You are on page 1of 11

CHERD-1723; No.

of Pages 11
ARTICLE IN PRESS
chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

The application of a multi-scale approach to the


manufacture of concentrated and highly
concentrated emulsions

Diego Pradilla, Watson Vargas, Oscar Alvarez ∗


Grupo de Diseño de Producto y de Proceso (GDPP), Departamento de Ingeniería Química, Universidad de los Andes,
Carrera 1 N 18A-12, Edificio Mario Laserna, Piso 7, Bogotá, Colombia

a b s t r a c t

In this work, a multi-scale approach is applied to the emulsification process of concentrated emulsions. Whereas
a vast number of studies have focused on either formulation or processing (Mason et al., 1997; Pal, 1999; Dimitrova
and Leal-Calderon, 2004; Derkach, 2009; Capdevila et al., 2010), we propose a transversal study that includes those
variables and also covers aspects of the rheological behavior, the droplet size and the near infra-red spectra (NIRS).
The analysis on oil-in-water (O/W) and water-in-oil (W/O) emulsions yielded results that depend on the energy
incorporated during the emulsification process. First, depending on the tip velocity and pumping capacity of the
impeller, a certain amount of energy can be incorporated to a given formulation (i.e., concentration of the dispersed
phase). Second, as a consequence, the properties (rheology and droplet size) (Zölzer and Eicke, 1993; Ewoldt and
McKinley, 2007; Evans et al., 2009) of the resulting emulsion are set and mathematical relationships are established.
Finally, it is possible to visualize via NIRS not only the changes in concentration (Aske et al., 2002; Sjöblom et al.,
2003; Araujo et al., 2008) but also the influence of the changes in droplet size and couple all three aspects.
The novelty of these results rests on the treatment of the energy as a transversal variable to the scales studied
instead of handling only the formulation variables.
© 2014 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords: Emulsions; Mixing; Rheology

1. Introduction a surfactant that stabilizes the droplets by ensuring a short


range repulsive interaction at the interface (Dimitrova and
Emulsions are a type of colloidal system in which droplets of Leal-Calderon, 1999; Datta et al., 2011). Highly concentrated
a liquid are dispersed in a second immiscible liquid. Emul- emulsions also known as high internal phase ratio emul-
sions are widely used in a variety of industrial applications: sions (Pal, 1996,1998a,b,1999,2006; Clausse et al., 2007) and
cosmetics, food, pharmacy, coatings, oil recovery and the so- concentrated emulsions are of special interest because of the
called liquid explosives (Jager-Lézer et al., 1998; Clausse et al., complexity of their rheological behavior resulting from the
1999; Dimitrova and Leal-Calderon, 2001,2004; Sjoblom, 2001; increasing relevance of the interactions between the droplets
Tadros, 2004; Pal, 2006; Masalova and Malkin, 2008; Capdevila as they come closer to one another. In most cases, this proxim-
et al., 2010; Bouyer et al., 2011; Masalova et al., 2011). These ity causes the spherical droplets to change in shape and take
products are non-equilibrium systems, which means that they the shape of polyhedrons.
are thermodynamically unstable and consequently two vari- A multi-scale approach could be defined as a way of creat-
ables must be considered for their formation: energy (which ing a link between what happens within a system (in this case,
must be incorporated through the emulsification process) and an emulsion) and its performance as a finished product. The


Corresponding author. Tel.: +57 1 3394949x3935; fax: +57 1 3324334.
E-mail addresses: d-pradil@uniandes.edu.co (D. Pradilla), wvargas@uniandes.edu.co (W. Vargas), oalvarez@uniandes.edu.co
(O. Alvarez).
Received 28 February 2014; Received in revised form 6 October 2014; Accepted 23 October 2014
http://dx.doi.org/10.1016/j.cherd.2014.10.016
0263-8762/© 2014 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
2 chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx

practical use of such link (or links) has been acknowledged by size and G is the value of the elastic modulus in the linear
some authors using several field specific examples (Jager-Lézer viscoelastic region. Eqs. (1) and (2) indicate how the elasticity
et al., 1998; Sagis, 2011; Charton et al., 2013). and the average droplet size change with respect to the incor-
Recently, Sagis (2011) pointed out that the dynamic behav- porated energy for W/O emulsions when the concentration of
ior of the interfaces of high surface to volume ratio systems, the dispersed phase is between 92 and 96 (wt%).
such as emulsions (also dispersions of vesicles, dispersion of The second part of this work, studies how average droplet
nanocapsules, etc.) is responsible for the macroscopic dynam- size affects the near infrared spectra (NIR) of O/W emulsions.
ics. Therefore, a study solely focused on the molecular scale Similarly, the effect of the concentration of the dispersed
(i.e., surface active material adsorption, desorption kinetics, phase on the different trends will also be studied for concen-
diffusion mechanisms, surface rheology, etc.) would not suf- trations of 30–80 (wt%).
fice if the goal of the study was to evaluate overall performance According to Sjøblom (Aske et al., 2002; Sjöblom et al., 2003),
of a product in different scenarios. In fact, interactions at the when the near infrared spectra (NIR) spectra of a water-in-oil
molecular scale are only a part of what defines the behavior emulsion is obtained, the water droplets within the emulsion
of emulsions. will scatter portions of the incoming light with an intensity
Analogously, at the macroscopic level, as highlighted by that follows Eq. (3):
Dimitrova and Leal-Calderon (2004), Babak et al. (2001) and
more recently by Derkach (2009), several factors affect the Id r6
behavior of emulsions: the volume fraction of the dispersed ∝ 2 4 (3)
I0 x 
phase, the droplet size, the polydispersity of the droplets, the
temperature and the emulsification process. This behavior
where r is the droplet radius, x is the sample thickness,  is
is usually studied using rheological measurements such as
the wavelength and Id /I0 is a measurement of the extinction
steady-state flow experiments to determine the viscosity as
of light. This means that for two water-in-oil emulsions that
a function of the shear rate and thus observing both the shear
have the same formulation but differ in their droplet size dis-
thinning behavior and the yield stress (Pal, 1999) -and oscil-
tributions, the emulsion with bigger droplets will scatter light
latory experiments in which the elastic and viscous moduli
with a greater total intensity than the emulsion with smaller
(G and G ) are obtained (Mason et al., 1997; Fa et al., 2004).
droplets. This will shift the spectra toward areas of higher
Also, the droplet size and droplet size distribution are usu-
absorbance for the emulsion with greater average droplet size.
ally obtained through optical microscopy or light diffraction
Araujo et al. (2008) observed similar changes in the NIR spectra
techniques.
of water-in-crude oil emulsions due to changes in the average
The overall goal of this work is to use a novel multi-scale
droplet size and concentration of water. However, the trends
approach, which analyzes the coupled effects between, first,
of the different spectra were opposite to those reported previ-
the formulation and the composition variables of a product
ously (Aske et al., 2002; Sjöblom et al., 2003).
(i.e., an emulsion), second, the process through which the
Finally, the third part of this work will use creep tests
product is made and third, the properties that the product
to study concentrated and highly concentrated W/O emul-
exhibits. These properties can be studied using a multi-level
sions and analyze the behavior of the compliance modulus,
method, meaning that the responses at different levels can be
J(t), when the concentrations of the dispersed phase and
analyzed. In this sense, the product will have properties at the
the applied stress within the linear viscoelastic region are
macroscopic level, described in this case by the rheology of the
changed.
system, the microscopic level, related to the average droplet
Creep testing has been commonly used to determine the
size, and the molecular level, represented in this case by the
yield stress at which some materials may fracture or to deter-
combinations and overtones in a near infrared spectrum. This
mine the viscosity of soft materials at low shear rates (Tadros,
novel approach allows the simultaneous study of aspects at
2004). Typically, oscillatory measurements are used to obtain
the three mentioned levels, covering the main variables that
reliable results in the short-time/high frequency domain and
influence emulsion preparation.
creep measurements are used to obtain information in the
The first part of this work establishes the relationships
long-time/low-frequency domain. However, creep testing can
between the energy incorporated through the emulsification
also be used to obtain information about the viscoelastic
process, the concentration of the dispersed phase and the
nature of concentrated and highly concentrated emulsions
properties at the three levels, using three different types
(Zölzer and Eicke, 1993; Baravian and Quemada, 1998; Jager-
of impellers to prepare oil-in-water (O/W) emulsions with
Lézer et al., 1998; Ewoldt and McKinley, 2007). Therefore, the
concentrations ranging from 80 to 90 (wt%). Using a simi-
information obtained through a creep experiment contains
lar approach, Alvarez et al. (2010) established a relationship
essentially the same information as oscillatory measure-
between the energy incorporated through the emulsification
ments, and accordingly, changes in the morphology of the
process and the macroscopic properties (the rheology) of
droplets can also be observed using a different technique.
water-in-oil (W/O) emulsions. They also established a second
relationship that relates the energy incorporated to the micro-
scopic structure of the system (the average droplet size). The 2. Materials and methods
relationships are shown in the following equations:
2.1. Materials
G ∝ E0.6
v (1)
Highly concentrated oil-in-water (O/W) and water-in-oil (W/O)
Rm ∝ E−0.3
v (2) emulsions were prepared with mineral oil (USP-grade), mili-Q
de-ionized water and two non-ionic surfactants provided by
In these equations, Ev corresponds to the energy incorporated Croda: Span 80® (Sorbitan Monooleate), oil soluble, HLB 4.3,
through the emulsification process, Rm is the average droplet and Tween 20® (polisorbate 20), water soluble, HLB 16.7.

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx 3

Table 1 – Relevant dimensions of the mixing system. Table 2 – The characteristic peaks of an emulsion.
Values taken from (Sjöblom, Aske et al., 2002).
Impeller Impeller Tank Height of
diameter diameter the finished Absorption band Wavelength
[mm] [mm] emulsion region [nm]
[mm]*
O H first overtone 1400–1450
“PB”: Pitch Blade 45◦ 104 133 48 O H combinations 1900–1975
“HF”: Hydrofoil 85 108 48 C H second overtone 1125–1225
“PR”: Propeller 65 83 48 C H combinations first overtone 1350–1450
C H first overtone 1625–1775

Relative to the bottom. C H combinations 1950–2450

to measure the temperature at all stages of the emulsification


2.2. Methods
process. The variation was less than 1 ◦ C in all experiments.
Part II: The procedure described in Part I was used with a
2.2.1. Formulation
minor change: only the propeller was used.
Stability tests showed that 4 (wt%) of total concentration of the
Part III: The procedure described in Part I was used with a
surfactant mixture was the most suitable for the study. At this
minor change: only the propeller was used.
value, the phase separation vs. time data showed a maximum
indicating the most stable system. This corresponds to a mass
ratio of Tween 20® to Span 80® for O/W emulsions of 7/3 and 2.3. Characterization
78/5 for W/O emulsions. The corresponding final HLB values
are 15 and 5 respectively. All measurements were performed 30 min after preparation to
Drop tests were performed after each preparation to verify allow the emulsions to adequately stabilize. High reproducibil-
the continuous phase. ity was achieved by testing duplicates for each emulsion.
Part I: O/W emulsions were prepared in a concentration The experimental error was no higher than 10%. The char-
range of 80–90 (wt%). acterization consisted of three parts: rheology, near infrared
Part II: O/W emulsions were prepared in a concentration spectroscopy and droplet size measurement.
range of 30–80 (wt%).
Part III: W/O emulsions were prepared in a concentration 2.3.1. Rheology
range of 75–85 (wt%). W/O emulsions were chosen for this part Part I: A controlled-stress rheometer (ARG2, TA instruments)
because their elastic response in the linear viscoelastic region equipped with a 20 mm-parallel plate geometry was used to
is more strongly pronounced than that of O/W emulsions. perform 2 oscillatory tests at a constant gap of 1 mm and
constant temperature of 20 ◦ C with a maximum deviation of
2.2.2. Emulsification process ±0.1 ◦ C. An extra set of oscillatory tests was carried out while
Before the preparation of the emulsions, the surfactants were varying the gap size (500–1500 ␮m) to ensure that there was
added to the corresponding soluble phase separately: Span 80® no wall-slip. These results were not included in the analysis.
was pre-mixed with oil and Tween 20® was pre-mixed with The first oscillatory test was a frequency sweep with a
water. pre-shear of 1 min at 1 s−1 , followed by the frequency step
Part I: A semi-batch process was used to prepare all emul- of 1–100 rad/s. The second oscillatory test was a stress sweep
sions. This process consisted of three steps: In the first with a pre-shear of 1 min at 1 s−1 , followed by the stress step
step, the continuous and the dispersed phase were pre- of 1–100 Pa. These tests allowed for the obtention of the value
homogenized separately for 15 min at 300 rpm. In the second of the elastic modulus (G ) in the linear viscoelastic region. In
step, the dispersed phase was added to the continuous phase our case, this means that G is independent of both the stress
at a slow, constant rate using a peristaltic pump (Fischer Sci- and the frequency, thus the value reported corresponds to the
entific, flow rate = 0.5 ml/s). In the third step, the emulsion was average in the plateau region. No creaming or other signs of
homogenized for 10 min at the same tip velocity (1.7 m/s). instability were observed during the experiments.
These three steps were performed in a mixing device (Light- Part II: No rheological measurements were performed.
nin LabMaster L1U10F) that allowed for the torque to be Part III: A controlled-stress rheometer (ARG2, TA instru-
recorded as a function of time, and from this data, it was pos- ments) equipped with a 20 mm-parallel plate geometry was
sible to calculate the power generated. The area under the used to perform creep tests at a constant gap of 1 mm and
power-time curve was later calculated (via integration); this constant temperature of 20 ◦ C, with a maximum deviation of
value corresponds to the energy consumed per unit volume ±0.1 ◦ C. An extra set of oscillatory tests was carried out while
for each emulsion. varying the gap size (500–1500 ␮m) to ensure that there was
Three different types of impellers were used: a pro- no wall-slip. These results were not included in the analysis.
peller, a 45◦ pitch blade turbine and a hydrofoil turbine. The The creep tests were carried for 300 s without recovery, and
impeller-to-tank diameter ratio was kept constant at 0.78. The the stress varied from 1 to 125 Pa.
off-bottom clearance was defined by placing the impeller at
the liquid-air interface prior to the incorporation of the dis- 2.3.2. Near infrared spectroscopy (NIRS)
persed phase. The total final height of the emulsion was also Part I: The absorbance of the emulsions was measured in the
kept constant at 48 mm relative to the bottom of the tank to near infrared region (1100–2500 nm) using a smartprobe ana-
avoid head pressure effects on the impeller. Table 1 shows rel- lyzer spectrophotometer (FOSS). The sample size was 12 ml.
evant dimensions of the mixing system. The tip velocity was Table 2 shows the different characteristic peaks and their cor-
also fixed at a constant value and a thermometer was used responding functional groups.

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
4 chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx

The parameter chosen to characterize the O/W emulsions


was the maximum located at 1625–1775 nm, which corre-
sponds to the first overtone of the C H (stretching vibrations).
This is a good representation of the oil droplets dispersed in
the continuous phase for the O/W emulsions. This parameter
has the advantage of being dependent on the concentration
of the absorbing group and largely independent of the rest of
the molecule, avoiding induced noise from other groups (Kelly
and Callis, 1990).
Part II: The procedure described in Part I was used without
any changes for this section.
Part III: No spectroscopy measurements were performed.

2.3.3. Average droplet size and droplet size distribution


Part I: The droplet size distribution was obtained using a Mas-
tersizer 3000 (Malvern Instruments), which uses the laser
diffraction technique to perform droplet size measurements Fig. 1 – Elastic modulus (G in the linear viscoelastic region:
in the range from 0.01 to 3500 ␮m. The governing principle 0.1–100 [rad/s] and 0.1–100 [Pa]) as a function of the energy
for these measurements is the Mie theory. To avoid multi- incorporated (homogenization step) to the system for O/W
ple scattering, adequate ranges of obscuration were selected emulsions with a concentration of the dispersed phase in
according to the laser specifications. All measurements were the range of 80–90 (wt%) prepared using three different
performed under the same conditions. impellers: 45◦ Pitch blade, Hydrofoil and Propeller. (The
Mie theory considers the diffraction, refraction and absorp- concentration values in ascending order are 80%, 82%, 85%,
tion events that occur when a beam of light impacts a particle. 87% and 90%).
For this reason, the optical parameters must be known. The
dispersant, which carries the particles or droplets to the mea-
elasticity of the system affects the amount of energy neces-
surement area, must not interact in any way with the droplets
sary for homogenization, that is, when the concentration of
to avoid changes in size; for this reason, water was the dis-
the dispersed phase increases, the energy incorporated also
persant for O/W emulsions (Sjöblom et al., 2003; Araujo et al.,
increases as long as the other variables related to the process
2008).
are kept constant (i.e., the geometry of the mixing system and
From the intensity reports, which are related to the vol-
the tip velocity of the impeller).
ume that the particles occupy, various mean diameters can be
The relationship between the rheological behavior of
obtained. One of these is the volume mean diameter, D[4,3],
concentrated and highly concentrated emulsions and the
which is the mean diameter used in this work.
average droplet size (droplet size distribution) has also been
Part II: The procedure described in Part I was used without
the subject of numerous publications (Jager-Lézer et al., 1998;
any changes.
Pal, 1998a,b; Malkin et al., 2004; Mougel et al., 2006; Pal,
Part III: No droplet size measurements were performed.
2006; Derkach, 2009; Capdevila et al., 2010). A comparison of
Figs. 1 and 2, confirms that when the elasticity of the system
3. Results and discussion

3.1. Part I: multi-scale approach

The first part of this discussion focuses on describing and


understanding the possible relationships between the energy
incorporated through the emulsification process and the
macroscopic properties, the microscopic structure and the
molecular response of oil-in-water emulsions. It is important
to highlight the fact that the energy incorporated through
the emulsification process corresponds to the energy that is
required to homogenize the product after the dispersed phase
has been added. In the preparation of concentrated and highly
concentrated emulsions, this stage requires more energy than
the incorporation step (Alvarez et al., 2010).
Different publications have reported that when the vol-
ume fraction of the dispersed phase increases, the elasticity
of the concentrated and highly concentrated emulsions also
increases (Salager, 1996; Langevin, 2000; Dimitrova and Leal-
Calderon, 2004; Pal, 2006). This increase in the elasticity is Fig. 2 – Average droplet size (D[4,3] ␮m) as a function of the
directly related to the degree of packing of the droplets, which energy incorporated (homogenization step) to the system
promotes the interactions among them. The consequence for O/W emulsions with a concentration of the dispersed
of the interactions between the droplets is an increase of phase in the range of 80–90 (wt%), prepared using three
the elastic modulus in the linear viscoelastic region. The different impellers: 45◦ Pitch blade, Hydrofoil and Propeller.
emulsions prepared for this work exhibited the behavior just (The concentration values in descending order are 80%,
described as shown in Fig. 1. Evidently, an increase in the 82%, 85%, 87% and 90%.)

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx 5

Table 3 – Values of the elastic modulus (G ), average droplet size (D[4,3]) and energy consumption for direct oil-in-water
emulsions with different concentrations of the dispersed phase prepared with three impellers: Pitch blade “PB”,
Hydrofoil “HF” and Propeller “PR”. Same data plotted in Figs. 1 and 2.
wt% Impeller Energy [J/ml] G [Pa] D[4,3] ␮m
80 PB 1.00 ± 0.06 3.85 ± 0.07 17.55 ± 0.07
HF 2.10 ± 0.06 5.62 ± 0.82 14.30 ± 0.56
PR 4.80 ± 0.06 9.37 ± 0.76 10.60 ± 0.07

82 PB 1.05 ± 0.01 5.86 ± 0.98 14.05 ± 0.49


HF 2.11 ± 0.02 7.78 ± 0.65 10.80 ± 0.28
PR 4.89 ± 0.03 10.16 ± 1.05 8.91 ± 0.39

85 PB 1.33 ± 0.06 14.97 ± 1.71 8.32 ± 0.67


HF 2.66 ± 0.07 22.75 ± 4.03 6.51 ± 0.50
PR 6.01 ± 0.28 29.14 ± 5.57 4.81 ± 0.03

87 PB 1.46 ± 0.12 31.00 ± 0.56 5.11 ± 0.11


HF 2.99 ± 0.16 35.25 ± 1.90 4.47 ± 0.15
PR 7.05 ± 0.04 53.75 ± 1.24 3.78 ± 0.06

90 PB 2.87 ± 0.04 169.47 ± 10.92 2.50 ± 0.26


HF 5.33 ± 0.09 220.11 ± 5.07 1.80 ± 0.17
PR 12.24 ± 0.17 322.57 ± 91.11 1.27 ± 0.04

(represented by the elastic modulus G ) increases, the average et al., 1998; Sjöblom et al., 2003), this is directly proportional to
droplet size (D[4,3]), decreases. This leads to the conclusion the elasticity of the system and the amount of energy needed
that increased energy consumption causes a decrease in for homogenization. Second, changes in the average droplet
the average droplet size. This conclusion is also valid when size cause shifts in the overall spectra of the emulsions. This
the concentration of the dispersed phase increases (see also result will be further studied in a later section of the present
Table 3 for direct value comparison). article.
Moreover, it can be observed that the changes at the macro- The effect of any variable related to the process can be
scopic level (indicated by the rheological properties) and the evaluated through the relationships described previously (Eqs.
microscopic level (indicated by the droplet size) due to the (1) and (2)). Such a variable can be, for instance, the type of
energy incorporated during the emulsification process, are impeller: pitch blade, hydrofoil or propeller. If the concen-
also evident in the molecular response of the emulsions. Fig. 3 tration of the dispersed phase is fixed at a given value (see
shows how the absorbance of the first overtone of the C H Figs. 1–3 and Table 3), it can be seen that the values of the
bond interactions behaves as a function of the incorporated elastic modulus, G , the average droplet size, D[4,3], and the
energy. This increment is a result of two conditions: first, an absorbance of the first overtone are similar and do not seem
increase in the phase volume of oil causes an increase in the to depend on the type of impeller used for preparation (i.e., all
absorbance of its characteristic peak (Gossen et al., 1993; Frake the values of G are within the same order of magnitude). For
better clarity and to avoid misinterpretations caused by solely
looking at Figs. 1 and 2, the data used to construct these figures
are shown in Table 3. From this table, it can be observed that
all the values of the elastic modulus are within the same order
of magnitude, which means they are comparable. Something
similar happens when looking at the values of the average
droplet size. However, it is also clear that the energy consump-
tion does, in fact, depend on the type of impeller. The values
of energy consumption for the propeller are always signifi-
cantly higher compared to the hydrofoil and the pitch blade.
This is a direct consequence of the relevance of the pumping
capacity as will be addressed further on this section. This also
means that there is a similar rheological behavior associated
with a similar microscopic structure and molecular response
for different energy consumption values.
To attempt an explanation of this result, it is important to
highlight two facts: first, the variables related to the geom-
etry of the mixing system were kept constant; that is, the
impeller-to-tank diameter, the total final emulsion height and
Fig. 3 – Total measured absorbance of the 1st overtone of the location of the impeller inside the mixing tank were kept
CH-bond interactions as a function of the energy constant. The tip velocity was also kept constant to assure
incorporated (homogenization step) to the system for O/W that the same shear was given by impellers with different
emulsions with a concentration of the dispersed phase in diameters. Second, the stage of the emulsification process that
the range of 80–90 (wt%) prepared using three different demands the most energy is the homogenization step (Alvarez
impellers: 45◦ Pitch blade, Hydrofoil and Propeller. (The et al., 2010).
concentration values in ascending order are 80%, 82%, 85%, A very important factor that influences this stage is the
87% and 90%.) ability that each impeller has to move the fluid across the

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
6 chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx

the slope of the straight line that best fits the three curves
Table 4 – The calculated pumping capacity for the three
impellers used in this study (pitch blade, hydrofoil and (three impellers) has a value of 4 (R2 = 0.99). Similarly, when the
propeller) shown in Fig. 1. concentration goes up to 85–90 (wt%), the slope of the straight
line that best fits the three impellers has a value of 3 (R2 = 0.99).
Type of impeller Pumping capacity (m3 /min)
Eq. (4) These two trends can be described through the proportional
relationships in Eqs. (5) and (6):
“PB”: Pitch Blade 45◦ 0.34
“HF”: Hydrofoil 0.17
G ∝ E4v (5)
“PR”: Propeller 0.10

G ∝ E3v (6)
mixing tank. This ability can be measured through the so-
called pumping capacity, which can be calculated using Eq. It is that the slope of the straight line decreases as the con-
(4) from reference Hemrajni and Tatterson (2004): centration of the dispersed phase increases. When this occurs,
the rheological properties associated with the incorporated
Q = Nq ND3 (4) energy decreases. Therefore the energy might split into three
paths once it has been incorporated during the emulsification
Q represents the circulation of the fluid within the mixing process: first, energy is required for droplet formation; second,
tank, Nq is the dimensionless pumping number, N is the energy is required to circulate the fluid across the mixing tank;
mixing velocity in rpm and D is the impeller diameter. The and third, when the droplets are small enough to come into
dimensionless pumping number can be calculated or found contact with one another, the interactions between them is
in different handbooks; for instance, the values used for the more relevant, therefore they dissipate energy due to droplet-
“PB” and “HF” impellers in this study were found in (Bakker droplet contact (high packing and change in morphology).
and Gates, 1995) and the value for the “PR” impeller was found We can now look for proportional relationships that are
in (Comments by Murray, 1948) and (Procházka and Landau, similar to those established in Eqs. (5) and (6) for the aver-
1961). It is important to highlight that these values are valid age droplet size and the energy (dotted lines in Fig. 2). For
for laminar flow. the concentration range of 80–85 (wt%), the straight line that
Table 4 shows the calculated pumping capacity for the sys- best fits the three impellers has a slope of −2.5 (R2 = 0.99), and
tem described in Figs. 1–3. It is evident that the values are similarly, the slope of the straight line that best fits the three
different for the three impellers used; therefore a possible impellers for concentrations in the range of 85–90 (wt%) has
explanation for why the curves in Fig. 1 do not overlap is a value of −1.5 (R2 = 0.99). Eqs. (7) and (8) describe these rela-
that each impeller uses the energy incorporated through the tionships and once again, the trend of the slope is evident:
mixing system differently. A portion of the energy is used for as the concentration increases, the slope of the line tends to
droplet formation as expected, whereas the other portion of decrease.
the energy is used for fluid circulation.
From Fig. 1 and Table 3, it can be concluded that the D ∝ E−2.5
v (7)
impeller with greatest pumping capacity is the one that
exhibits the lowest energy consumption for the studied emul- D ∝ E−1.5
v (8)
sions. Similarly, there is a qualitative relationship between
the energy consumed and the pumping capacity: a low These results provide a clearer insight into the relationship
pumping capacity partially accounts for the increase in energy between the process, the product and properties of con-
consumption when emulsions with similar properties are centrated and highly concentrated emulsions and how the
obtained. An impeller that has a high pumping capacity value energy is related to the different formulation and composition
will allow for a better circulation of the fluid within the mixing variables. Eqs. (7) and (8) show that for systems that are mod-
tank for concentrated and highly concentrated emulsions. erately concentrated (80 wt%), where the elasticity is low and
Different authors (Salager, 1996; Tadros, 2004; Alvarez et al., the droplets are somehow larger, the relationship between the
2010; Capdevila et al., 2010) have implicated the tip velocity elastic modulus (or the average droplet size) and the energy
as a key factor in the emulsification process, mainly because is strongly correlated, but when the concentration increases
this factor can control the amount of shear available for to 90 wt%, this relationship becomes less relevant, indicating
droplet formation. If we now take into consideration the that highly concentrated systems require much more energy
results presented in Figs. 1–3, a novel methodology can be used to be circulated across the mixing tank.
for the preparation of concentrated and highly concentrated The last relationship that can be established using the
emulsions. This methodology (the multi-scale approach) may multi-scale approach is that between the incorporated energy
be used to manufacture emulsions with certain desirable and the properties at the molecular level represented by the
characteristics while minimizing the energy consumption by near infrared spectra of the emulsions. Fig. 3 shows the behav-
controlling the tip velocity and selecting an impeller that ior of the absorbance of the first overtone of the C H bond
yields the highest pumping capacity. These parameters may interactions as a function of the incorporated energy. The con-
be used as criteria for scale-up processes. centration range is the same as those reported in Figs. 1 and 2.
If we now look more closely at the relationships between When the concentration increases, the absorbance
the incorporated energy and the rheological behavior (that is, increases. This result was expected and has been studied by
the microscopic structure and the molecular response), we can Sjöblom et al. (2003): When there is an increase in the amount
follow an approach similar to the one proposed by Alvarez of any substance, the characteristic peaks that best describe
(Alvarez et al., 2010) and establish equations similar to Eqs. it will increase because of the increase in the number of
(1) and (2). In Fig. 1, two trends proposed are represented by bonds that vibrate. When Fig. 3 is evaluated with Figs. 1 and 2,
the dotted lines: when the concentration range is 80–85 (wt%), it can be noted that the increase in the absorbance is also

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx 7

will cause the beam of light to change its trajectory at some


angle. These two phenomena may occur simultaneously, or
one may dominate over the other, and this influences the
total absorbance measured by the device. There are several
other phenomena that make smaller contributions to the total
measured absorbance, such as diffraction, partial absorption,
back scattering and multiple scattering, but they will not be
included in the following analysis (Bohren and Huffman, 1983;
Aske et al., 2002; van de Hulst, 2012).
The predominance of one phenomenon over the other
depends, primarily, on the size and shape of the particle that
interacts with the beam of light. Araujo et al. (2008) reported
a trend in the behavior of the near infrared spectra of water-
in-crude oil emulsions with different average droplet sizes.
They established that the total measured absorbance of the
spectra of emulsions with smaller average droplet sizes was
Fig. 4 – Near infrared spectra of oil-in-water emulsions
higher than the total measured absorbance of the spectra of
with a concentration of the dispersed phase of 30 (wt%).
emulsions with larger average droplet sizes. The behavior of
Average droplet size also shown (D[4,3] ␮m).
spectra 3, 4 and 5 in Fig. 4 is similar to the reported trend.
This trend is attributed to the predominance of the light
related to a decrease in the average droplet size; therefore, absorption phenomenon by the oil particles in this droplet
this technique is also able to quantify changes in elasticity size range. If the average droplet size is within or close to the
and the morphology of droplets and can also be used for this value of the wavelength (Rayleigh limit), or in other words,
purpose. when the volume of the particle approaches 0, the extinction
If we now look for relationships between the absorbance of of light due to absorption is proportional to V/ whereas the
the first overtone (the oil droplets) and the incorporated energy extinction of light due to scattering is proportional to 4 V2 (van
(dotted lines in Fig. 3), Eqs. (9) and (10) can be established: de Hulst, 2012) where V is the volume of the particle and 
is the wavelength. For spectra 3, 4 and 5 in Fig. 4, in which
AbsC H ∝ E−0.17
v (9) the average droplet size is approximately 2–7 ␮m, the domi-
nant phenomenon is light absorption, which thus contributes
AbsC H ∝ E−0.10
v (10) more to the total measured absorbance. This is because the
size range of the droplets is very similar to the wavelength of
These equations also have a good fit for the slopes of the the incoming light (1–2.5 ␮m).
straight lines (R2 = 0.99), and once again, it can be seen that Analogously, spectra 1 and 2 in Fig. 4 correspond to emul-
the slope decreases as the concentration increases. sions having an average droplet size of at least one order
These results provide better insight about the behavior of of magnitude higher than the incident wavelength (diameter
these types of colloidal systems and how the process, the 10 ␮m). Within this range, the predominant phenomenon
product and the properties are related. As stated recently is light scattering. As pointed out by van de Hulst (2012), the
by Capdevila et al. (2010), for concentrated and highly con- incident beam of light, which forms a plane wave front with
centrated systems such as O/W and W/O emulsions, the infinite extent, may be thought to consist of separate rays of
formulation and composition variables are of paramount light that pursue their own path. This allows the rays actually
importance in product design and product scale-up because hitting the particle and rays passing along the particle to be
of the highly non-Newtonian nature of these fluids. differentiated and therefore these rays are said to be localized.
When this occurs, the probability of multiple beams of light
3.2. Part II: near infrared spectroscopy and droplet size hitting one particle is much greater for large particles than for
small particles because, in the latter case, neither absorption
The second part of the discussion focuses on describing and nor scattering or diffraction are likely to occur. This causes an
understanding the possible relationship between the changes increase in the total measured absorbance simply because the
in the average droplet size of the O/W emulsions and the scattered light leaves the particle at angles that make it dif-
molecular response represented by the near infrared spec- ficult to reach the detectors. Because the scattering particles
trum. To accomplish this, O/W emulsions with the same are homogeneous, absorbing and spherical, there is no dou-
concentration of the dispersed phase were prepared by vary- ble fraction or dichroism of any kind (van de Hulst, 2012). The
ing the tip velocity (different shear) to generate different above theoretical implications hold very well for particles and
droplet size distributions. mediums with refractive index close to 1 as in this study.
The near infrared spectra of O/W emulsions with a 30 (wt%) Of note in the spectra shown in Fig. 4, is the behavior when
concentration of the dispersed phase is presented in Fig. 4 average droplet size changes in either region. In the first region
along with the values for the average droplet size, D[4,3]. The (spectra 3–5 in Fig. 4), where the average droplet size can be
only difference between the emulsions is the average droplet compared to the wavelength of the incident light and light
size meaning that the formulation is constant. absorption dominates, if the average droplet size decreases,
When the incident light beam meets particles in its way, the total measured absorbance increases; this trend is similar
two main phenomena may occur: first, the particles can absorb to that reported by Araujo et al. (2008). In the second region
the light, and the molecular bonds within the particles will (spectra 1 and 2 in Fig. 4), where the average droplet size is
vibrate at a frequency similar to that of the incoming light. much higher than the wavelength and light scattering dom-
Second, the particles can scatter the light, and the particles inates, if the average droplet size decreases then the total

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
8 chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx

Fig. 5 – Near infrared spectra of oil-in-water emulsions


with a concentration of the dispersed phase of 50 (wt%).
Average droplet size also shown (D[4,3] ␮m). Fig. 6 – Near infrared spectra of oil-in-water emulsions
with a concentration of the dispersed phase of 80 (wt%).
measured absorbance also decreases; this trend is similar to Average droplet size also shown (D[4,3] ␮m).
that reported by Sjöblom et al. (2003) and Kallevik et al. (2000).
With these results it is possible to observe a change in the solving the field equations is relatively straightforward. If the
behavior of the spectra depending on whether light absorption boundary conditions are not defined, the constitutive models
or light scattering is the dominant phenomenon. This change become highly problematic to solve. There are methods for
in the behavior is due to the changes in the average droplet solving the equations using approximations; however, the
size. results deviate from the results for a spherical particle simply
Even more interesting than the change in the behavior of because of the changes in the methods for solving the equa-
the spectra of O/W emulsions due to the changes in the aver- tions and because the spherical problem is a limiting case
age droplet size is how these two trends are influenced by the (Bohren and Huffman, 1983; van de Hulst, 2012). When dealing
concentration of the dispersed phase. Fig. 5 shows the spectra with non-spherical particles (or a mixture of both spherical
of O/W emulsions of a given concentration of the dispersed and non-spherical particles), the equations become specific,
phase of 50 (wt%) with different average droplet sizes. In this where each case must be treated differently (which yields
case, it is possible to see the same trends described above, different constitutive models) and some approximations
where spectra 6 and 7 in Fig. 5 are comparable to spectra 1 and must be considered. It is highly unlikely that a constitutive
2 in Fig. 4. In this case, when the average droplet size decreases, model for each particle within the medium (i.e., emulsion)
the total measured absorbance also decreases. Similarly, spec- can be produced and solved.
tra 8 and 9 in Fig. 5 are comparable to spectra 3–5 in Fig. 4, and It is important to note that near infrared spectroscopy
show that when the average droplet size decreases, the total can be used to establish two different trends depending on
measured absorbance increases. the predominance of one phenomenon over the other (light
Finally, when the concentration of the dispersed phase absorption or light scattering); this behavior depends on the
increases, the average droplet size decreases. This causes the changes in the average droplet size. The relevance of this find-
droplets to be closer to one another to the point that some or ing is that this technique can be used together with statistical
all of them are no longer spheres and become polyhedral (Pal, calibration methods, such as partial least squares, to adapt
1996). This is true for emulsions with an 80 (wt%) concentra- it as a quick on-line quality measurement for industries in
tion of the dispersed phase; these spectra are shown in Fig. 6. which it is necessary to know in advance the droplet size or
The average droplet size is also shown. It is clear from Fig. 6 the average droplet size of the product (e.g., cosmetics or bev-
that the two trends described for diluted and moderately con- erage emulsions). It has been shown here that the two trends
centrated emulsions are still identifiable, and the behavior of described for the near infrared spectra are independent of the
spectra 11 and 12 in Fig. 6 are comparable to spectra 6 and 7 in concentration of the dispersed phase; this result can also be
Fig. 5 or spectra 1 and 2 in Fig. 4. Similarly, spectra 10 and 13 used for industries that deal with highly concentrated systems
in Fig. 6 are comparable to spectra 8 and 9 in Fig. 5 or spectra (e.g., petrochemical or food emulsions).
3–5 in Fig. 4.
However, the distinction between the two regions in Fig. 6 3.3. Part III: creep tests and viscoelasticity
is not as evident as in Fig. 5 or Fig. 4. This is due to the
change in the morphology of the oil droplets, and this change For the last part of this work, we will again focus on the
requires a completely different theoretical analysis of how rheological behavior of concentrated and highly concentrated
the light behaves once it meets a particle with non-spherical emulsions. We will show one method of relating the changes
shape (van de Hulst, 2012). This is mainly because knowl- that occur at the microscopic level (i.e., changes in the struc-
edge of the specific shape or morphology of the particle, in ture/morphology of the droplets) and the response at the
addition to the homogeneous nature of its optical constants, macroscopic level (i.e., rheological properties) using creep test-
allows for the construction of a mathematical problem (field ing. The results will be also related to the viscoelastic nature
equations and constitutive models) that involve defined of the emulsions.
known-boundary conditions. Because a spherical particle Creep tests were used to study water-in-oil emulsions
of one material is thought to be isotropic, the method for where the concentration of the dispersed phase was between

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx 9

Fig. 7 – Compliance modulus J(t) as a function of time Fig. 8 – Compliance modulus J(t) as a function of time
obtained through creep tests (imposed stress 50 Pa) for obtained through creep tests for water-in-oil emulsions at
water-in-oil emulsions having a concentration of the different values of the imposed stress () 1–125 [Pa]. The
dispersed phase from 70 to 85 (wt%). The value of the concentration of the dispersed phase is fixed at 85 (wt%).
imposed stress is within the linear viscoelastic region
(0.1–100 [rad/s] and 0.1–100 [Pa]).

70 and 85 (wt%). Fig. 7 shows the compliance modulus J(t) as a


function of time, and it is clear from this figure that the dura-
tion of the experiments was sufficient to reach the terminal
regime. When this regime is attained, it is possible to identify
the three typical regions of a viscoelastic material proposed by
Tadros (2004): the elastic response (at low times), the retarded
viscoelastic response (at intermediate times) and the steady
state viscous response (at latter times).
When the concentration of the dispersed phase is higher
than 70 (wt%), it is possible to observe the so-called “ringing”
phenomenon (perfectly reproducible damped oscillations)
located at intermediate times, more specifically, at the
retarded viscoelastic response (Fig. 7). This phenomenon has
been reported in some publications (Zölzer and Eicke, 1993;
Baravian and Quemada, 1998; Ewoldt and McKinley, 2007), and
it is a result of the coupled effect between the elasticity of the
sample and the inertia of the instrument. When the concen-
tration of the dispersed phase is 70 (wt%), which is below the
maximum packing fraction for spherical droplets, the oscil-
lations are not present and the behavior of the compliance
modulus is smooth as expected. However, when the concen-
tration increases, the phenomenon is evident. This suggests
that ringing is a consequence of three variables. The first vari-
able is the inertia of the instrument, which is unavoidable in
controlled stress rheometers. The second variable is the elas-
ticity of the sample, which is a result of the increase in the
concentration of the dispersed phase.
Highly concentrated water-in-oil emulsions are systems in
which the elastic response clearly dominates the viscoelastic
nature of the emulsion. Fig. 9 shows that the elastic modulus
(G ) is at least two orders of magnitude higher than the loss
Fig. 9 – (a) Elastic and loss moduli (G and G ) obtained
modulus (G ). The third variable is the changes in morphology
through a frequency sweep for a water-in-oil emulsion with
(droplet deformation) as the concentration increases which
a concentration of the dispersed phase of 85 (wt%) showing
cause an increase in the interaction forces as the droplets
the linear viscoelastic region. The frequency sweep was
come closer to one another when the droplets reach the
performed at a fixed stress of 1 Pa (b). Elastic and loss
maximum packing fraction. In this sense, ringing can be
moduli (G and G ) obtained through a stress sweep for a
thought of as a manifestation of the viscoelastic nature
water-in-oil emulsion with a concentration of the dispersed
of concentrated and highly concentrated W/O emulsions.
phase of 85 (wt%) showing the linear viscoelastic region.
It is important to highlight that the behavior depicted in
The stress sweep was performed at a fixed frequency of
Figs. 8 and 9 for an emulsion with a concentration of the
1 rad/s. For both figures 20 points per decade per duplicate
dispersed phase of 85 (wt%) is similar (data not shown) if the
were taken.
concentration is lower, for instance: 70 (wt%). This means

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
10 chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx

that all the emulsions in this study are analyzed within the the same sense, to have a preliminary notion of the proper-
linear viscoelastic region where the elastic modulus (G ) is ties (elastic modulus, viscosity, average droplet size) of the
independent of the stress (and in this particular case, also emulsion after preparation.
independent of the frequency). Furthermore, it was established how the energy incor-
If the analysis was to be taken further and if creep data porated through the emulsification process depends on two
were to be transformed into oscillatory data (G and G ), it variables: first, the tip velocity by which droplet formation is
would be necessary to separate the response associated to the set through shear and second, the pumping capacity of the
inertia of the instrument. This goes beyond the scope of this impeller which determines the circulation of the emulsion
work. However the calculations and the appropriate rheologi- on the mixing tank. This novel result leads to a new way of
cal models have been previously reported elsewhere; see for considering emulsion manufacture. Instead of solely consid-
instance (Zölzer and Eicke, 1993; Baravian and Quemada, 1998; ering the design of an impeller for a given application, it is
Ewoldt and McKinley, 2007). proposed to choose an impeller with the suitable pumping
Fig. 8 shows an interesting result that supports the fact capacity that allows a fixed formulation to produce an emul-
that ringing is a manifestation of the viscoelastic nature of sion with desired properties. For instance, if the droplet size
W/O emulsions, where the compliance modulus is plotted as distribution is a key parameter of a given product, Figs. 1–3 can
a function of time with various values for the imposed stress be used to select the appropriate impeller based on the energy
and a fixed concentration of 85 (wt%). When the imposed required.
stress in the creep experiment is low (i.e., 1 Pa) the oscilla- Through the use of a multi-scale approach, it was also
tions are diffused and not reproducible; similarly, when the possible to extract relevant information about the different
imposed stress is high (i.e., 125 Pa) ringing not present. In this properties of O/W and W/O emulsions. First, although near
figure, it should be noted that ringing becomes clear and repro- infrared spectroscopy (NIRS) can be used to track changes in
ducible in a stress range that corresponds to the stress range the amount of a given substance, for instance oil (Aske et al.,
of the linear viscoelastic region, where the elastic response 2002; Sjöblom et al., 2003; Araujo et al., 2008) it was possible to
of the emulsion (i.e., the elastic modulus) is independent elucidate the influence of the droplet size in the spectra of the
of the stress (and in our case, also independent of the fre- emulsions. Figs. 4–6 show that the predominance of two phe-
quency). It is important to emphasize that in our case, the nomena depending on how close the average droplet size is to
stress range applied during the creep tests must be within the the incident wavelength causes different trends in the spectra.
linear viscoelastic region otherwise the response of the emul- These two phenomena are light absorption and light scatter-
sions would not primarily be elastic. This was corroborated ing. Second, the phenomenon commonly known as “ringing”
through oscillatory tests to determine the linear viscoelas- (Zölzer and Eicke, 1993; Ewoldt and McKinley, 2007; Evans et al.,
tic region presented in Fig. 9a, which shows a frequency and 2009) was observed for W/O emulsions. This is attributed not
Fig. 9b, which shows a stress sweep for W/O emulsions with a only to the decrease in droplet size but also to the phys-
concentration of the dispersed phase of 85 (wt%). ical deformation (packing) that droplets undergo when the
concentration of the dispersed phase increases. Furthermore,
ringing is only present at a specific range of applied stress.
4. Conclusions
References
Research on concentrated and highly concentrated emul-
sions has been traditionally focused on relating formulation
Alvarez, O.A., Choplin, L., et al., 2010. Influence of semibatch
variables (i.e., amount of surfactant, concentration of the emulsification process conditions on the physical
dispersed phase) to limited aspects of their properties (Jager- characteristics of highly concentrated water-in-oil emulsions.
Lézer et al., 1998; Derkach, 2009; Capdevila et al., 2010; Ind. Eng. Chem. Res. 49 (13), 6042–6046.
Masalova et al., 2011). The main results of this type of approach Araujo, A.M., Santos, L.M., et al., 2008. Evaluation of water
have led to the development of some scale-up parameters content and average droplet size in water-in-crude oil
based on traditional dimensionless and characteristic num- emulsions by means of near-infrared spectroscopy. Energy
Fuels 22 (5), 3450–3458.
bers. However, the main flaw of this analysis is that it fails
Aske, N., Kallevik, H., et al., 2002. Water-in-crude oil emulsion
to establish adequate relationships for non-Newtonian fluids. stability studied by critical electric field measurements.
In this study, a first attempt to tackle this issue through the Correlation to physico-chemical parameters and
use of a multi-scale approach is made (Alvarez et al., 2010). near-infrared spectroscopy. J. Petrol. Sci. Eng. 36 (1–2), 1–17.
Relationships between the energy incorporated through the Babak, V., Langenfeld, A., et al., 2001. Rheological properties of
emulsification process and the properties at three different highly concentrated fluorinated water-in-oil emulsions. In:
Koutsoukos, P. (Ed.), Trends in Colloid and Interface Science
levels were established (Eqs. (5)–(10)): first, the macroscopic
XV, 118. Springer, Berlin Heidelberg, pp. 216–220.
level represented by the elastic modulus G , second, the micro-
Bakker, A., Gates, L.E., 1995. Properly choose mechanical agitators
scopic level represented by the average droplet size D[4,3] and for viscous liquids. Chem. Eng. Prog. 91 (12), 25–34.
third, the molecular level represented by the near infrared Baravian, C., Quemada, D., 1998. Using instrumental inertia in
spectrum. controlled stress rheometry. Rheol. Acta 37 (3), 223–233.
In this sense, an intertwined connection between the Bohren, C.F., Huffman, D.R., 1983. Absorption and Scattering of
energy incorporated, the formulation of the given system and Light by Small Particles. Wiley.
Bouyer, E., Mekhloufi, G., et al., 2011. Stabilization mechanism of
the properties is proposed as a way of producing emulsions
oil-in-water emulsions by ␤-lactoglobulin and gum arabic. J.
with given characteristics. The description of how the energy Colloid Interface Sci. 354 (2), 467–477.
incorporated influences the properties depending on the dif- Capdevila, M., Maestro, A., et al., 2010. Preparation of Span
ferent formulation gives a novel insight on product design. 80/oil/water highly concentrated emulsions: influence of
This behavior, represented in Figs. 1–3, can be used to pro- composition and formation variables and scale-up. J. Colloid
duce tailor-made emulsions given a fixed formulation; or, in Interface Sci. 345 (1), 27–33.

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016
CHERD-1723; No. of Pages 11
ARTICLE IN PRESS
chemical engineering research and design x x x ( 2 0 1 4 ) xxx–xxx 11

Charton, S., Kacem, A., et al., 2013. Actinides oxalate constituents of finished gasolines. Anal. Chem. 62 (14),
precipitation in emulsion modeling: from the drop scale to 1444–1451.
the industrial process. Chem. Eng. Res. Des. 91 (4), 660–669. Langevin, D., 2000. Influence of interfacial rheology on foam and
Clausse, D., Daniel-David, D., Gomez, F., Komunjer, L., Pezron, I., emulsion properties. Adv. Colloid Interface Sci. 88 (1–2),
Dalmazzone, C., Noïk, C., 2007. Emulsion stability and 209–222.
interfacial properties – application to complex emulsions of Malkin, A.Y., Masalova, I., et al., 2004. Effect of droplet size on the
industrial interest. In: Tadros, T.F. (Ed.), Colloid Stability and rheological properties of highly-concentrated w/o emulsions.
Application in Pharmacy, vol. 3. Wiley-VCH Verlag GmbH & Rheol. Acta 43 (6), 584–591.
Co. KGaA, Weinheim, Germany. Masalova, I., Foudazi, R., et al., 2011. The rheology of highly
Clausse, D., Pezron, I., et al., 1999. Stability of W/O and W/O/W concentrated emulsions stabilized with different surfactants.
emulsions as a result of partial solidification. Colloids Surf. A: Colloids Surf. A: Physicochem. Eng. Aspects 375 (1–3),
Physicochem. Eng. Aspects 152 (1–2), 23–29. 76–86.
Comments by Murray, A.B.,1948. The choice of suitable Reynolds Masalova, I., Malkin, A.Y., 2008. Master curves for elastic and
number for model propeller experiments. In: Fifth plastic properties of highly concentrated emulsions. Colloid J.
International Conference of Ship Tank Superintendents 70 (3), 327–336.
London. Captain Harold Saunders, Chairman. Mason, T.G., Lacasse, M.D., et al., 1997. Osmotic pressure and
Datta, S.S., Gerrard, D.D., et al., 2011. Rheology of attractive viscoelastic shear moduli of concentrated emulsions. Phys.
emulsions. Phys. Rev. E – Stat. Nonlin. Soft Matter Phys. 84 (4). Rev. E – Stat. Phys. Plasmas Fluids Relat. Interdiscipl. Top. 56 (3
Derkach, S.R., 2009. Rheology of emulsions. Adv. Colloid Interface B), 3150–3166.
Sci. 151 (1–2), 1–23. Mougel, J., Alvarez, O., et al., 2006. Aging of an unstable w/o gel
Dimitrova, T.D., Leal-Calderon, F., 1999. Forces between emulsion emulsion with a nonionic surfactant. Rheol. Acta 45 (5),
droplets stabilized with Tween 20 and proteins. Langmuir 15 555–560.
(26), 8813–8821. Pal, R., 1996. Effect of droplet size on the rheology of emulsions.
Dimitrova, T.D., Leal-Calderon, F., 2001. Bulk elasticity of AIChE J. 42 (11), 3181–3190.
concentrated protein-stabilized emulsions. Langmuir 17 (11), Pal, R., 1998a. A novel method to correlate emulsion viscosity
3235–3244. data. Colloids Surf. A: Physicochem. Eng. Aspects 137 (1–3),
Dimitrova, T.D., Leal-Calderon, F., 2004. Rheological properties of 275–286.
highly concentrated protein-stabilized emulsions. Adv. Pal, R., 1998b. Scaling of viscoelastic properties of emulsions.
Colloid Interface Sci. 108–109, 49–61. Chem. Eng. J. 70 (2), 173–178.
Evans, R.M.L., Tassieri, M., et al., 2009. Direct conversion of Pal, R., 1999. Yield stress and viscoelastic properties of high
rheological compliance measurements into storage and loss internal phase ratio emulsions. Colloid. Polym. Sci. 277 (6),
moduli. Phys. Rev. E – Stat. Nonlin. Soft Matter Phys. 80 (1.). 583–588.
Ewoldt, R.H., McKinley, G.H., 2007. Creep ringing in rheometry or Pal, R., 2006. Rheology of high internal phase ratio emulsions.
how to deal with oft-discarded data in step stress tests. Rheol. Food Hydrocolloids 20 (7), 997–1005.
Bull. 76 (1.). Procházka, J., Landau, J., 1961. Studies on mixing. XII.
Fa, N., Babak, V.G., et al., 2004. The release of caffeine from Homogenation of miscible liquids in the turbulent region.
hydrogenated and fluorinated gel emulsions and cubic Collect. Czech. Chem. Commun. 26 (12),
phases. Colloids Surf. A: Physicochem. Eng. Aspects 243 (1–3), 2961–2973.
117–125. Sagis, L.M.C., 2011. Dynamic properties of interfaces in soft
Frake, P., Luscombe, C.N., et al., 1998. Mass median particle size matter: experiments and theory. Rev. Mod. Phys. 83 (4),
determination of an active compound in a binary mixture 1367–1403.
using near-infrared spectroscopy. Anal. Commun. 35 (4), Salager, J.-L., 1996. Guidelines to handle the formulation,
133–134. composition and stirring to attain emulsion properties on
Gossen, P.D., MacGregor, J.F., et al., 1993. Composition and particle design (type, drop size, viscosity and stability). Surfact. Solut.
diameter for styrene/methyl methacrylate copolymer latex Surfact. Sci. Ser. 64, 261–295.
using UV and NIR spectroscopy. Appl. Spectrosc. 47 (11), Sjoblom, J., 2001. Encyclopedic Handbook of Emulsion
1852–1870. Technology. Taylor & Francis.
Hemrajni, R., Tatterson, G., 2004. Chapter 6: mechanically stirred Sjöblom, J., Aske, N., et al., 2003. Our current understanding of
vessels. handbook of industrial mixing. In: Science and water-in-crude oil emulsions. Recent characterization
Practice. John Wiley & Sons, New Jersey, pp. 345–390. techniques and high pressure performance. Adv. Colloid
Jager-Lézer, N., Tranchant, J.F., et al., 1998. Rheological analysis of Interface Sci. 100–102 (Suppl.), 399–473.
highly concentrated w/o emulsions. Rheol. Acta 37 (2), Tadros, T., 2004. Application of rheology for assessment and
129–138. prediction of the long-term physical stability of emulsions.
Kallevik, H., Hansen, S.B., et al., 2000. Crude oil model emulsion Adv. Colloid Interface Sci. 108–109, 227–258.
characterised by means of near infrared spectroscopy and van de Hulst, H.C., 2012. Light Scattering by Small Particles. Dover
multivariate techniques. J. Dispers. Sci. Technol. 21 (3), Publications.
245–262. Zölzer, U., Eicke, H.F., 1993. Free oscillatory shear measurements –
Kelly, J.J., Callis, J.B., 1990. Nondestructive analytical procedure for an interesting application of constant stress rheometers in
simultaneous estimation of the major classes of hydrocarbon the creep mode. Rheol. Acta 32 (1), 104–107.

Please cite this article in press as: Pradilla, D., et al., The application of a multi-scale approach to the manufacture of concentrated and highly
concentrated emulsions. Chem. Eng. Res. Des. (2014), http://dx.doi.org/10.1016/j.cherd.2014.10.016

You might also like