.23g of 1% aqueous
NaCI solution
14.62g n-dodecane
0.67g Lutensit A-BO
0.60g Lutensol AO 3

Further reading:
How to Study Microemulsions
Kahlweit, Strey, Haase, Kunieda,
Schmeling, Faulhaber, Borkovec,
Eicke, Busse, Eggers, Funck,
Richmann, Magid, Sdermann,
Stilbs, Winkler, Dittrich und Jahn,
J. Coll. Surf. Sci., 118, 2, 1987, 436
General Patterns of the Phase
Behavior of Mixtures of H 2O,
Nonpolar Solvents, Amphiphiles,
and Electrolytes. 1 Kahlweit,
Strey, Firman, Haase, Jen,
Schmcker, Langumir 1988, 4, 499
General Patterns of the Phase
Behavior of Mixtures of H 2O,
Nonpolar Solvents, Amphiphiles,
and Electrolytes. 2 Kahlweit,
Strey, Schomcker, Haase,
Langumir 1989, 5, 305
Nonionic Surfactants with Linear
and Branched Hydrocarbon Tails:
Composition Analysis, Phase
Behavior, and Film Properties in
Bicontinous Microemulsions
Frank, Frielinghaus, Allgaier, Prast,
Langumir 2007, 23, 6526
Interfacial Tensions and Solubilizing Ability of a Microemulsion
Phase that Coexists with Oil and
Brine Huh, J. Coll. Interface Sci.,
71, 2, 1979, 408

Microemulsions with extremely


polar oils

Interfacial Tensions in Micro emulsions Wennerstrm, Balogh,


Olsson, Coll. Surf. A: Physicochem.
Eng. Aspects, 291, 2006, 69

Even with extremely polar oils,


microemulsions can be formulated
provided that surfactants with a
correspondingly high HLB value
are selected.

Uses and Applications of Micro emulsions Paul, Moulik,


Current Science, 80, 8, 2001, 990

43.5g
3g
66g
37.5g

water
NaCl
ortho-sec-butylphenol
Lutensit TC-CS 40
(HLB = 41)

At 23C one obtains three phases,


the middle phase is the micro emulsion.

39

BASF SE
Care Chemicals & Formulators Europe
Carl-Bosch-Strae 38
67056 Ludwigshafen Germany
Phone: +49 621 60-0
Fax:
+49 621 60-42525
e-mail: industrial-formulators-eu@basf.com
www.care-chemicals-formulators.basf.com
BTC Speciality Chemical
Distribution GmbH
Maarweg 163 /165
50825 Kln Germany
Phone: +49 221 9 54 64-0
www.btc-de.com
North America
BASF Corporation
100 Campus Drive
07932 Florham Park, NJ USA
Phone: 800 526-1072
+1 973 245-6000
Fax:
+1 973 245-6002
www.performance.basf-corp.com
South and Central America

Safety
We know of no ill effects that could have resulted from using our products for the purpose for which they are intended and from processing
them in accordance with current practice. According to the experience
we have gained up to now and other information at our disposal, our
products do not exert any harmful effects on health, provided that they
are used properly, due attention is given to the precautions necessary
for handling chemicals, and the information and advice given in our
safety data sheet are observed.
Labeling
Details about the classification and labeling of our products and further
advice on safe handling are contained in the current safety data sheets.
Note
This document, or any answers or information provided herein by BASF,
does not constitute a legally binding obligation of BASF. While the
descriptions, designs, data and information contained herein are presented in good faith and believed to be accurate, it is provided for your
guidance only. Because many factors may affect processing or application/use, we recommend that you make tests to determine the suitability
of a product for your particular purpose prior to use. It does not relieve
our customers from the obligation to perform a full inspection of the
products upon delivery or any other obligation. NO WARRANTIES OF
ANY KIND, EITHER EXPRESS OR IMPLIED, INCLUDING WARRANTIES
OF MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE,
ARE MADE REGARDING PRODUCTS DESCRIBED OR DESIGNS, DATA
OR INFORMATION SET FORTH, OR THAT THE PRODUCTS, DESIGNS,
DATA OR INFORMATION MAY BE USED WITHOUT INFRINGING THE
INTELLECTUAL PROPERTY RIGHTS OF OTHERS. IN NO CASE SHALL
THE DESCRIPTIONS, INFORMATION, DATA OR DESIGNS PROVIDED BE
CONSIDERED A PART OF OUR TERMS AND CONDITIONS OF SALE.

08_100110e-00 07.2010 supersedes edition EMV 0091 - 00 e 09.2008

Europe, Africa, West Asia

BASF S.A.
AV. Faria Lima, 3600/3624
04538-132 So Paulo SP Brazil
Phone: +55 11 3043-2233
Fax:
+55 11 3043-6989
www.basf.com.br
Asia Pacific

= Registered trademark of BASF group

BASF East Asia


Regional Headquarters Ltd.
45th Floor, Jardine House
1 Connaught Place
Central Hong Kong China
Phone: +852 273 10111
Fax:
+852 273 49631
www.basf.com.sg

Anda mungkin juga menyukai