Sie sind auf Seite 1von 11

Journal of Membrane Science 169 (2000) 107–117

Membrane emulsification — a literature review


Simon M. Joscelyne, Gun Trägårdh ∗
Food Engineering, Lund University, Box 124, 22100 Lund, Sweden
Received 29 April 1999; received in revised form 22 October 1999; accepted 22 October 1999

Abstract
Membrane emulsification is a simple method that has received increasing attention over the last 10 years, with potential
applications in many fields. Experimental studies which have focused mainly on investigations of process parameters such as,
membrane type, average pore size and porosity, crossflow velocity, transmembrane pressure and emulsifier, are reviewed. By
careful choice of these parameters, emulsions with narrow emulsion droplet size distributions have been produced, with average
droplet sizes ranging between 2 and 10 times the nominal membrane pore diameter. The effects of individual parameters are
reasonably well understood, particularly at a qualitative level. Results can be explained by a direct influence of the membrane
pore size, diameter and distribution. Interfacial tension and the action of wall shear stress are also important.
In comparison with conventional turbulence based methods, such as homogenization and rotor-stator systems, less energy
is needed to produce droplets of a given size using membrane emulsification. However, one of the main limiting factors with
regard to industrial scale-up can be the often low level of dispersed phase flux through the membrane, especially for small
submicron droplets. ©2000 Elsevier Science B.V. All rights reserved.
Keywords: Membrane emulsification; Process parameters; Flux; Droplet size; Ceramic membranes

1. Introduction This method involves using a low pressure to force


the dispersed phase to permeate through a membrane
Emulsions play an important role in the formula- having a uniform pore-size distribution into the con-
tion of foods, cosmetics and pharmaceuticals. They are tinuous phase. The distinguishing feature is that the
traditionally prepared using colloid mills, rotor-stator resulting droplet size is controlled primarily by the
systems and high pressure homogenizers. The choice choice of membrane and not by the generation of tur-
is usually dictated by the application of the resulting bulent droplet break-up. The technique is highly at-
emulsion, the apparent viscosity, the amount of me- tractive given its simplicity, potentially lower energy
chanical energy required and the heat-exchange de- demands, need for less surfactant and the resulting nar-
mands [1]. But, over the last 10 years or so, there has row droplet-size distributions [3]. It is applicable to
been an increasing interest in a technique for making both oil-in-water (o/w) and water-in-oil (w/o) emul-
emulsions known as ‘membrane emulsification’ [2]. sions.
Work carried out in the field of membrane emul-
∗ Corresponding author. Tel.: +46-46-222-9821, sification is reviewed, discussing both theoretical
fax: +46-46-222-4622. and practical aspects, as well as indicating present
E-mail address: gun.tragardh@livstek.lth.se (G. Trägårdh). limitations.

0376-7388/00/$ – see front matter ©2000 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 6 - 7 3 8 8 ( 9 9 ) 0 0 3 3 4 - 8
108 S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117

Fig. 1. Schematic diagram of membrane emulsification process.

2. Process of membrane emulsification — may occur depending on the flow rate of continuous
overview phase and the contact angle between the droplet and
membrane surface [5]. The final droplet size and size
The process of membrane emulsification is shown distribution are not only determined by the pore size
in Fig. 1. The dispersed phase is pressed through the and size distribution of the membrane but also by the
pores of a microporous membrane, while the continu- degree of coalescence, both at the membrane surface
ous phase flows along the membrane surface. Droplets and in the bulk solution.
grow at pore outlets until, on reaching a certain size, A schematic picture of a typical small-scale, mem-
they detach. This is determined by the balance between brane emulsification apparatus, representative of those
the drag force on the droplet from the flowing contin- in the literature for making o/w emulsions, is shown in
uous phase, the buoyancy of the droplet, the interfa- Fig. 2. The system incorporates a tubular microfiltra-
cial tension forces and the driving pressure [4]. The tion membrane, a pump, a feed vessel, and a pressur-
droplet at a pore tends to form a spherical shape under ized (N2 ) oil container. The oil phase (to be dispersed)
the action of interfacial tension, but some distortion is pumped under gas pressure through the pores of the

Fig. 2. Schematic picture of a simple small scale membrane emulsification system.


S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117 109

membrane into the aqueous continuous phase which often +/−15% [6]. Ceramic ␣-Al2 O3 (e.g., Mem-
circulates through the middle of the membrane. braflow, Germany and Membralox, SCT France) or
Although much of the early work on membrane ␣-Al2 O3 coated with titainia oxide or zirconia oxide
emulsification is of a rather empirical nature, more have been used. By multicoating, the pore size can
systematic studies have been done and the important be reduced and the distribution made narrower. All
process controlling parameters and conditions identi- the above-mentioned membrane types come in many
fied. These are membrane pore size and distribution, different nominal pore sizes, typically ranging from
membrane porosity, membrane surface type, emulsi- about 0.05–14 ␮m and do not deform or compact
fier type and concentration, dispersed phase flux, ve- under the applied pressures.
locity of the continuous phase and transmembrane It has been generally observed that the droplet size
pressure. Transmembrane pressure 1P is defined as: (dd ) of an emulsion can be related to the pore size
(dp ) of the membrane by a linear relationship for given
Pd − (Pc,1 + Pc,2 )
1P = (1) operating conditions:
2
where Pd is the pressure of the dispersed phase out- dd = x dp (3)
side the membrane. Pc,1 and Pc,2 are the pressures
at both ends of the membrane module. The minimum where x can range typically from 2–10 [5,7,8].
emulsification pressure, i.e., the pressure at which the ‘Monodispersed’ emulsions can be produced if the
dispersed phase outside the membrane must be at to membrane pore-size distribution is sufficiently nar-
just begin to permeate through the membrane, can be row. The presence of a small number of coarse pores
calculated from the equation for capillary pressure: have been shown to lead to a bimodal distribution [9].
4 γ cos θ As a rule, the dispersed phase should not wet the
Pc = (2) membrane pores. This means that hydrophilic mem-
dp
branes are suited to making o/w emulsions and hy-
where dp is the pore diameter, γ is the oil/water inter- drophobic membranes for w/o emulsions. Nakashima
facial tension, and θ is the contact angle between the et al. [10] have shown that using a hydrophobized
dispersed phase and the membrane surface. Pc is also MPG membrane for making an o/w emulsion leads to
referred to as the critical pressure. a polydispersed emulsion with a larger average droplet
It is generally necessary to find a balance between size than when using a hydrophilic membrane. Sim-
all of the above-mentioned parameters to achieve the ilar behaviour [10] was found when sodium dodecyl
desired result, often considered as the formation of an sulfate (SDS) was used as the emulsifier with a posi-
emulsion with the smallest droplet size and narrowest tively charged membrane surface. The SDS adsorbed
size distribution. In reality, this will be dictated by to the membrane surface rendering it hydrophobic.
the demands of the product and the rate of emulsion The w/o emulsions have also been made using hy-
production. drophilic membranes. However, the resultant droplet
size is less than the pore size and is apparently de-
pendent on the structure of the pore outlets and not
3. Process parameters strictly the diameter [11].
Membranes purchased in a hydrophilic form can be
3.1. Membrane pore-size, distribution, porosity and made hydrophobic by chemical surface modification.
surface type This may create problems in certain circumstances
where silane coupling agents are banned, as in the
Many of the published investigations have been food industry. It may also be necessary to repeat the
made using tubular micro-porous glass (MPG) mem- surface treatment after each cleaning cycle. Katoh et
branes (Asahi Glass Company, Japan) and Shirasu al. [12] came up with the ingenious idea of pre-soaking
porous glass (SPG). These membranes are reputed a hydrophilic membrane in the oil phase to render
as having cylindrical, interconnected, uniform mi- it hydrophobic. Pre-soaking membranes in solutions
cropores. They have narrow pore-size distributions of surfactants or emulsifiers has also been shown to
110 S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117

improve dispersed phase flux [4]. The decreased


interfacial tension in the pores leads to better filling.
The porosity of a membrane is also important be-
cause it determines the distance between adjacent
pores, this distance increases as the porosity de-
creases. The closer the pores are together (at high
porosities) the greater the likelihood of droplet co-
alescence at the membrane surface before droplets
detach. Schröder et al. [4] found that a ratio of the
droplet size to pore diameter of >1.6, for a membrane
porosity of 0.3, led to a significant degree of coales-
cence. On the other hand, if the porosity is too low
then the dispersed phase flux may be insufficient for
viable emulsion production.
Fig. 3. Effect of wall shear stress on droplet size for oil-in-water
Some investigations have been done on droplet for-
emulsions produced using membrane emulsification, shown
mation from single pores. Kawakatsu et al. [13,14] schematically.
have used a silicon microchannel plate and Peng and
Williams [5] glass capillaries, to observe in situ forma-
tion of individual emulsion droplets using microscope rapidly and the size distributions became much
video systems. These investigations allow fundamen- broader because of coalescence at the membrane sur-
tal processes of droplet formation and detachment ki- face. Muschiolik and Dräger [17] and Muschiolik et
netics to be studied. The results are in agreement with al. [18] found that the droplet size was halved on in-
the findings using membranes. creasing the wall shear stress from 0.3 to 1.5 Pa using
a 0.5 ␮m MPG membrane. Katoh et al. [12] found
that a minimum wall shear stress of at least 0.48 Pa
3.2. Velocity of continuous phase
was needed to produce monodisperse oil droplets,
also using MPG membranes.
Droplets formed at the membrane surface detach The action of shear force may vary depending on
under the influence of the flowing continuous phase. the concentration of dispersed phase in the circulating
Typically, the crossflow velocity lies between 0.8 solution. If there is a significant change, this could
and 8 m s−1 . Its influence is often expressed in terms influence the droplet size. It has been shown [9] that up
of wall shear stress (Pa). The droplet size decreases to 30% oil in an o/w emulsion can have no observable
sharply as the crossflow velocity increases from rest influence.
and reaches a size where it becomes more or less
independent of the flow velocity [15,16], shown
schematically in Fig. 3. The largest change in droplet 3.3. Emulsifier
size occurs at small shear stresses. The flow velocity
can therefore be used to control droplet size. For very Emulsifiers have two main roles to play in the
small droplets, the sublayer thickness may start to formation of an emulsion. Firstly, they lower the
limit the effect of crossflow velocity. interfacial tension between oil and water. This fa-
Schubert and Schröder [15], using ceramic ␣-Al2 O3 cilitates droplet disruption and in the case of mem-
membranes to make o/w emulsions, found that the branes lowers the minimum emulsification pressure.
smaller the nominal membrane pore size the smaller Schröder and Schubert [15] have suggested that the
the minimum droplet size and the smaller the shear interfacial tension force is one of the essential forces
stress required to reach this diameter. A wall shear holding a droplet at a pore. They found that larger
stress >2 Pa was needed for the 0.1 and 0.5 ␮m droplets are produced the higher the equilibrium in-
membranes and >20 Pa for the 0.8 ␮m membrane. terfacial tension. Secondly, emulsifiers stabilize the
At smaller shear stresses the droplet size increased droplets against coalescence and/or aggregation. This
S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117 111

will depend on both the type of emulsifier and the icon microchannel to make o/w emulsions. The ad-
concentration. dition of surfactant had the effect of broadening the
Schröder et al. [4] considered the effects of dynamic pressure range over which stable monodispersed emul-
interfacial tension on droplet formation. They found sions could be produced. Similar observations were
that the faster emulsifier molecules adsorbed at newly made by Katoh et al. [12] who investigated the effects
formed interfaces, the smaller the emulsion droplets. of added surfactant in both oil and water phases dur-
Comparing fast adsorbing SDS to slow Tween 20, the ing production of o/w emulsions. With 0.3% sucrose
difference in droplet size was about a factor of six. ester in the water phase and 0.5% sorbitan ester in
Katoh et al. [19] investigated different emulsifiers, the oil phase the flux was increased for a given trans-
SDS, sucrose esters (SE), polyglycerol esters and membrane pressure when compared to the case of no
sodium caseinate in the water phase of corn oil and emulsifier in the oil phase.
kerosene o/w emulsions. Monodispersed emulsions Kandori et al. [21] made w/o emulsions using
could be prepared at surfactant concentrations where poly(oxyethylene-oxypropylene) surfactant, PE-64,
the interfacial tension had decreased to a constant and hydrophilic SPG membranes. The droplet size
minimum value (0.3% wt for the SE and 1.5% wt was smaller than the average membrane pore size and
for the sodium caseinate). However, there were dif- the droplets were monodispersed. The results were
ferences observed between sodium caseinate and the independent of surfactant concentration which varied
other emulsifiers. The droplet sizes for the latter re- from 2 to 10%. Under these conditions the interfa-
mained constant on increasing further the emulsifier cial tension was nearly zero and it was proposed that
concentration, whereas the droplet size decreased for spontaneous emulsification may provide part of the
the caseinate. explanation. For 7.5% added PE-64, the emulsions
Ban et al. [20] studied the preparation of o/w emul- showed no indication of creaming or coalescence af-
sions using poly(oxyethylene) hydrogenated castor oil ter a month, which was attributed to the viscoelastic
(EC-n) as an emulsifier in the aqueous phase and adsorbed surfactant film. The importance of interfa-
methylphenylpolysiloxane as the oil. By mixing dif- cial tension was confirmed in further investigations
ferent EC-n (n = 5 and 50) surfactants different HLB [11] which were carried out using SPG membranes of
values were produced from 9.8 to 12.8. The smallest different pore diameters and various copolymer-type
emulsion droplets were formed for a HLB value of surfactants PE-n (where n = 61, 62 and 64). The size
12.8 at 10% total surfactant concentration. The flux of of the water droplets and their stability was strongly
oil through the membrane was faster the shorter the dependent on the interfacial tension between the wa-
chain length or average chain length for the mixed sur- ter and oil phases. Stable, monodispersed emulsions
factant system, i.e. at lower HLB values, and was also could only be produced at an interfacial tension of
dependent on the concentration of surfactant. It in- <1 mN m−1 .
creased as the surfactant concentration increased from Muschiolik and Dräger [17] and Muschiolik et al.
0.5 to 2% but thereafter decreased at 5 and 10%. Pre- [18] investigated whey protein concentrate as an emul-
treatment of the membrane by immersion for 20 h in sifier for making o/w emulsions containing 10% soy-
the aqueous phase resulted in an increased flux, prob- bean oil-in-water. Proteins with different degrees of
ably as a result of better filling of membrane pores denaturation were studied. With the less denatured
from the lowered interfacial tension. protein (i.e., the more surface active) it was possible
Nakashima et al. [10] found that monodispersed to produce a monodisperse emulsion.
emulsions could be produced using SDS at concen-
trations much less than the critical micelle concen-
tration (cmc). At very small levels (0.5 mmol dm−3 ) 3.4. Emulsification pressure and flux of dispersed
larger droplets were formed because of coalescence. phase
The minimum emulsification pressure decreased as the
concentration of SDS was increased up to the cmc. The emulsification pressure controls the flux of
Similar behaviour was found by Kawakatsu et al. [14] dispersed phase through the membrane. Higher pres-
using SDS as the water phase emulsifier and a sil- sures are needed for membranes of smaller nominal
112 S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117

pore size, given that all other parameters are the the effects of temperature on membrane emulsifica-
same, because of the higher capillary pressures (see tion. The emulsification temperature will usually be
Eq. (2)). Typical values of transmembrane pressure dictated by the requirements of a product. In cases
for emulsification using 0.2, 0.5, and 0.8 ␮m pore [23] making o/w emulsions, where experimental op-
size membranes lie between about 20 and 500 kPa erating conditions have been significantly above room
for making o/w emulsions. Increasing transmembrane temperature (65◦ C), the temperature was employed
pressure increases the flux of dispersed phase through to make the oil phase more fluid, and to dissolve the
a membrane in accordance with Darcy’s law [4]. Any emulsifier. For w/o emulsions, heating the continu-
deviations from Darcy’s law that occur at low applied ous phase so that there is a substantial decrease in
pressures are because not all pores may open. It is viscosity, would make it easier to circulate.
difficult to predict emulsification pressures. Too high In studies by Katoh et al. [24], making w/o food
pressures lead to jets of oil and very large droplets. emulsions, it was found that changes in the viscosity
Too low pressures make the emulsification time long. of the continuous phase led to changes in the average
Williams et al. [9] have suggested that between 2 emulsion droplet size. Increasing the viscosity led to
and 10 times the minimum membrane pressure is an increase in size. It was noted that the flux was very
probably usable range. low. They [25] also encountered problems of crushing
Most of the studies already discussed above have of the highly viscous emulsions by the shear stress of
been limited to conditions suitable for producing pumps and rotors.
monodispersed emulsions. With such constraints, oil
fluxes for production of o/w emulsions have been 3.6. Effect of pH
found to range from about 2–20 l m−2 h−1 using hy-
drophilic membranes having an average pore size In general, pH is another parameter which is product
from 0.2 ␮m. Fluxes for 0.8 ␮m membranes have been dependent. However, membrane surface properties are
higher, up to about 40 l m−2 h−1 . In contrast fluxes often pH dependent. They may exhibit an iso-electric
of 2300 and 200 l m−2 h−1 [12] have been observed point at a given pH, where the surface has no net
for making w/o emulsions using 1.0 and 0.5 ␮m hy- charge. Huisman et al. [26] determined the iso-electric
drophilic membranes, respectively. These membranes points of a number of ceramic membranes which were
were pre-soaked in oil phase to make them hydropho- between pH 5.2 and >8. At pHs above and below these
bic. points, the membranes were negatively and positively
An increase in flux is often at the expense of droplet charged, respectively. The charge on the membrane
size, which may increase. Schröder et al. [4] have can have a profound influence on which surface ac-
shown that this increase depends on the type of emul- tive agents adsorb. It may be that a hydrophilic mem-
sifier or rather the adsorption kinetics and in the for- brane is rendered hydrophobic and as already seen
mation times of the droplets. If the formation time is above (see Membrane Pore–Size, Distribution, Poros-
long compared to the time of decreasing interfacial ity and Surface Type). This could have disastrous con-
tension, then the interfacial tension dynamics have lit- sequences on the resulting emulsion droplet size.
tle or no influence on the droplet size and the size
remains unchanged on increasing flux.
4. Comparison with other methods

3.5. Temperature and viscosity Some comparisons between membrane emulsifica-


tion and so-called ‘conventional’ rotating stirrer meth-
Temperature can be an important parameter in ods and homo-mixers can be found in the literature
emulsification affecting both the viscosity of the dis- [27–33]. But, these tend to focus on the superior sta-
persed and continuous phases and also the nature of bility and uniform droplet size and possibilities for
the emulsifier as a consequence of phase inversion producing tailor-made size distributions by mixing
temperature [22] and its solubility. To the authors’ two monodispersed emulsions, rather than analysis
knowledge there have been no systematic studies of and discussion of production rates and energy require-
S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117 113

ments. More systematic studies should be done ad- successfully in both batch or continuous mode. How-
dressing these aspects. Schubert [34] has compared the ever, to date there is only one documented product
performance of different types of continuous emulsi- produced using membrane emulsification, ‘Yes light’,
fication equipment, namely high pressure homogeniz- a low fat spread [7,12].
ers, rotor-stator systems and membrane emulsification, A small industrial membrane system such as the
in terms of energy densities Ev (J m−3 ) given the cri- Alfa Laval MFS-38 one has a membrane area of up
teria of smallest drop size and narrowest distribution: to 7.6 m2 [36]. For the oil fluxes reported earlier this
corresponds to throughput capacities of between 15
P
Ev = (4) and 300 l h−1 . The output capacities of industrial ho-
V mogenizers vary from about 100–20 000 l h−1 [37] and
where P is the effective power input and V is the typically lie around 1000 l h−1 [1] in the case of a
volume flow rate of emulsion. In all cases higher Manton–Gaulin commercial homogenizer. Thus pro-
energy densities were needed to produce smaller duction of 1000 l of o/w emulsion containing 10%
droplets. Droplet size ranged from about 100 ␮m oil would take typically an hour using an homoge-
down to 0.2 ␮m. For membrane systems the droplet nizer compared to between 20 min and 8 h using a
size increased for a given energy density as the con- membrane system. For preparation of small submicron
centration of dispersed phase (expressed as volume droplets, membranes having a pore size of 0.2 ␮m or
fraction) increased from 0.05 to 0.8. Energy densi- smaller would be needed. The low fluxes through such
ties for membrane emulsification (between 103 and membranes could be a limiting factor for large-scale
106 J m−3 ) were some 100 times less than those de- applications. Any advantages of low energy density
manded by high pressure homogenization (between may be lost at the expense of long production times.
105 and 108 J m−3 ) and some 10 times less than In practice, it is important to avoid a significant
rotor-stator methods (range of 105 –107 J m−3 ). At a tail in the droplet-size distribution function. It is the
similar energy density, membranes produce smaller presence of particularly large droplets in the fresh
droplets, because less energy is lost as heat. However, emulsion, and the growth in the number of such
if the dispersed phase viscosity and/or volume fraction droplets on storage, which generally have the predom-
is large, emulsification systems based on turbulent inant influence on the perceived instability. Thus, a
flow are more suitable. Droplet size distributions of balance between an acceptably high flux and a degree
turbulence-based methods tend to be broad given the of particle coalescence could be found, at the expense
nature of turbulence generation, making them more of monodispersity. Joscelyne and Trägårdh [23] used
sensitive to creaming and Ostwald ripening. crossflow velocity to keep oil droplets under 1 ␮m for
The shear stresses calculated for a membrane sys- an oil flux of 140 kg m−2 h−1 through a 0.1 ␮m pore
tem are much less than in the other methods, Mus- size membrane.
chiolik and Dräger. [17]. This makes it is possible to
use shear sensitive ingredients. Schubert and Schröder
[35] calculated shear stresses in the pores and found 6. Other applications
them comparable with those along the membrane sur-
face. Membrane emulsification has been used to make
monodispersed multiple emulsions, both o/w/o and
w/o/w emulsions. These have applications as drug car-
5. Industrial scale production riers for controlled release [38] and in the food indus-
try for microencapsulation of flavours [39]. Mine et
Membrane systems are particularly suitable for al. [8] successfully prepared w/o/w emulsions contain-
large scale production because it is easy to scale-up ing edible soybean oil. First a 30% (v/v) w/o emul-
and to add more membrane modules. Reproducibility sion was made having a mean droplet diameter of
of membrane systems is good and energy utiliza- 0.54 ␮m using a homogenizer. This was then added to
tion is efficient. Williams et al. [9] have shown that a water phase containing a emulsifier through a hydro-
pilot-scale membrane emulsification can be operated philic microporous glass membrane. Emulsions with a
114 S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117

Table 1
A summary of typical results and conditions for producing emulsions with small droplet sizes and droplet-size distributions obtained from
the literature. An underlined membrane pore size indicates that the value of dispersed phase flux applies to that membrane pore size
Membrane Pore characteristics/ Surface type Experimental w/o or o/w Droplet Ref.
nominal pore size conditions size, dp
(␮m)
SPG micro-porous 0.57, 1.1, 2.3 hydrophilic oil phase: corn oil
glass 1. water phase: water + 0.3% wt SE o/w 5dp [19]
flux: up to 30 l m−2 h−1
2. water phase: water + 2.0% wt sodium caseinate o/w 10dp
crossflow velocity > 0.2 ms−1
SPG micro-porous 0.98, 2.70, 4.70 hydrophilic oil phase: toluene + 2–10% wt
glass PE-64 copolymer w/o <dp [11„21]
Membraflow 0.1, 0.2, 0.8 hydrophilic oil phase: vegetable oil
ceramic
1. water phase: water +2% wt SDS o/w 3–4dp [15]
2. water phase: water + 0.1% wt Tween 20 o/w 8–12dp
flux: 10–40 l m−2 h−1
Membraflow 0.2, 0.8 hydrophilic oil phase: vegetable oil
ceramic
1. water phase: water + 0.7% wt LEO-10 o/w 15dp [4]
2. water phase: water + 0.1% wt lacprodan 60 o/w 50dp
flux: 5 l m−2 h−1
Single glass 5–200 hydrophilic oil phase: mineral oil + 3% wt iso stearic acid o/w 4dp [5]
capillaries water phase: water + 3% wt tri-
ethanolamine + 0.3% wt sodium nipastat
crossflow velocity: 0.5 ms−1
Ceramic 0.2 and 0.5 hydrophilic oil phase: mineral oil
water phase: water + 36.4% wt sorbital o/w 4dp [9]
(70%) + 2.16% wt dobanol 91-8 + 0.04%
wt formalin
flux: 7.4 l m−2 h−1
crossflow velocity: 1 ms−1
MPG microporous 0.2 and 0.5 hydrophilic oil phase: soybean oil
glass water phase: water + 2% milk protein concentrate o/w 8dp [17]
flux: 2.6 l m−2 h−1
crossflow velocity: 0.22 ms−1
SPG microporous 2.56 hydrophobic oil phase: soybean oil + 0.5% wt Span 80 w/o 3dp [2]
glass water phase: water + 1% wt NaCl
flux: 2.6 l m−2 h−1
Microporous glass 0.52 hydrophilic oil phase: kerosene
water phase: water + 0.04% wt SDS o/w 3dp [10]
flux: 0.005 l m−2 h−1
Ceramic 0.1 hydrophilic oil phase: vegetable oil + 8% v/v monoglyceride o/w 3dp [23]
water phase: skim milk
crossflow velocity 9 ms−1
flux: 140 l m−2 h−1
Microporous glass 0.36 and 1.36 hydrophilic oil phase: soybean oil
water phase: water + egg yolk proteins o/w 3dp [8]
1.0 hydrophobic oil phase: soybean oil
water phase: water + egg yolk proteins w/o 3dp
+ polyglycerol esters
flux: 0.005 l m−2 h−1
Silicon 6 ␮m diameter hydrophilic or oil phase: sunflower oil + 0.3% wt o/w and w/o 4dp [13„14]
microchannel channel hydrophobic sorbitan monolaurate
water phase: water + SDS
S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117 115

concentration of water droplets of 30–50% showed 8. Symbols


good stability over a period of 6 weeks. Kawashima
et al. [40] produced a multiple w/o/w emulsion with a 1P Transmembrane pressure (Pa)
smaller droplet size and a better distribution from an Pd Pressure of dispersed phase outside
original emulsion which was extruded through a poly- membrane (Pa)
carbonate membrane, before redispersing it in aqueous Pc,1 , Pc,2 Pressure at ends of membrane module
surfactant solution. W/o/w emulsions containing wa- (Pa)
ter soluble anticancer drugs have been prepared [38]. Pc Capillary or critical pressure (Pa)
In this case a w/o emulsion was produced by sonicat- γ Oil-water interfacial surface tension
ing a mixture of oil and saline containing the drug to (N m−1 )
form submicron-sized droplets. This was then passed θ Contact angle between membrane
through a hydrophilic membrane into a glucose solu- surface and dispersed phase
tion to produce the final w/o/w emulsion. Such emul- dp membrane pore diameter
sions have applications in the field of general target- dd Droplet diameter
ing chemotherapy. Okochi and Nakano [32] prepared Ev Energy density (J m−3 ) (kW m−3 h−1 )
w/o/w emulsions encapsulating cytarabine, doxoru- P Effective power input
bicine and vancomycin. Release of these drugs was V Volume flow rate of emulsion
slower from membrane produced emulsions than from
emulsions made by stirring methods.
The membrane emulsification method has also Acknowledgements
been used to produce uniform silica hydrogel [41]
and polymer microspheres [2]. Such polymer spheres This work was financed by Arla FoU AB, Advanced
have uses as stationary phases in HPLC [42] oper- Technology with Tetra Pak Processing Systems AB
ating in reversed phase mode. They show excellent and the Swedish National Board for Industrial and
column stability and give a relatively low pressure Technical Development through the Swedish Founda-
drop. Porous polymethyl methacrylate and crosslinked tion for Membrane Technology.
polystyrene spheres made using SPG membranes have
been used as immobilizing carriers of glucoamylase
[43]. References

[1] M.J. Lynch, W.C. Griffin, Food emulsions, in: Emulsions and
Emulsion Technology, Surfactant Science Series, vol. 6(1),
Marcel Dekker, New York, 1974, Chapter 5, pp. 249–289.
7. Conclusions [2] K. Kandori, Applications of microporous glass membranes:
membrane emulsification, in: A.G. Gaonkar (Ed.),
Membrane emulsification has been shown to be Food Processing: Recent Developments, Elsevier Science,
Amsterdam, 1995, pp. 113–142.
a relatively simple and reliable process for making
[3] E. Dickinson, Emulsions and droplet size control, in: D.J.
emulsions. The size of the droplets is largely depen- Wedlock (Ed.), Controlled Particle, Droplet and Bubble
dent on the size of the membrane pores, which means it Formation, Butterworth–Heinemann, Oxford, 1994, Chapter
is easy to select the membrane most suitable for the ap- 7, pp. 189–216.
plication. A summary of typical results and conditions [4] V. Schröder, O. Behrend, H. Schubert, Effect of dynamic
interfacial tension on the emulsification process using
for producing emulsions with narrow droplet-size dis- microporous ceramic membranes, J. Colloid Interface Sci.
tributions is presented in Table 1. 202 (1998) 334–340.
As a final comment, membrane emulsification may [5] S.J. Peng, R.A. Williams, Controlled production of emulsions
be more appropriate at the present time for production using a crossflow membrane, Part. Part. Syst. Charact. 15
of special ‘high technology’ products and applications, (1998) 21–25.
[6] Membrane emulsification by microporous glass, MPG Product
where there is a need for a high degree of droplet Specification Sheet.
size uniformity as opposed to a general method for [7] R. Katoh, Y. Asano, A. Furuya, K. Sotoyama, M. Tomita,
emulsification. Preparation of Food Emulsions using Membrane Emulsi-
116 S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117

fication System, in: Proceedings of the 7th International emulsification technique, Colloids and Surfaces 55 (1991)
Symposium on Synthetic Membranes in Science, Tübingen, 73–78.
Germany, 1994, 407. [22] D.J. Shaw, Emulsions and Foams, in: Introduction to Colloid
[8] Y. Mine, M. Shimizu, T. Nakashima, Preparation and and Surface Chemistry, 3 ed., Butterworths, London, 1980,
stabilization of simple and multiple emulsions using Chapter 10, pp. 232–244.
microporous glass membrane, Colloids Surfaces B: [23] S.M. Joscelyne, G. Trägårdh, Food emulsions using
Biointerfaces 6 (1996) 261–268. membrane emulsification: conditions for producing small
[9] R.A. Williams, S.J. Peng, D.A. Wheeler, N.C. Morley, droplets, J. Food Engineering 39 (1999) 59–64.
D. Taylor, M. Whalley, D.W. Houldsworth, Controlled [24] R. Katoh, Y. Asano, A. Furuya, K. Sotoyama, M. Tomita, S.
production of emulsions using a crossflow membrane Part II. Okonogi, Preparation of food emulsions using a membrane
Industrial scale manufacture, Chem. Eng. Res. Des. 76 A (8) emulsification system 2. Conditions for preparation of W/O
(1998) 902–910. food emulsions using a membrane emulsification system,
[10] T. Nakashima, M. Shimizu, M. Kukizaki, Membrane Nippon Shokuhin Kagaku Kaishi 44 (1997) 44–49.
emulsification by microporous glass, Key Engineering Mater. [25] R. Katoh, Y. Asano, A. Furuya, K. Sotoyama, M. Tomita,
61/62 (1991) 513–516. S. Okongi, Preparation of food emulsions using a membrane
[11] K. Kandori, K. Kishi, T. Ishikawa, Formation mechanisms of emulsification system 3. A method of preventing dispersed
monodispersed W/O emulsions by SPG filter emulsification droplets from crushing using a membrane emulsification
method, Colloids and Surfaces 61 (1991) 269–279. system, Nippon Shokuhin Kagaku Kokaku Kaishi 44 (3)
[12] R. Katoh, Y. Asano, A. Furuya, K. Sotoyama, M. (1997) 233–237.
Tomita, Preparation of food emulsions using a membrane [26] I. Huisman, G. Trägårdh, Ch. Trägårdh, A. Pihlajamäki,
emulsification system, J. Membr. Sci. 113 (1996) 131–135. Determining the zeta potential of microfiltration membranes
[13] T. Kawakatsu, Y. Kikuchi, M. Nakajima, Regular-seized using the electroviscous effect, J. Membr. Sci. 156 (1999)
cell creation in microchannel emulsification by visual 153–158.
[27] S. Okonogi, R. Kato, Y. Asano, H. Yuguchi, R. Kumazawa,
microprocessing method, J. Am. Oil Chemists’ Soc. 74 (3)
K. Sotoyama, K. Takahashi, M. Fujomoto, Morinaga
(1997) 317–321.
Milk Industry Co Ltd., Methods for producing emulsions,
[14] T. Kawakatsu, H. Komori, N. Oda, T. Yonemato, Relation
low-fat spread and oil-in-water-in-oil type spread, US Patent
between the concentration of a surfactant and pressure for
5,279,847 (1994).
droplet creation and effect on droplet size in microchannel
[28] S. Okonogi, R. Kumazawa, M. Kato, Y. Asano, Morinaga
O/W emulsification Kagaku Kogaku Ronbunshu 24 (2) (1998)
Milk Industry Co. Artificial milk preparation by pressure
313–317.
emulsification using membrane modules, JP 4-320643 1992.
[15] V. Schröder, H. Schubert, Emulsification using microporous,
[29] Y. Asano, R. Kumazawa, R. Kato, H. Yuguchi, K. Sotoyama,
ceramic membranes, in: Proceedings of the First European
K. Takahashi, M. Fujimoto, Morinaga Milk Industry Co
Congress on Chemical Engineering (ECCE 1), Florence, Italy,
Inc., Method for producing emulsions, low-fat spread and
4–7 May 1997, pp. 2491–2494.
oil-in-water-in-oil type spread, Eur. Patent Appl. 91303236.3
[16] V. Schröder, Z. Wang, H. Schubert, Production of oil-in-water (1991).
emulsions by microporous membranes, in: Proceedings of [30] M. Tomita, S. Kazuyoshi, R. Kato, Y. Asano, K. Takahashi,
the Third International Symposium on Progress in Membrane Morinaga Milk Industry Co Inc., A spread and a method
Science and Technology, Euromembrane 1997, University of for production of said spread, Eur Patent Appl. 92308937.9
Twente, 1997, 439. (1992).
[17] G. Muschiolik, S. Dräger, Emulsionsbildung mittels [31] S. Okonogi, R. Katoh, H. Yuguchi, Y. Asano, Morinaga Milk
mikroporösen Glas, Deutsche Milchwirtschaft 46 (19) (1995) Industry Co Inc., Method for producing emulsions, Eur Patent
1041–1048. Appl. 95200750.8 (1991).
[18] G. Muschiolik, S. Dräger, I. Scherze, H.M. Rawel, M. Stang, [32] H. Okochi, M. Nakano, Comparative study of two preparation
Protein-stabilized emulsions prepared by the micro-porous methods of W/O/W emulsion: stirring and membrane
glass method, in: E. Dickingson (Ed.), Food Colloids: emulsification, Chem. Pharm. Bull. 45 (8) (1997) 1323–1326.
Proteins, Lipids and Polysaccharides, Royal Society of [33] K. Sotoyama, Y. Asano, K. Ihara, K. Takahashi, K. Doi,
Chemistry, Cambridge, 1997, pp. 393–400. Preparation of W/O food product emulsions by the membrane
[19] R. Katoh, Y. Asano, A. Furuya, M. Tomita, Conditions emulsification method and assessment of stability, Nippon
for preparation of O/W food emulsions using a membrane Shokuhin Kagaku Kogaku Kaishi 45 (4) (1998) 253–260.
emulsification system, Nippon Shokuhin Kagaku Kogaku [34] H. Schubert, Advances in the production of food emulsions,
Kaishi 42 (1995) 548–555. in: Proceedings of Engineering and Food at ICEF
[20] S. Ban, M. Kitana, A. Yamasaki, Preparation of O/W 7, ICEF-Conference, Academic Press, Brighton, 1997,
emulsions with poly(oxyethylene) hydrogenated castor oil by AA82–AA87.
using SPG membrane emulsification, Nippon Kagaku Kaishi [35] H. Schubert, V. Schröder, Mechanical processes with
8 (1994) 737–742. minimum energy input 1. Emulsions, in: Proceedings of 4th
[21] K. Kandori, K. Kishi, T. Ishikawa, Preparation of EFFast- Conference, Köln, Germany, 6–9 November 1996,
monodispersed W/O emulsions by Shirasu-porous-glass filter in press.
S.M. Joscelyne, G. Trägårdh / Journal of Membrane Science 169 (2000) 107–117 117

[36] Alfa-Laval Filtration Systems, Cross-flow microfiltration loop, water/oil/water emulsion droplets with porous membrane, J.
Type MFS-38 data sheet. Colloid Interface Sci. 145 (1991) 512–523.
[37] L.W. Phipps, The High Pressure Dairy Homogenizer, [41] K. Kandori, K. Kishi, T. Ishikawa, Preparation of uniform
Technical Bulletin 6, NIRD, Reading, England, 1985. silica hydrogel particles by SPG filter emulsification method,
[38] S. Higashi, M. Shimizu, T. Setoguchi, Preparation of new Colloids Surfaces 26 (1992) 259–262.
Lipiodol emulsion containing water soluble annticancer [42] K. Hosoya, K. Yoshizako, K. Kimata, N. Tanaka,
agent by membrane emulsification technique, Drug Delivery Chromatographic properties of uniformly sized macroporous
Systems 8 (1993) 59–61. polymer particles prepared using SPG emulsification,
[39] A.G. Gaonkar, Kraft General Foods Inc., Stable Chromatography 18 (4) (1997) 226–227.
multiple emulsions comprising interfacial gelatinous layer, [43] S. Omi, K. Kaneka, A. Nakayama, K. Katami, T. Taguchi,
flavor-encapsulating multiple emulsions and low/no fat food M. Iso, M. Nagai, M.G.-H. Ma, Application of porous
products comprising the same, US Patent 5332595 (1994). microspheres prepared by SPG emulsification as immobilizing
[40] Y. Kawashima, T. Hino, H. Takeuchi, T. Niwa, K. carriers of glucamylase (GLuA), J. Appl. Polym. Sci. 65 (13)
Horibe, Shear-induced phase inversion and size control of (1997) 2655–2664.

Das könnte Ihnen auch gefallen