You are on page 1of 16

chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Droplet breakage and coalescence models for the


flow of water-in-oil emulsions through a valve-like
element

João F. Mitre a,1 , Paulo L.C. Lage a,∗ , Marcos A. Souza b , Eli Silva b ,
Luiz Fernando Barca b , Antonio O.S. Moraes a , Raquel C.C. Coutinho c ,
Elizabeth F. Fonseca c
a Programa de Engenharia Química – COPPE, Universidade Federal do Rio de Janeiro, PO Box 68502, Rio de Janeiro,
RJ 21941-972, Brazil2
b Núcleo de Separadores Compactos (NUSEC), Instituto de Engenharia Mecânica, Universidade Federal de Itajubá,

Av. BPS 1303, Itajubá, MG 37500-903, Brazil


c CENPES/PETROBRAS, Rio de Janeiro, RJ 21949-900, Brazil

a b s t r a c t

An experimental and theoretical study was carried out regarding the evolution of droplet size distributions for the
flow of water in oil emulsions through a valve-like element that simulates a mixing valve. The water droplets had
diameters in the 0.1–100 ␮m range, which includes droplets that are either smaller or larger than the Kolmogorov
length scale for the experimental conditions. Droplet breakage and coalescence models that can be used for this
size range were proposed. A simplified population balance model was developed to interpret the experimental data
and solved by the method of classes. The model parameters were estimated by the orthogonal distance regression
method. The simplifying assumption of the model was verified by global optimization using a parallel implementa-
tion of the particle swarm optimization method. The agreement between simulated and experimental droplet volume
size distributions was good. The predictions of the DeBroukere mean diameter, d43 , were unbiased with mean errors
of 8%.
© 2014 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords: Mathematical modeling; Parameter estimation; Population balance; Emulsion; Coalescence; Breakage

1. Introduction key variables in defining the separation efficiency (Raghavan


et al., 1971; Paiva, 2004; Eow and Ghadiri, 2002; Cunha, 2008).
During oil production, the mixture of water and oil is sub- The oil is received at the refinery with a small fraction of
mitted to large shear rates generating water–oil emulsions water but with a large concentration of salts. Before the desalt-
(Raghavan et al., 1971; Paiva, 2004; Eow and Ghadiri, 2002; ing equipment, fresh water is added to wash the oil, increasing
Cunha, 2008) which have a large salt content. Electrostatic the concentration of water droplets. A globe valve is com-
demulsification of water-in-oil emulsions is largely applied in monly used to promote adequate mixing by generating droplet
the oil industry to desalt the crude dead oil prior to its refin- breakage and coalescence.
ing (Manning and Thompson, 1995; Stewart and Arnold, 2008, The emulsion flow is a polydisperse multiphase flow
2009; Noïk et al., 2006). Droplet sizes and salt concentration are that can be simulated by couplinga population balance


Corresponding author. Tel.: +55 21 2562 8346; fax: +55 21 2562 8300.
E-mail address: paulo@peq.coppe.ufrj.br (P.L.C. Lage).
Available online 4 April 2014
1
Present address: Departamento de Engenharia Química e de Petróleo, Escola de Engenharia, Universidade Federal Fluminense, Campus
Praia Vermelha, Rua Passos da Pátria, 156, Bloco D, Sala 305, Niterói, RJ 24210-240, Brazil.
2
http://www.peq.coppe.ufrj.br/areas/tfd/.
http://dx.doi.org/10.1016/j.cherd.2014.03.020
0263-8762/© 2014 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
2494 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

Nomenclature
xk pivot of class k
a(v, v ) aggregation frequency of the particles with vol- y vector of continuous phase variables
ume v and v z spatial position
A Hamaker constant Ż particle velocity vector
Ac flow cross section area
b(v) breakage frequency of the particle with volume Greek symbols
v ˛ vector of model parameters
BB production of particles by breakage ˇ dimensionless constant equal to 8.2
BC production of particles by coalescence ı error in explanatory variable
c minimum clearance in the test section  variation
Cb parameter of the breakage frequency model  dissipation rate of turbulence energy
Cc parameter of the coalescence frequency model ε error in response variable
Ce parameter of the coalescence efficiency model  Kolmogorov length scale
Ca capillary number  collision frequency
d particle diameter  coalescence efficiency
dc,mb critical diameter of Martínez-Bazán et al.  dynamic viscosity
(1999a) kinematic viscosity
deq equivalent diameter in the collision of drops i
explanatory variable
and j, d−1 −1 −1
eq = 0.5(di + dj )
density
dpq mean diameter given by interfacial tension
 ∞ ∞ 1/(p−q) ς mean number of daughter particles
0
dp f (v)dv/ 0 dq f (v)dv
 water volume fraction
DB destruction of particles by breakage
˚ multiresponse model
DC destruction of particles by coalescence
 response variable
Dpq mean diameter given by
 ∞ ∞ 1/(p−q)
0
dp F(v)dv/ 0 dq F(v)dv Subscripts
e relative error crit critical value
e mean of the relative errors d dispersed phase
|e| mean of the magnitude of the relative errors exp experimental
emax maximum error in magnitude in at the inlet of the test section
f particle number distribution function max maximum
F particle volume distribution function n refers to normalized distribution
G objective function out at the outlet of the test section
h film thickness sim simulated
H particle net production rate v refers to volume distribution
L turbulent length scale
l depth of the test section (l = 5 mm)
M number of valid experimental runs
N zeroth-order sectional moment model (Ramkrishna, 2000) with computational fluid dynam-
NT particle number density ics (Marchisio et al., 2003; Marchisio and Fox, 2005; Silva et al.,
n number of classes 2008; Silva and Lage, 2011). The accuracy of this approach
P(v |v) daughter particle breakage conditional proba- heavily relies on the adequacy of the employed breakage and
bility coalescence models, which are not generally applicable and
p pressure have adjusted parameters that usually depend on the charac-
Q emulsion volumetric flow teristics of the multiphase system (Lasheras et al., 2002; Liao
Re Reynolds number and Lucas, 2009, 2010; Mitre et al., 2010).
ReT Reynolds number based on the Taylor length Coalescence is believed to occur in three stages (Marrucci,
scale 1969; Friedlander, 1977; Coulaloglou and Tavlarides, 1977; Ross
S collision cross-section of particles with diame- et al., 1978; Kim and Lee, 1987; Prince and Blanch, 1990;
ters d and d Chesters, 1991; Kamp et al., 2001; Liao and Lucas, 2010).
StK Stokes number based on Kolmogorov time scale First, there is a collision between two fluid particles, which
t time become separated by a thin fluid film. Then, the fluid in this
tres particle residence time film is drained until it reaches a critical thickness at which
u velocity fluctuation molecular attractive forces dominate, promoting the film rup-
ud particle characteristic velocity ture and leading to coalescence. In turbulent flows, particle
U mean velocity collisions are mainly controlled by particle–turbulence inter-
v particle volume actions, which also cause intense particle deformation that
Ve effective volume of the valve-like element leads to fluid particle breakage (Hinze, 1955; Lasheras et al.,
W mass flow rate 2002).
W matrix of weights of the errors in the objective The breakage and coalescence modeling in turbulent flows
function depends on the size ratio between the fluid particle and the
We Weber number Kolmogorov length scale, . If the particle is smaller than , it is
in the sub-Kolmogorov size range. If the particle size is smaller
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2495

than the dissipation-inertial size limit (∼20), it is in the tur- leading to a different breakage dependence on the Weber
bulence dissipation range (Pope, 2000). On the other hand, number. Recently, Maniero et al. (2012) modeled and simulated
if the particle is larger than this limit and smaller than the the deformation of single droplets in the flow downstream of
integral scale, it is in the inertial subrange of the turbulence a duct orifice. They simulated the continuous-phase flow by
spectrum. DNS and used a Lagrangian drop tracking model that included
There are several models for breakage and coalescence of the dynamics of drop deformation. The Taylor Reynolds num-
fluid particles in the inertial subrange of the turbulence spec- ber was about 40 and the 2 mm droplets were in the inertial
trum (Coulaloglou and Tavlarides, 1977; Hsia and Tavlarides, subrange. The results were in agreement with the experimen-
1983; Tsouris and Tavlarides, 1994; Prince and Blanch, 1990; tal data of Galinat et al. (2005, 2007). These results indicates
Chen et al., 1998; Martínez-Bazán et al., 1999a,b; Hagesaether, that for low turbulent flows and millimeter-sized droplets,
2002, 2002b; Andersson and Andersson, 2006) but, for the sub- breakage is a non-local phenomena.
Kolmogorov size range, there is just one model for coalescence There are several works that analyzed the behavior of
(Chesters, 1991) but no model for breakage. Besides, there is liquid–liquid dispersions in turbulent flows, but we focused on
no model that can be continuously applied for all length scales those works that developed breakage or coalescence models
(Chesters, 1991; Kamp et al., 2001; Joshi, 2001; Lasheras et al., for liquid–liquid emulsions.
2002; Jakobsen and Dorao, 2005; Frank et al., 2007; Liao and Chen et al. (1998) carried out experiments in a continu-
Lucas, 2009, 2010). ous flow screw-loop reactor using a water–paraffin oil system
There are a few reviews of breakage and coalescence of in order to validate their breakage model. The droplet size
fluid particles (Chesters, 1991; Lasheras et al., 2002; Liao and distributions of the emulsions were determined by a laser
Lucas, 2009, 2010) that describe and compare most of the diffraction method. The droplet diameter ranges from 0.4 to
existing models. The coalescence models are based on similar 200 ␮m. Chen et al. (1998) modified the model of Coulaloglou
hypotheses being quite similar to each other. However, there and Tavlarides (1977) to include the internal viscous defor-
are some inconsistencies in almost all of the breakage models mation forces. The experimental data was used to model the
(Lasheras et al., 2002). probability distribution of the daughter droplets. Chen et al.
Coalescence in liquid–liquid and gas–liquid systems are (1998) adjusted the 6 model parameters and stated that the
considered to behave differently only because of their dif- model describes the process fairly well. Droplet coalescence
ferent interface behaviors (Chesters, 1991). The interface can was neglected.
be assumed to be rigid or deformable. Being deformable, the Raikar et al. (2009) experimentally analyzed the produc-
interface may be immobile, partially mobile or fully mobile tion of soybean oil-in-water emulsions in a high pressure
(Chesters, 1991; Liao and Lucas, 2010). homogenizer. Droplet size distributions were measured in a
Some models (Coulaloglou and Tavlarides, 1977; Hsia and light-scattering equipment. They used two models for the
Tavlarides, 1983; Tsouris and Tavlarides, 1994; Chen et al., droplet breakage rate. The first one assumes that breakage
1998) were developed for liquid–liquid dispersions and others originates from droplet collision with turbulent eddies, being
(Prince and Blanch, 1990; Luo and Svendsen, 1996; Martínez- derived following Chen et al. (1998) but neglecting the effect
Bazán et al., 1999a,b; Hagesaether, 2002b; Wang et al., 2003) of the dispersed phase viscosity on drop deformations. The
for gas–liquid systems. However, there is no feature that dis- second model assumed that breakage occurs due to turbulent
tinguishes the application of these models for gas–liquid or shear. They solved a pure breakage PBE model by a sec-
liquid–liquid systems. Martínez-Bazán et al. (1999a,b) pointed tional method using a 100-node uniform mesh. Their model
out that their breakage model should be applied only to assumed binary breakage with a truncated normal distribu-
gas–liquid systems due to the assumption of negligible inter- tion as the daughter droplet probability distribution. There
nal viscous deformation forces. Some works (Eastwood et al., are 4 parameters in this model that were estimated from the
2004; Andersson and Andersson, 2006; Maaß et al., 2010) drop volume distributions. The model predict only qualitative
showed that the internal flow is critical to differentiate trends of the evolution of the drop size distribution but failed
between the breakage of drops and bubbles and that this may to predict the effect of the pressure drop in the equipment.
explain the large number of particles formed in drop breakage. Afterwards, Raikar et al. (2010) improved the model by
Several works analyzed single drop breakage experiments. including a model for drop breakage due to laminar shear,
Eastwood et al. (2004) analyzed the breakage of drops in a considering non-binary breakage with the Hill-Ng distribu-
turbulent water jet, showing that drops usually are consid- tion. They showed that the formation of multiple daughter
erable deformed prior to the breakage event, specially for drops could explain much better the experimental data on
highly viscous dispersed phases. This behavior was attributed the droplet size distribution, which were best represented
to drop interaction with large turbulent scales. Their data with a mean number of daughters equal to 20. Their results
showed the occurrence of multiple daughters on drop break- were good for large surfactant–oil ratios, when the effect of
age. Andersson and Andersson (2006) observed drop breakage coalescence was small. However, the accuracy of the experi-
events in a static mixer, confirming that the binary breakage mental data was not enough to determine the values of the
is not a good assumption for drops. two parameters of the model for the breakage rate due to
Galinat et al. (2005, 2007) analyzed the drop breakage down- laminar shear.
stream of a pipe restriction. Their single drop experiments Recently, Maindarkar et al. (2012) incorporated the coales-
indicated that the number of daughter droplets is an increas- cence model of Coulaloglou and Tavlarides (1977) to Raikar
ing number of the global Weber number, which they defined et al. (2010) model. The breakage rate was assumed to be the
using the pressure drop through the orifice. For dispersed sum of the rates originated from the mechanisms of drop-
phase fractions lower than 20%, Galinat et al. (2007) found eddy collisions and turbulent shear (Raikar et al., 2009). The 6
out that the breakage in the dispersions was similar to that model parameters were adjusted and the best fit was obtained
observed in the single droplet experiments. For larger dis- considering a mean value of 80 daughters in the breakage
persed phase fractions, the turbulence field was modified, process.
2496 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

Raikar et al. (2011) aimed to increase the model applica-


bility over a wide range of pressures. The most significant
improvements were obtained by increasing the mean num-
ber of daughter drops in a breakage event to 20–150 and by
introducing a maximum stable diameter, below which drops
could not break (Vankova et al., 2007). They concluded that
the usage of an appropriate number of daughter drops and a
critical droplet diameter were essential to model the emulsion
formation.
All the above cited works analyzed breakage and coales-
cence in emulsions that were not made with crude oil.
Recently, Keleşǒglu et al. (2013) determined droplet size dis-
tributions for water-in-crude oil emulsions with digital video
microscopy and nuclear magnetic resonance, but they did not
attempted to analyze its evolution during the emulsion flow.
Our aim was to develop breakage and coalescence models
for the turbulent flow of water-in-crude oil emulsions through
valve-like elements. Experimental data were obtained by flow-
ing the emulsion through a duct containing a variable accident
that acts like a valve. A population balance model was used
to relate the inlet and outlet droplet size distributions. A new
breakage model was proposed based on the theoretical work of
Cristini et al. (2003) and a coalescence model based on previous Fig. 1 – The test section: (a) side view of the whole section
knowledge was employed. The parameters of these models and (b) side view of the flow path restricted by the gates.
were adjusted considering the experimental uncertainty of The duct depth is 5 mm. ˝ is the central valve opening
all variables. To the authors’ knowledge, this is the first time distance. The effective volume of the valve-like element
that breakage and coalescence models have their parameters was modeled as the sum of volumes V1–V5.
estimated for an emulsion with an actual crude oil.

experimental tests. This limited the size of the cross section


area of the test section.
2. Materials and methods

2.1. Premises of the experimental setup design 2.2. The test section

As stated above, we were interested in studying the effect of The test section consists of a duct of square cross section with
a valve-like element, similar to a mixing valve, on the size 5 mm in side and 200 mm in length. The central part of the
distribution function of water droplets in a emulsion of dead duct has three movable gates with 5 mm width. These gates
crude oil. This imposes severe restrictions on the design of the can be precisely put at chosen positions using dial indicators
experimental setup. (Digimess, model 121.302) with 0.01 mm accuracy, generating a
First, in order to isolate the effect of the valve-like element, large variety of geometric configurations for this valve model.
the flow in the test section must be laminar in its absence Fig. 1 shows the details of this test section.
and, when it is present, it must generate enough turbulence
to break the droplets. Therefore, a test section with mov- 2.3. Experimental setup
able gates was designed and the operational conditions were
defined to meet these requirements. The flowchart of the experimental apparatus is shown in Fig. 2.
Second, as the crude oil is completely opaque, there is no A 5000-l oil tank (T-01) has a 60 l/min double diaphragm pump
optical access to measure the flow variables, as the turbulence (DP-01, Versa-matic, model E1) for homogenization. From the
characteristics or the droplet size distributions. Therefore, tank, a progressive cavity pump (P-01, Weatherford with rotor
the analysis must consider only global variables, that is, the HD 20, 6.8 l/min) makes the oil flows through a 18 kW heater
droplet size distribution functions at the inlet and outlet of (H-01, Eltra, model ADO-180) to facilitates its emulsification in
the test section and an estimate of the turbulence induced by the mixer (M-01, an inverted flow venturi). Water is fed to the
the valve-like element. A procedure to sample the emulsion mixer from an insulated 100-l tank (T-02) with a 3 kW electrical
to determine the droplet size distributions was developed and resistance by a second progressive cavity pump (P-02, Netzsch
validated. with rotor NM11). Pumps P-01 and P-02 are driven using fre-
Finally, the dead crude oil must not contain any additives, quency converters (Adleepower, model AS2-IPM) in order to
demulsifiers or surfactants, in order not to affect the coales- independently control their flow rates to fix the concentration
cence and breakage phenomena. The addition of demulsifiers of the water in the emulsion within 1% of the desired value.
in crude oil prior to refining is common practice (Al-Sabagh It should be pointed out that the emulsion generator always
et al., 2011) but this impedes the generation of the emul- operates in the same conditions in order to minimize vari-
sion and, thus, the analysis of the coalescence and breakage ations in the produced emulsion. Two pumps (P-03, P-04,
phenomena. Therefore, special care was taken to get a large Weatherford with rotor HD 15, up to 3 l/min) in parallel,
sample of a crude oil without any additives. However, due driven using frequency converters (Adleepower, model AS2-
to storage limitations, the emulsion flow rate had to be kept IPM), feed part of the overall emulsion flow rate to the test
smaller than 6 l/min to allow the execution of all the planned section. A 500 W line heater (H-02) placed just before the test
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2497

Drain
T-01

Crude
Oil DP-01

P-03
Drain
(emulsion)

P-01 H-02
(crude oil) TI TI

T-02
H-01
Test Section
TI Hot
Water
TI P-04
(emulsion) PI PI PI
PI

Shear
M-01
Valve Sample A Sample B
P-02
(water)
PI

T-01 Crude Oil Tank Progressive Cavity Pump "T" Mixer

T-02 Water Tank (heated)


PI Pressure Indicator
TI Temperature Indicator Diaphragm Pump
Globe Valve
Check Valve
Ball Valve
Heater

Fig. 2 – Schematics of the experimental setup.

section controls the emulsion temperature in the test section the emulsion was determined from a sample by the volumetric
within ±1 ◦ C for all runs. (Mettler Toledo, Model DL38) and coulometric (Mettler Toledo,
Preliminary tests showed that the emulsion generated in Model DL37) Karl Fischer titration methods (Paiva, 2004). Oil
the conditions described above had droplets too small to break properties at the desired operational temperature are given in
further in the test section. Therefore, before the test section, a Table 1.
tube with 12.7 mm in diameter and 20 meters long arranged in
a spiral with a diameter of 2 m (see Fig. 2) was included. This 2.4. Experimental procedure
long tube with large cross section area proved to be adequate
to promote droplet coalescence in order that the droplet sizes In order to guarantee the homogenization of the oil in the tank,
at the inlet of the test section were large enough to break by pump DP-01 was used to circulate it during 2 h at 60 l/min.
the turbulence generated by the valve-like element. Then, oil and water were fed to the mixer at the specified flow
There are drains before and after the test section. The rates for the given experimental point. Finally, the gates of the
emulsion flow can be completely deviated from the main line test section were positioned to the desired gaps.
to each drain in order to sample the emulsion flow for the After the stabilization of temperature, flow rate and pres-
experimental determination of the drop size distribution func- sure drop across the test section, an additional waiting period
tion. of 5 min was taken before the emulsion was sampled. Emul-
Flow rate in the test section is measured by an oval gear sion samples were first obtained for the test section outlet and
flow meter (Metroval, model OI-06Ag19/F5) with a 0.3% accu- then for its inlet, in order not to disturb the flow, because the
racy. The pressure drop in the test section was determined sampling procedure diverts the overall flow rate to the corre-
by pressure transmitters (Yokogama, model EJX430A-EAS4G- sponding drain.
912EB) with 2% accuracy. The operating temperature was
measured before and after the test section by two resistance
thermometers (PT-100 Ohm). Table 1 – Properties at 60 ◦ C.
Oil density, viscosity and water–oil surface tension were
Property Value
determined as a function of temperature in the range 55–65 ◦ C
by a glass density meter with accuracy better than 5 kg/m3 , a (kg/m3 ) 865
Brookfield LV-DVII+viscometer with 1.0% accuracy and a Du  (Pa s) 0.0115
(m2 /s) 1.33 × 10−5
Noüy ring tensiometer (Kruss, K9 model) with accuracy better
(N/m) 0.0223
than 0.1 mN/m, respectively. The water mass concentration in
2498 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

The samples were collected manually by taking emulsion 4.3 kg/min, respectively. The pressure loss in the test section
drops using a glass stick and then put in a beaker containing varied in the ranges 0.04–0.18 bar when the gates were fully
a highly viscous mineral oil to prevent water droplets coales- opened and 0.5–8.40 bar for the other conditions. Therefore,
cence. This oil was transparent, making possible the use of the except for the blank experiments, most of the pressure drop
particle analyzer by laser diffraction. The beakers with mineral in the test section was caused by the valve-like element.
oil were put to rest for 24–48 h in order to eliminate the bubbles From the 103 runs shown in Table 2, 19 were those with
that may be entrained in the sampling procedure. The sam- fully opened gates that were performed to guarantee that the
ples were homogenized using a magnetic stirrer very gently in valve-like element was the main factor affecting the evolution
order not to produce drop coalescence. Then, the normalized of the droplet size distribution in the test section. They cannot
droplet size distribution was determined for each sample. be used in the estimation of the model parameters. Other six
This determination was carried out using a laser diffraction runs were eliminated due to inconsistencies in the measured
particle sizer (Malvern® Mastersizer 2000). Such equipment drop size distributions. Therefore, 78 valid experimental runs
was modified to allow that all sample volume can be flowed were used for parameter estimation.
through the detector by gravity without any agitation or suc-
tion in order to avoid changing the droplet size distribution. 3. Population balance model
Under ordinary conditions, this analyzer provides an error of
10% in the determinations of the normalized volume distribu- The mean number of particles per unit volume of the parti-
tion function of droplets and errors around 2% for the mean cle state space is the particle number distribution function, f.
diameters. Considering only one internal variable, the particle volume, v,
f (t, v, z) is called the particle number size distribution func-
2.5. Experimental conditions tion (PSDF). In order to predict the evolution of the PSDF,
it is necessary to solve the population balance equation
The test conditions are based on a full factorial experimental (PBE). Considering breakage and coalescence of particles and
planning with two levels of water concentration, three levels neglecting nucleation and particle growth, the PBE can be writ-
of mass flow rate and five levels of positioning of the gates. ten in the following Lagrangian form (Ramkrishna, 2000):
The temperature of all runs were nominally set at 60 ◦ C.
The two external gates of the test section were kept in Df ∂f  
the same positions, defining openings of 2.0 mm for all runs. = + Ż · ∇z f (v, z) = H( f ; v, z, y) − f (v, z)∇z · Ż (1)
Dt ∂t
The opening of the central gate, ˝, was varied among four
positions (see Fig. 1). This simulates the characteristics of a where Df/Dt is the material derivative of f, Ż is the velocity
globe valve. The chosen values of ˝ were 0.5, 1.0, 1.5 and of the particle with volume v at spatial position z, y is the
2.5 mm. Besides, a blank experiment, where all gates were vector of continuous phase properties and H is the particle
fully opened, was added to the experimental planning in order net production rate:
to periodically check if the drop size distributions before and
after the valve-like element were approximately equal. The H = BB − DB + BC − DC (2)
blank experiments were also used to show that the procedure
of sampling and determination of the droplet size distribu-
where BB , DB , BC and DC represent, respectively, the parti-
tions gives reproducible results.
cle production (birth) and destruction (death) by breakage
The nominal water concentrations were selected to be 8
and the birth and death by coalescence, which are given by
and 15% (w/w), which reflect typical industrial operational
(Ramkrishna, 2000):
conditions. The selected emulsion mass flow rates were 1.7,
3.0 and 4.3 kg/min. These values were chosen in order to guar- 
  ∞
antee that the flow was laminar in the test section with all BB ( f ; v, z, y) = ς(v , y)P(v|v , y)b(v , y)f (v , z)dv (3)
gates fully opened. v

For each experimental condition, up to 5 replicates were  


attempted. The experiments were grouped by mass flow rate DB ( f ; v, z, y) = b(v, y)f (v, z) (4)
and water concentration, but the position of the central gate

was randomized. In order to identify the experimental data,   1
v

a code was established to each run, given by: ptXgY, where BC ( f ; v, z, y) = f (v − v , z)f (v, z)a(v − v , v , y)dv (5)
2 0
X and Y are integer numbers varying in the 1–10 and 1–15

ranges, respectively. In this code, Y is related to the replicate at   ∞
a mass flow rate level and X is related to the gate positioning DC ( f ; v, z, y) = f (v, z)f (v , z)a(v, v , y)dv (6)
and water concentration levels. Table 2 associates each run 0

code to its conditions. The dash in this table indicates that


In the above equations, it was assumed that the breakage
the corresponding run was eliminated from the analysis due
and coalescence functions are affected by the position only
to detected experimental problems.
through y. From now on, in order to simplify the notation, y
From the 150 performed runs, 47 were discarded due to
and z are omitted.
the problems detected at the experimental level. For the
Considering volume conservation in the breakage process,
remaining 103 runs, the temperature varied within 56–62 ◦ C,
the conditional daughter particle probability density function,
whereas the measured water concentrations varied in the
P(v |v), has the property (Ramkrishna, 2000):
ranges of 8.1–8.5 and 14.5–14.6% (w/w) for the 8 and 15%
(w/w) nominal values, respectively. The measured emulsion  v
mass flow rates varied within the ranges of 1.67–1.72, 2.97–3.01 ς(v) P(v |v)dv = v (7)
and 4.25–4.37 kg/min for the nominal values of 1.7, 3.0 and 0
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2499

Table 2 – Code map for all experimental runs.


Nominal mass Nominal water concentration (w/w, %)
flow rate (kg/min)

15 8

˝ Open 2.5 1.5 1.0 0.5 0.5 1.0 1.5 2.5 Open
Y\X 1 2 3 4 5 6 7 8 9 10

4.3 1 – – – – – – – – – –
2 pt1g2 pt2g2 – – pt5g2 pt6g2 pt7g2 pt8g2 pt9g2 pt10g2
3 pt1g3 pt2g3 pt3g3 pt4g3 pt5g3 – pt7g3 pt8g3 pt9g3 pt10g3
4 pt1g4 pt2g4 pt3g4 pt4g4 pt5g4 pt6g4 pt7g4 pt8g4 pt9g4 –
5 – pt2g5 pt3g5 pt4g5 – pt6g5 pt7g5 – pt9g5 pt10g5

1.7 6 – – – – – – – – – –
7 pt1g7 pt2g7 pt3g7 pt4g7 pt5g7 pt6g7 pt7g7 pt8g7 pt9g7 pt10g7
8 pt1g8 pt2g8 pt3g8 pt4g8 pt5g8 – pt7g8 – – –
9 pt1g9 pt2g9 pt3g9 pt4g9 pt5g9 pt6g9 – pt8g9 pt9g9 pt10g9
10 pt1g10 pt2g10 – – – – pt7g10 pt8g10 pt9g10 pt10g10

3.0 11 pt1g11 pt2g11 pt3g11 pt4g11 pt5g11 pt6g11 pt7g11 pt8g11 pt9g11 pt10g11
12 pt1g12 pt2g12 pt3g12 – pt5g12 pt6g12 pt7g12 pt8g12 pt9g12 –
13 – pt2g13 pt3g13 pt4g13 – pt6g13 – pt8g13 pt9g13 pt10g13
14 – – pt3g14 pt4g14 pt5g14 – – pt8g14 pt9g14 pt10g14
15 pt1g15 pt2g15 pt3g15 – pt5g15 pt6g15 pt7g15 pt8g15 pt9g15 –

3.1. Simplified model for experimental analysis Therefore, Ve was firstly considered to be equal to the volume
shown in this figure. Afterwards, this assumption was verified.
As the PB-CFD simulation demands a large computational Moreover, as the experimental pressure drop in the test
effort, its usage for determining the parameters of the break- section is almost completely due to the valve-like element, the
age and coalescence models by optimization is nowadays mean dissipation rate of turbulent energy, , was estimated by
impractical. Besides, only the normalized PSDFs at the out-
let and inlet were known. Thus, we decided to use a simplified p
= (10)
Lagrangian model based on Eq. (1). tres
Our simplified model assumes that the phases are incom-
pressible and all droplets move with the continuous phase where is the oil density.
velocity. In this case, the mixture continuity equation is
∇z · Ż = 0 and the last term in Eq. (1) vanishes, giving for the
3.2. The coalescence model
steady-state:

Df   In an attempt to develop a model that can be applied to any


= Ż · ∇z f (v, z) = H( f ; v, z, y) (8) range of droplet sizes observed experimentally, the coales-
Dt
cence model was designed to be used for particles in the
These assumptions lead to phase fractions that are uniform dissipation and inertial size ranges.
in the test section. According to Chesters (1991) the coalescence frequency, a,
Furthermore, as the PSDF change occurs due to the pres- is given by
ence of the valve-like element, this model also assumes that
the final distribution can be obtained by integrating Eq. (8) in
a(d, d ) = (d, d )(d, d ) (11)
t up to the mean hydrodynamic residence time, tres , which is
defined by
where  is the coalescence efficiency and  is the collision
Ve frequency, which is given by
tres = (9)
Q

where Ve is the effective volume of the valve-like element (d, d ) = Cc Sur , (12)
and Q is the volumetric flow of the emulsion. Even though
droplet coalescence and breakage occur in different regions, where the sectional area, S, and the relative velocity, ur , are
this model assumes the same region for both phenomena, given by
which is equivalent of using mean coalescence and breakage
functions as usually done in well-mixed reactors (Coulaloglou 
(d + d )
2
S= (13)
and Tavlarides, 1977). 4
The effective volume of the valve-like element can be
considered another parameter of the model, whose value is and
different for each experimental run. However, previous stud-
ies (Lage et al., 2006) showed that the droplets break mainly ud + ud
ur = (14)
nearby the valve-like element, in a region whose volume is 2
somewhat constant and may be approximated by the sum
of the volumes V1–V5 which are represented in Fig. 1(b). being Cc a parameter which must be adjusted experimentally.
2500 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

The collision frequency was estimated using the character- According to Martínez-Bazán et al. (1999a,b), the breakup
istic velocity modeled as (Chesters, 1991): frequency is given by

⎧ ⎧ 
⎨ d  , for d <  ⎪
⎪ 2Wecrit

⎨ ˇ(d)2/3 −
ud = (15) d
⎩ 1/3 b(d) = Cb , for d > dc,mb (19)
(d) , for d≥ ⎪

d


0, for d ≤ dc,mb
Four alternative models were tested for the coales-
cence efficiency (Chesters, 1991), corresponding to different
where ˇ = 8.2, Wecrit = 6 (the Weber number in this model is
hypotheses regarding the droplet interface behavior: rigid
defined as We = ˇ(d)2/3 d/2 ) and the critical particle diameter,
interface, deformable partially mobile interface, deformable
dc,mb is given by
immobile interface and deformable mobile interface. From
these models, only the first two gave good results and, thus,  2We 3/5
only them are described in the following. dc,mb = crit
−2/5 (20)
ˇ
The coalescence efficiency for droplets with rigid interfaces
is modeled as (Chesters, 1991): Cristini et al. (2003) theoretically developed an expression
 to the breakage frequency that can be written as
1 hi
 = exp −Ce ln (16) √
4 hf  3  / d
b(d) = Cb,c Ca , Ca = (21)
2
where hi and hf are, respectively, the initial and final thick-
for Ca > Cacrit , being zero otherwise.
nesses of the thin film formed between the droplets under
It is essential that the breakage model is a continuous func-
collision. The value of hf can be estimated by
tion of its variables. Therefore, it was assumed that there is a
1/3 transition diameter such that the breakage frequencies given
Adeq by Eqs. (19) and (21) have the same value. This transition diam-
hf = (17)
16 eter, dl , was assumed to be given by

 3 6/5
As the value of hi is unknown, the ratio hi /hf is sometimes
dl = dc,mb ∼
= 1.627 dc,mb (22)
assumed to be an unknown constant which leads to a constant 2
coalescence efficiency which is incorporated in the adjusted
parameter Cc . This hypothesis is usually necessary when the which is the point where the breakage frequency given by Eq.
values of  are in a narrow range which makes the estima- (19) reaches its maximum value.
tion of Ce very difficult. This approach was used in the present Thus, by equating the breakage frequency given by Eqs. (19)
work. and (21) at d = dl , a relation between the empirical parameters
The coalescence efficiency for droplets with deformable Cb,c and Cb can be derived, which can be written as:
and partially mobile interfaces is modeled as (Chesters, 1991):
63.927
 d 4/5
Cb,c = Cb 11/5
Ca−4/5 (23)
Wecrit 2
√    √
3 d 3/2 deq  / deq
 = exp −Ce Caeq , Caeq =
8  hf 2 where the value of ˇ was used.
(18) Using Eq. (23), the combined model for breakage frequency
is given by

where hf was given by Eq. (17) with A = 0.6 × 10−20 J. This value ⎧


Cb 2/3
ˇ(d) −
2Wecrit
, for d > dl
was chosen because it is the center value of the experimen- ⎪
⎪ d d

tal range of the Hamaker constant for water-water interaction  4/5
b(d) = 63.927  2.2 d
through a hydrocarbon phase given by Israelachvili (2011) in ⎪
⎪ Cb Ca , for d ≤ dl and Ca > Cacrit


11/5
Wecrit 2
his Table 13.3. Any error in this A value is compensated by the ⎩
adjusted Ce value. 0, Ca < Cacrit
(24)

3.3. The breakage model


which have three empirical parameters, Cb , the critical Weber,
The breakage model was also designed to be used for any range
Wecrit , that determines dl , and the critical capillary number,
of particle sizes. It is composed by the Martínez-Bazán et al.
Cacrit , defined by
(1999a) model and one developed based on the work of Cristini
et al. (2003). √
 / dcrit
The breakage model of Martínez-Bazán et al. (1999a) was Cacrit = (25)
2
selected because it can be applied to non-binary particle
breakage. Moreover, the model of Martínez-Bazán et al. (1999a) Cacrit was modeled by assuming that dcrit is the minimum
has extensive theoretical basis and experimental validation attainable droplet diameter at the outlet of the test section
when compared with other models that could have been as most of the data are for breakage dominant conditions.
selected (Lasheras et al., 2002; Liao and Lucas, 2009), even Cristini et al. (2003) suggested that Cacrit is inversely pro-
though it was developed for bubbles. portional to the flow Reynolds number. Therefore, we tried
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2501

to use Cacrit ∝ Re−C


max , where Remax was defined based on the
Re Following Kumar and Ramkrishna (1996), the method of
maximum fluid velocity in the valve-like element: classes was applied to Eq. (8) leading to the following expres-
sion:
(Q/Ac,min )c Q
Remax = = (26)
l

DNk 
i≥j
where Ac,min = lc is the smallest flow sectional area (l = 5 mm). = (1 − 0.5ıij )ijk ai,j Ni Nj
However, this formula could not represent the minimum Dt
i,j
attainable droplet diameter. xk−1 ≤ (xi + xj ) ≤ xk+1
Since the probability of breaking the smallest droplet in the
distribution depends on its interaction with a high energy tur- 
n 
n
− Nk Ni ai,k + ςi ϕk,i bi Ni − Nk bk (30)
bulent eddy, this probability increases with the period of time
i=1 i=k
in which the droplet is in the effective volume of the valve-
like element. Therefore, it is reasonable to assume that Cacrit
depends on the Stokes number defined by the ratio between where  i,j,k and ϕk,i are defined as:
the residence time and the turbulent time scale. Thus, we
⎧ x − (x + x )
assumed the following functional dependency: ⎪ k+1 j i
⎨ x − x , for xk ≤ (xj + xi ) ≤ xk+1
k+1 k
i,j,k = (31)
Cacrit = CCa StK Re−C
max ,
Re
tres
StK = √ (27) ⎪
⎩ (xj + xi ) − xk−1 , for x ≤ (x + x ) ≤ x
/ xk − xk−1 k−1 j i k

where CCa and CRe are experimental parameters. The min-


imum attainable droplet diameter at the outlet of the  xk+1  xk
test section could be correlated with CCa = 1.65 × 10−4 and xk+1 − v v − xk−1
ϕk,i = P(v | xi ) dv + P(v | xi ) dv
CRe = 3/20. xk
xk+1 − xk xk−1
xk − xk−1
Recent works showed that the breakage of drops is not (32)
binary (Eastwood et al., 2004; Galinat et al., 2005; Andersson
and Andersson, 2006; Raikar et al., 2010; Maaß et al., 2010), Further details about these equations can be found in the orig-
which must be considered in the modeling of the daughter inal reference.
particle probability distribution. In this work, several models It is also of interest the particle volume size distribution,
were attempted. The only one that gave good results is the F(v) = vf (v), as well as its zeroth-order sectional moments:
simple model that assumes that ς equal daughter droplets are
formed upon breakage:  vk+1
Nv,k = F(v)dv ∼
= xk Nk (33)
v
 vk
 
P(v |v) = ı v − (28)
ς
4.2. Determination of the inlet and outlet discrete
This introduces the mean number of daughter droplets, ς, as PSDFs
an adjusted parameter in the model.
The particle size analyzer provides only the first order sec-
4. Numerical procedure tional moments of the normalized particle size distributions
for all size classes, which can be given either for the number,
4.1. Solution of the population balance model Nn , or the volume, Nv,n , distributions. However, the inlet con-
ditions for Eq. (30) are the Nk values of the non-normalized
Among the several numerical methods that have been devel- particle number size distribution, defined by Eq. (29). How-
oped for the solution of the population balance equation, this ever, the dispersed phase volume fraction, , can be calculated
work used the method of classes of Kumar and Ramkrishna from the measured water concentration and the properties of
(1996) enforcing the conservation of number and volume of both phases, allowing the computation of the droplet number
the particle population. This method was chosen because it density by:
allows the direct usage of the particle size classes defined by
the particle size analyzer. Therefore, the experimental data is  6
NT = ∞ = (34)
used to define the discrete drop size distributions without any vf̂ (v)dv d330
0
manipulation.
The sectional moment of the k class, Nk , includes all parti-
where f̂ is the normalized drop number distribution and d30 is
cles with volume between vk and vk+1 , being defined by
the mean volume diameter, which is given by the particle size
 vk+1 analyzer. Then, the Nk values can be determined by:
Nk = f (v)dv (29)
vk
 vk+1  vk+1
Nk = f (v)dv = NT f̂ (v)dv = NT Nn,k (35)
vk vk
This class is represented by the pivot xk = 0.5(vk + vk−1 ). The
classes vk are defined by the particle size analyzer (Malvern®
Mastersizer 2000), being given by the geometric grid defined by where Nn,k is the zeroth order sectional moment of the nor-
vk+1 /vk = 1.514, k = 0, . . ., n, v0 = 5.24 × 10−7 ␮m3 (d0 = 0.01 ␮m), malized particle number distribution for class k as given by
with n = 100 classes. the particle analyzer.
2502 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

4.3. Orthogonal distance regression For the comparison between the experimental and sim-
ulated results of a variable i of experiment i, we used the
The parameter estimation problem was solved using the relative error defined by:
routine ODRPACK95 (Zwokak et al., 2004). This routine imple-
ments the orthogonal distance regression, which consists of i,sim − i,exp
ei = (38)
minimizing the orthogonal distance between the experimen- i,exp
tal data and the prediction of a model that considers the errors
both on the response and explanatory variables. to calculate the following mean errors:
Consider a multiresponse model ˚ : Rp+m → Rq given by
1 1
M M
e= ei , |e| = |ei | (39)
i = ˚(
i + ıi ; ˛) − εi (36) M M
i=1 i=1

where
i ∈ Rm , m≥1 are the multivariate explanatory variables,
The value of |e| is the mean of the magnitude of the rela-
i ∈ Rq , q ≤ 1 is the multiresponse dependent variables, ˛ ∈ Rp
tive errors and it represents how well a variable is reproduced
are the model parameters, ıi and εi are the estimated errors in
by the adjusted model, whereas e is the mean of the relative

i and i , respectively, and i = 1, . . ., M is the experiment index.


errors that should be zero for an unbiased model when M is
The orthogonal distance regression implies in the mini-
large.
mization of the objective function:

4.4. Global optimization



M 
M

G= εTi Wεi εi + ıTi Wıi ıi (37)


i=1 i=1
The global optimization algorithm called AIU-PPSO (Asyn-
chronous and Immediate Update Parallel Particle Swarm
where Wıi and Wεi are the matrices of weights associated Optimization) developed by Moraes (2011) was applied to esti-
with ıi and εi , respectively. They were assumed to be diagonal mate the parameter values of the population balance model,
whose elements were given by the inverse of the variances of including the effective volumes of the valve-like element for
the corresponding variables. all experiments Ve,i , i = 1, . . . M, as new model parameters.
The numerical procedure used in ODRPACK95 is described This global optimization was performed to confirm the
in Boggs et al. (1987, 1989) and Zwolak et al. (2007). The solution validity of the procedure described in the previous section
is found iteratively using a trust region Levenberg–Marquardt for locating the global minimum using the ODRPACK and the
method with bound constraints and scaling. Therefore, it is a assumption used to calculate the values of the effective vol-
method of local minimization of the objective function. This ume of the valve-like element. For comparing the modeled Ve,i
work used the default values for the scaling. Further details values and their estimated values determined by the AIU-PSO
can be found in the literature (Boggs et al., 1987, 1989; Zwokak algorithm, error definitions given by Eqs. (38) and (39) were
et al., 2004; Zwolak et al., 2007). used.
The explanatory variables were chosen to be the effective More details of the AIU-PPSO are given in the work of
volume of the valve-like element, Ve , the mass flow rate, the Moraes (2011).
mean temperature at the test section, the pressure loss in
the test section and the zeroth-order sectional moments of 5. Results and discussion
the number particle size distribution, Nk , at the inlet of the
test section. In an alternative analysis, Nk were substituted by 5.1. Analysis of the experimental flow conditions
Nv,k . The response variables were the zeroth-order sectional
moments of the corresponding distributions at the outlet of Considering the oil properties shown in Table 1, the Reynolds
the test section. The parameters are those related to the break- number (Rel = lU/ ) for the pure oil flow in the test section with
age and coalescence models (Cc , Ce , Cb and ς) being just 3 all gates opened (blank experiments) are 493, 870 and 1246,
parameters for the model with constant coalescence efficiency for the 1.7, 3.0 and 4.3 kg/s flow rates, respectively. Therefore,
(Cc , Cb and ς). the flow is laminar and very little effect on the droplet size
As explained later, the experimental runs could not be distributions was expected for the blank experiments.
replicated. Therefore, the variances of the explanatory vari- For the flow in the test section with the gates partially
ables were assumed to be the square of the experimental closed, the experimental conditions of three different cases
errors. For the mass flow rate, the mean temperature and the were used to estimate the Taylor scale Reynolds number, ReT .
pressure drop in the test section, we used the measurement Two cases were chosen because they have the minimum and
errors. For the Nk or Nv,k , the 10% error reported by the particle maximum flow rates and the same minimum clearance, c. As
sizer manual was used. A 20% error was assumed for the Ve a matter of fact, most experiments has c = 1 mm due to the
value of each experimental run. small horizontal aperture between the gates. The third case
Due to the large computational cost of using the ODR- was one of those with the largest pressure drop, caused by the
PACK95 considering the experimental errors in all explanatory smallest gap in the central gate, ˝ = 0.5 mm.
variables, an exploratory optimization study was first carried Again, pure oil flow was considered with the properties
out without considering such errors. The minimal point of this shown in Table 1. Table 3 summarizes the calculations. The
study was the initial condition used in the main optimization maximum velocity, Umax , is calculated using the minimum
run that includes all the errors in all explanatory variables. area normal to the flow in the test section, Ac,min = cl. The
The validity of this procedure in obtaining the global mini- estimation of the turbulent variables follows Pope (2000),
mum point was verified by performing a global optimization including the usage of his spectra model to determine the ratio
as explained in the next section. between the longitudinal integral scale, L11 , and the length
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2503

Table 3 – Turbulence conditions for pure oil flow. (a) 12


inlet
Variable Conditions of run outlet
10
pt6g2 pt9g2 pt9g9
8 blank experiments:
Data pt1g2

Nv,n (%)
c (mm) 0.50 1.00 1.00 pt1g3
6 pt1g4
Ac,min (mm2 ) 2.50 5.00 5.00
W (kg/min) 4.3 4.3 1.7
4
Umax (m/s) 33.1 16.6 6.6
p (105 N/m2 ) 8.22 3.37 0.54
2
Estimates
u = 0.2Umax (m/s) 6.6 3.3 0.55
2
0
k = 1.5(u ) (m2 /s2 ) 66. 16. 2.6
5 2 3
 (10 m /s ) 6.7 2.5 0.16 101 102 103
 (␮m) 7.7 9.9 19.7 d (micra)
Le = k3/2 / (mm) 0.80 0.27 0.26
ReLe = k1/2 Le / 487. 82. 32.
L11 /Le a 0.65 0.80 0.90 (b) 12
inlet
L11 (mm) 0.52 0.22 0.24 outlet
Ret = u L11 / 259. 54. 23. 10
ReT = u LT / 72. 33. 22.

a
8 blank experiments:
Estimated from Fig. 6.24 of Pope (2000). pt10g2
Nv,n (%) pt10g3
6 pt10g5
scale Le = k3/2 /. The rate of dissipation of turbulent kinetic
energy was estimated from the pressure loss in the test sec- 4
tion, using Eq. (10), which is described in Section 3.
The estimates for the Taylor Reynolds number indicates 2
that most of the experiments were for low turbulence flow
0
conditions (ReT  30). In this case, the inertial subrange does
not really exist. The Kolmogorov length scale, , varied 101 102 103
between 7 and 20 ␮m and L11 was in the [220, 520] ␮m range. d (micra)
As the largest droplets were 100 ␮m in diameter, all droplets
Fig. 3 – Normalized droplet volume size distributions for
should have sizes in the turbulence dissipation range.
the blank experiments for runs (a) pt1 and (b) pt10. The
bars represent the standard deviation calculated from the
5.2. Analysis of the blank experiments
three experiments.

The blank experiments corresponding to the largest flow rate


were selected for this analysis. They correspond to the runs concentration of 15% (w/w) and mass flow rate of 3 kg/min.
pt1g2, pt1g3 and pt1g4 for the 15% (w/w) water-in-oil emulsion Ideally, all these distributions should be equal. The lack of
and the runs pt10g2, pt10g3 and pt10g5 for the emulsion with reproducibility was due to two reasons. First, at each startup
8% (w/w) water content. of the experimental apparatus, the generated emulsion was
The normalized droplet volume size distributions mea- somewhat different, even though all controlled variables (tem-
sured for these runs are given in Fig. 3. The relative frequency perature, water concentration and mass flow rates) were
of each size class was plotted against the size of the droplet basically equal. It was then verified that the physical prop-
with the mean volume in the corresponding class. From, erties of the stored oil suffered 5–10% changes due to aging.
this figure, it can be seen that the method for sampling and This was partially taken into account in the data analysis by
measurement of the droplet size distribution is reproducible
within 10–25% of the relative frequency of each size class. d (μm)
Fig. 3(b) shows that this agreement is better for the 8% (w/w) 101 102
concentration of water, being in the error range of the dis- 25.0 Experimental data
pt1g11 pt3g13
tribution determination. Therefore, the effects of coalescence pt1g12 pt3g14
20.0
and breakage were negligible for this case. On the other hand, pt1g15 pt3g15
Nout × 10-11 ( - )

pt2g11 pt4g11
Fig. 3(a) clearly shows that the inlet and outlet distributions 15.0 pt2g12 pt4g13
pt2g13 pt4g14
for the 15% (w/w) water-in-oil emulsion were somewhat differ- pt2g14 pt5g11
ent, especially for the smallest droplets. Although this could 10.0 pt2g15 pt5g12
pt3g11 pt5g14
be attributed to some degree of coalescence, the most proba- 5.0
pt3g12 pt5g15
ble cause is the inadequate sampling of the smallest droplets,
as discussed in the following. 0.0

10-16 10-15 10-14 10-13 10-12 10-11 10-10


5.3. Preliminary analysis of the experimental size 3
v (m )
distributions
Fig. 4 – Experimental droplet number size distributions at
Fig. 4 shows the droplet number size distributions at the the inlet of the test section for W = 3.0 kg/min and  = 15%
inlet of the test section for all experimental runs with water (w/w) conditions.
2504 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

d (μm) Table 4 – Model parameter estimates obtained by


100 101 102 ODRPACK95 considering (S2 and S3) or not (S1) the
6.0 6.0
pt2g4 inlet (exp) experimental errors in the explanatory variables.
5.0 outlet (exp) 5.0 Results S3 were obtained using the estimates for the
Nout × 10−14 ( − )

Nin × 10−12 ( − )
4.0 η dc,mb 4.0 effective volume of the valve-like element given by the
AIU-PPSO algorithm (Moraes, 2011).
3.0 3.0
Parameters S1 S2 S3
2.0 2.0
1.0 Model with constant coalescence efficiency
1.0
102 Cc 1.00 ± 0.05 1.05 ± 0.23 1.88 ± 0.06
0.0
0.0 102 Cb 0.98 ± 0.06 1.06 ± 0.80 1.08 ± 0.07
ς 26.0 ± 0.9 26.9 ± 9.4 32.7 ± 16.8
10−19 10−17 10−15 10−13
3
v (m ) Coalescence efficiency model given by Eq. (18)
102 Cc 0.90 ± 0.47 – –
Fig. 5 – Experimental droplet number size distributions at 102 Ce 1 (constant) – –
the inlet and outlet of the test section. 102 Cb 0.81 ± 0.31 – –
ς 31.2 ± 9.5 – –

determining the oil physical properties again after the g10 modified to consider just this size range. The model for droplet
group of experimental runs. collisions became:
Second, the experimental procedure used to sample and
3
(d + d ) 
measure the droplet size distributions may introduce some (d, d ) = Cc (40)
perturbations due to micro-bubbles entrainment and insuffi- 8
cient sampling of the smallest droplet sizes. This is the most
where Cc is an experimentally adjusted parameter and the
probable reason why the number size distributions did not
coalescence efficiency was assumed constant or given by Eq.
reproduce well (see Fig. 4). On the other hand, as the vol-
(18), which introduces the adjusted parameter Ce . The break-
ume size distributions depend only slightly on the smallest
age frequency model was simplified to:
droplets, they could be obtained with fairly good reproducibil-
ity, as shown in Fig. 3. This occurs because the number of
⎧  4/5

⎨ Cb 63.927  Ca2.2 d
droplets needed for the convergence of the number size distri- 11/5
, for Ca > Cacrit
b(d) = We 2 (41)
bution is much larger than what is necessary for the volume ⎪ crit
⎩ 0, for Ca ≤ Ca
size distribution convergence. Therefore, droplet number size crit
distributions should not be used for estimating the model
11/5
parameters, as they are mostly influenced by the smallest where Wecrit = 6 was used leading to 63.927/Wecrit = 1.241.
droplet sizes. However, we decided to analyze both the num- Although this is a typical value for bubbles and not for droplets
ber and volume size distributions. In order to do that, the runs (Galinat et al., 2005; Raikar et al., 2010; Maaß et al., 2010), this
in the same operational conditions could not be used as actual has no consequence because the adjusted value of Cb corrects
replicates. Therefore, each experimental run was considered for any inadequacy. The critical capillary number is still given
unique and the experimental accuracy of 10% was assumed by Eq. (27) and the daughter droplet probability distribution is
for Nk and Nv,k . given by Eq. (28).
Fig. 5 shows the particle size distributions at the inlet and
outlet of the test section for run pt2g4. The calculated values 5.4. Parameter estimation with ODRPACK
of the Kolmogorov length scale, , and the critical diameter
in the Martínez-Bazán et al. (1999a) model, dc,mb , were also The parameter estimation procedure described in Section 4.3
represented for the conditions of this run. dc,mb was evalu- was applied to the population balance models given by Eqs. (8),
ated using Eq. (20) with Wecrit = 6. Fig. 5 shows the occurrence (40), (41), (27) and (28) for the model with constant coalescence
of strong droplet breakage as the emulsion flows through the efficiency and also using Eq. (18) for the coalescence effi-
test section. It also clearly shows that most of the particles ciency model for drops with deformable and partially mobile
at the inlet of the test section are smaller than  and all of interfaces. The discrete particle volume distribution functions
them are smaller than dc,mb calculated using Wecrit = 6. For this (Nv,k values) at the test section inlet and outlet were used
value of dc,mb , the Martínez-Bazán et al. (1999a) model predicts as explanatory and dependent variables, respectively. The
a zero breakage frequency, as given by Eq. (19). Therefore, Wecrit alternative use of the discrete particle number distribution
was included as a new parameter to be estimated in the pop- functions (Nk values) did not result in any set of parameter
ulation balance model using the breakage model described in values with statistical significance. This was expected due to
Section 3.3. the errors associated to the measurement of the number dis-
However, in any optimization run, the estimate of Wecrit tribution functions.
always corresponded to a dl value larger than the largest Table 4 shows the optimized parameter values and their
droplet in the experimental PSDFs. From Eq. (24), it is clear 95% confidence intervals obtained using all the valid experi-
that only the part of the proposed breakage model based on mental data, considering or not the errors in the explanatory
the work of Cristini et al. (2003) was active. This was expected variables. For the constant coalescence efficiency model, the
because the analysis of the flow turbulence indicates that values of the parameters are almost the same for both
there is no distinct inertial subrange. optimizations, but the confidence intervals are much larger
Therefore, we could conclude that all droplets in our when the errors in the explanatory variables were taken into
experimental conditions were in the turbulence dissipation account. Considering the coalescence efficiency model for
size range and the coalescence and breakage models were droplets with deformable and partially mobile interfaces, it
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2505

was not possible to determine all the model parameters with


Table 5 – Errors in the prediction of mean diameters at
statistical significance. This is related to the global character- the outlet of the test section using the model with
istics of the employed population balance model. However, constant coalescence efficiency.
for PB-CFD simulations, it is very important to have a non-
pq Dpq dpq
constant coalescence efficiency. Therefore, the Ce parameter
was fixed at several values and an attempt was made to e (%) |e| (%) emax (%) e (%) |e| (%) emax (%)
estimate the other parameters. This procedure was success- 10 −0.01 8.1 25.6 189 189 780
ful only for Ce = 0.01 and when the errors in the explanatory 20 −1.69 9.3 29.6 142 142 546
variables were not considered. The results show in Table 4 30 −1.95 10.3 31.4 88 88 288
indicates that the values for the other parameters, Cc , Cb and 32 −2.37 12.6 35.3 21.3 21.3 50.6
ς, agree with those obtained for the constant coalescence effi- 43 −0.03 14.0 39.5 −0.01 8.1 25.6
ciency model, within their confidence intervals.
The large value of the number of daughter particles may peak, as shown in Fig. 6(b), which is not important to the vol-
be substantiated by other works (Eastwood et al., 2004; Galinat ume particle distribution but is very important to the number
et al., 2005; Andersson and Andersson, 2006; Raikar et al., 2010; particle distribution. Nonetheless, the mean diameter d43 is
Maaß et al., 2010), but it is important to note that the model the most important one for analyzing the desalting equip-
used for the daughter conditional probability distribution is ment and it was very well predicted. The mean diameter d32
the simplest possible. is important in polydisperse multiphase flow simulations. As
Due to space limitations, it is not possible to present all shown in Table 5, it could be predicted with a mean error of
results here. The comparison of simulated and experimental 21.3%, which is reasonably good.
data for all runs are available in Mitre (2010). Some results are
presented below. 5.5. Results from the global optimization
Fig. 6 shows the droplet volume size distributions at the
inlet and outlet of the test section obtained experimentally The global optimization AIU-PPSO algorithm (Moraes, 2011)
and by simulation using the population balance model with was applied to the estimation of Cc , Cb , ς and the Ve,i , i = 1, . . . M
constant coalescence efficiency. These results were calculated for the population balance model with the constant coales-
using an unique set of estimated values for the parameters of cence efficiency. Since M is large, the global optimization in a
the breakage and coalescence models using all 78 experimen- space with M + 3 dimensions has a very large computational
tal runs (set S1 of Table 4). Besides, it should be pointed out cost, whose order is about 3 months of CPU time in a 64-node
that, when errors are considered in the Nv,in variables, the inlet cluster. Therefore, the global optimization of the population
distribution obtained by the optimized model can be different balance model with the other coalescence efficiency model
from the experimentally measured one. These simulated inlet was not carried out.
distributions are also given in Fig. 6 where it can be seen that After the global minimization, the parameters obtained by
they agree to the experimental ones within the experimental the AIU-PPSO algorithm were fed to the ODRPACK optimiza-
errors. Fig. 6(a) presents the results with the best agreement, tion program, including the estimated values of Ve,i , i = 1, . . . M
where the droplet volume size distribution at the outlet is uni- and considering the errors of all explanatory variables. The
modal. Fig. 6(b) shows typical results where the outlet droplet resulting parameter values and their 95% confidence inter-
volume size distribution is bimodal. Despite the good agree- vals are also shown in Table 4. This procedure guaranteed that
ment obtained in the main peak, the secondary peak could the same objective function was used to generate these and
not be predicted by the model. This part of the curve is not the previous results (S2). Therefore, we can compare directly
relevant to the drop volume distribution, but is the most rel- the parameter values, their 95% confidence intervals and the
evant part for the corresponding drop number distribution. objective function value.
The inlet drop volume distribution for case pt7g2, shown in It is clear from Table 4 that the confidence intervals for Cc
Fig. 6(c), has an unusual secondary peak. But it had little influ- and Cb are much smaller when the optimized values for Ve,i ,
ence in the droplet volume size distribution predicted at the i = 1, . . . M were used. Besides, it can be seen that the values
outlet. Fig. 6(d) shows the worst result for which the main peak of Cb and ς agree within their confidence intervals, while the
of the outlet size distribution was not predicted well. Almost Cc values do not, even though they have the same magnitude.
identical results were obtained using the 4-parameter model, This result can be explained as follows. The model assumption
with a non-constant coalescence efficiency. These results are that breakage and coalescence occur in the same region, with
not shown here. volume Ve , about the valve-like element is good for breakage,
Mean diameters are usually needed for equipment design but not so good for coalescence. As high velocity collisions lead
and to couple the population balance solution with CFD sim- to bouncing, the coalescence should occur mostly after the
ulation. They can be calculated using the particle number size valve-like element in a larger region than what was assumed
distribution, dp,q , or by the particle volume size distribution, for breakage. Therefore, the optimization of the Ve,i values for
Dp,q . Since F(v) = vf (v), we have Dp,q = dp+3,q+3 . Table 5 shows all experiments affected more the Cc value than the Cb value.
the errors for the most common mean diameters. Regarding By comparing the values of the objective function, Eq. (37),
the mean diameters based on the particle volume distribu- it could be concluded that the use of the adjusted Ve,i val-
tion, Dpq , it can be concluded that the model is not biased ues produced simulated droplet volume size distributions that
because the values of e are almost zero. Besides, the mean were approximately 20% better than those obtained with the
relative errors are smaller than 15% for all mean diameters. empirical rule used to fix these values. Besides, the mean of
On the other hand, the mean diameters based on the droplet the magnitude of the relative errors between the modeled and
number distribution, dpq , were not well determined except for estimated values of Ve,i was approximately 23%, whereas the
d4,3 = D1,0 . As the parameters have been estimated using the mean of the relative errors was just 8%. This indicates that
values of Nv,k , the model fails to predict the small secondary the empirical rule to fix the Ve,i values has some error, but
2506 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

d (μm) d (μm)
(a) (b)
101 102 100 101 102
15.0
8.0 inlet (exp) inlet (sim) 8.0 inlet (exp) inlet (sim)
outlet (exp) outlet (sim) outlet (exp) outlet (sim) 15.0

Nv,in × 103 ( − )
Nv,in × 10 ( − )
Nv,out × 10 ( − )

Nv,out × 10 ( − )
6.0 6.0 10.0
pt6g7 pt4g13 10.0

3
3

3
4.0 4.0
5.0
2.0 5.0
2.0

0.0 0.0
0.0 0.0

−17 −15 −13 −11 −19 −17 −15 −13 −11


10 10 10 10 10 10 10 10 10
v (m3) v (m3)

d (μm) d (μm)
(c) 100 101 102 (d) 100 101 102
15.0
8.0 inlet (exp) inlet (sim) inlet (exp) inlet (sim)
outlet (exp) outlet (sim) 8.0 outlet (exp) outlet (sim) 15.0
Nv,out × 10 ( − )

Nv,out × 10 ( − )

Nv,in × 103 ( − )
Nv,in × 10 ( − )
6.0 10.0
pt7g2 6.0 pt2g4 10.0
3

3
3

4.0
4.0
5.0
5.0
2.0
2.0

0.0 0.0 0.0


0.0
−19 −17 −15 −13 −19 −17 −15 −13 −11
10 10 10 10 10 10 10 10 10
3 3
v (m ) v (m )

Fig. 6 – Simulated and experimental droplet volume size distributions using the constant coalescence efficiency model: (a)
best, (b) typical, (c) unusual and (d) worst results.

it is almost unbiased. Besides, no trend could be perceived


between the Ve,i errors and the experimental operational con- d (μm)
ditions. For instance, the mean values of the magnitude of the (a) 101 102
30.0
relative errors between the modeled and estimated values of inlet (exp) outlet (sim) 14.0
25.0 outlet (exp)
Ve,i were 25%, 28% and 16% for the smallest, intermediate and 12.0
largest emulsion flow rates, and the corresponding values of
Nv,out × 10 ( − )

Nv,in × 10 ( − )
20.0 pt4g11 10.0
the mean relative errors were −10%, 19% and 10%, respectively.
3

8.0

3
15.0
6.0
5.6. Results for other models 10.0
4.0
5.0 2.0
Several combinations of coalescence and breakage models
0.0 0.0
were evaluated during this study. More than 40 combina-
−18 −16 −14 −12 −10
tions of models were tested using the coalescence models 10 10 10 10 10
3
of Chesters (1991), the breakage models of Martínez-Bazán v (m )
et al. (1999a) and that derived from Cristini et al. (2003)
work with and without some modifications. These models
d (μm)
were employed with parameters multiplying the breakage (b) 1 2
10 10
and coalescence frequencies, which were estimated using the
inlet (exp) inlet (sim)
same procedure described above for each one of the combi- outlet (exp) outlet (sim) 20.0
15.0
nations. All combinations of models that were tested can be
Nv,in × 103 ( − )
Nv,out × 10 ( − )

found in Mitre (2010). pt4g8 15.0


Most of the combinations of models could not represent 10.0
3

the experimental data nor gave parameters with statistical 10.0


significance. For example, when using the original equations 5.0
of Cristini et al. (2003), Eq. (21), the 95% confidence level of 5.0
Cb,c was almost twice the value of the parameter itself. Bad 0.0
results were also obtained when trying other models for the 0.0
−18 −16 −14 −12 −10
critical capillary number, even when considering the model for 10 10 10 10 10
3
breakage frequency developed in the present work, Eq. (41). v (m )
Regarding the probability distribution of daughter particles,
Fig. 7 – Simulated and experimental droplet volume size
the models of binary breakage with equal size daughters or an
distributions for binary breakage with (a) daughters of
uniform distribution of daughter sizes were also considered as
equal size and (b) uniform distribution of daughter sizes.
well as the models of Martínez-Bazán et al. (1999a), Hsia and
chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508 2507

Tavlarides (1983) and Hesketh et al. (1991). A ternary breakage Chesters, A.K., 1991. The modelling of coalescence processes in
model with uniform distribution of daughter sizes was devel- fluid–liquid dispersions: a review of current understanding.
oped and analyzed. In all cases, the model parameters were Trans. IChemE 69 (Pt A), 259–270.
Coulaloglou, C.A., Tavlarides, L.L., 1977. Description of interaction
estimated again.
processes in agitated liquid–liquid dispersions. Chem. Eng.
The results considering the simple models for the prob- Sci. 32, 1289–1297.
ability distribution of daughter particles assuming binary Cristini, V., Blawzdziewicz, J., Loewenberg, M., Collins, L., 2003.
breakage are shown in Fig. 7. The employed breakage and Breakup in stochastic Stokes flows: sub-Kolmogorov drops in
coalescence models are those described in Section 5.3 using isotropic turbulence. J. Fluid Mech. 492, 231–250.
a constant coalescence efficiency. Fig. 7(a) shows a result for Cunha, P.S.M.D., 2008. Mathematical Modeling of the Crude Oil
Electrostatic Dehydration Process. Universidade Federal do
assuming binary breakage with daughters particles of equal
Rio de Janeiro, EQ, RJ, Brazil (Master’s thesis), (in Portuguese).
size whereas Fig. 7(b) shows a result for uniform probabil-
Eastwood, C.D., Armi, L., Lasheras, J.C., 2004. Breakup of
ity distribution of daughter droplets. It can be noted that the immiscible fluids in turbulent flows. J. Fluid Mech. 502,
simulated distributions did not agree with the experimental 309–333.
data. Eow, J.S., Ghadiri, M., 2002. Electrostatic enhancement of
coalescence of water droplets in oil: a review of the
technology. Chem. Eng. J. 85, 357–368.
6. Conclusions Frank, T., Zwart, P.J., Krepper, E., Prasser, H.-M., Lucas, D., 2007.
Validation of CFD models for mono- and polydisperse
This work presented an experimental and theoretical study air–water two-phase flows in pipes. Nucl. Eng. Des. 238,
on the evolution of droplet size distribution in the flow of 647–659.
water in oil emulsions through valve-like elements. A sim- Friedlander, S.K., 1977. Smoke, Dust and Haze. Wiley, New York.
Galinat, S., Masbernat, O., Guiraud, P., Dalmazzone, C., Noïk, C.,
plified population balance model was developed to interpret
2005. Drop break-up in turbulent pipe flow downstream of a
the experimental data.
restriction. Chem. Eng. Sci. 60, 6511–6528.
Breakage and coalescence models that could be applied to Galinat, S., Torres, L.G., Masbernat, O., Guiraud, P., Risso, F.,
the entire range of droplet sizes were proposed. However, the Dalmazzone, C., Noïk, C., 2007. Breakup of a drop in a
experimental data had the majority of their droplets in the tur- liquid–liquid pipe flow through an orifice. AIChE J. 53, 56–68.
bulence dissipation size range and, therefore, only the models Hagesaether, L., 2002. Coalescence and Break-up of Drops and
applicable for this size range were used to correlate the droplet Bubbles. Norwegian University of Science and Technology,
Trondheim, Noruega (Ph. D. thesis).
volume size distribution data. The models have 3 or 4 parame-
Hagesaether, L., 2002b. A model for turbulent binary breakup of
ters depending if a constant coalescence efficiency is assumed
dispersed fluid particles. Chem. Eng. Sci. 57, 3251–3267.
or not. Thus, caution is recommended for the application of Hesketh, R.P., Etchells, A.W., Fraser Russell, T.W., 1991.
these models for droplet sizes much larger than the turbulent Experimental observations of bubble breakage in turbulent
dissipation size range. flow. Ind. Eng. Chem. Res. 30, 835–841.
Statistically significant parameter values could be esti- Hinze, J.O., 1955. Fundamentals of the hydrodynamic mechanism
mated by local and global optimization methods. The of splitting in dispersions processes. AIChE J. 1, 289–295.
Hsia, M.A., Tavlarides, L.L., 1983. Simulation analysis of drop
population balance model using the simplified breakage and
breakage, coalescence and micromixing in liquid–liquid
coalescence models generated unbiased results for the mean stirred tanks. Chem. Eng. J. 26, 189–199.
droplet diameters based on the particle volume size distribu- Israelachvili, J.N., 2011. Intermolecular and Surface Forces, 3rd ed.
tion that show very good agreement with the experimental Academic Press, San Diego.
data. The model also gave good results for the particle volume Jakobsen, H.L.H., Dorao, C.A., 2005. Modeling of bubble column
size distributions at the outlet of the test section. reactors: progress and limitations. Ind. Eng. Chem. Res. 44,
5107–5151.
Joshi, J.B., 2001. Computational flow modelling and design of
Acknowledgements bubble column reactors. Chem. Eng. Sci. 56, 5893–5933.
Kamp, A.M., Chesters, A.K., Colin, C., Fabre, J., 2001. Bubble
The authors would like to acknowledge the financial sup- coalescence in turbulent flows: a mechanistic model for
port from PETROBRAS, CNPq (Grants nos. 485817/2007-1, turbulence-induced coalescence applied to microgravity
301672/2008-3, 140883/2006-1, 100253/2011-3) and ESSS. bubbly pipe flow. Int. J. Multiphase Flow 27, 1363–1396.
Keleşǒglu, S., Pettersen, B., Sjöblom, J., 2013. Flow properties of
water-in-north sea heavy crude oil emulsions. J. Petrol. Sci.
References Eng. 100, 14–23.
Kim, W.K., Lee, K.L., 1987. Coalescence behavior of two bubbles in
Al-Sabagh, A.M., Kandile, N.G., Noor El-Din, M.R., 2011. Functions stagnant liquids. J. Chem. Eng. Jpn. 20, 448–453.
of demulsifiers in the petroleum industry. Sep. Sci. Technol. Kumar, S., Ramkrishna, D., 1996. On the solution of population
46, 1144–1163. balance equations by discretization – I. A fixed pivot
Andersson, R., Andersson, B., 2006. Modeling the breakup of fluid technique. Chem. Eng. Sci. 51, 1311–1332.
particles in turbulent flows. AIChE J. 52, 2031–2038. Lage, P.L.C., Rodrigues, R.C., Mitre, J.F., Souza, M.N.D., Barca, L.F.,
Andersson, R., Andersson, B., 2006. On breakup of fluid particles Souza, M.A., Silva, E., Varella, S., Coutinho, R.C.C., 2006.
in turbulent flows. AIChE J. 52, 2020–2030. Droplet evolution in the flow of water in oil emulsions
Boggs, P.T., Byrd, R.H., Schnabel, R.B., 1987. A stable and efficient through duct accidents – analysis of droplet breakage. In:
algorithm for nonlinear orthogonal distance regression. J. Sci. Proceedings of the 2006 CFDOIL, pp. 1–41.
Stat. Comput. 8, 1052–1078. Lasheras, J.C., Eastwood, C., Martínez-Bazán, C., Montañés, J.L.,
Boggs, P.T., Byrd, R.H., Donaldson, J.R., Schnabel, R.B., 1989. 2002. A review of statistical models for the break-up of an
Algorithm 676 – ODRPACK: software for weighted orthogonal immiscible fluid immersed into a fully developed turbulent
distance regression. ACM Trans. Math. Softw. 15, 348–364. flow. Int. J. Multiphase Flow 28, 247–278.
Chen, Z., Prüss, J., Warnecke, H.-J., 1998. A population balance Liao, Y., Lucas, D., 2009. A literature review of theoretical models
model for disperse systems: drop size distribution in for drop and bubble breakup in turbulent dispersions. Chem.
emulsion. Chem. Eng. Sci. 53, 1059–1066. Eng. Sci. 64, 3389–3406.
2508 chemical engineering research and design 9 2 ( 2 0 1 4 ) 2493–2508

Liao, Y., Lucas, D., 2010. A literature review on mechanisms and Prince, M.J., Blanch, H.W., 1990. Bubble coalescence and breakup
models for the coalescence process of fluid particles. Chem. in air-sparged bubble columns. AIChE J. 36, 1485–1499.
Eng. Sci. 65, 2851–2864. Raghavan, R., Marsden Jr., S.S., Stanford, U., 1971. Theoretical
Luo, H., Svendsen, H.F., 1996. Theoretical model for drop and aspects of emulsification in porous media. Soc. Petrol. Eng. 11,
bubble breakup in turbulent dispersions. AIChE J. 42, 153–161.
1225–1233. Raikar, N.B., Bhatia, S.R., Malone, M.F., Henson, M.A., 2009.
Maaß, S., Hermann, S., Kraume, M., 2010. Determination of Experimental studies and population balance equation
breakage rates with single drop experiments. In: 4th models for breakage prediction of emulsion drop sizes
International Conference on Population Balance Modelling, distributions. Chem. Eng. Sci. 64, 2433–2447.
Berlin, Germany. Raikar, N.B., Bhatia, S.R., Malone, M.F., McClements, D.J.,
Maindarkar, S., Raikar, N., Bongers, P., Henson, M., 2012. Almeida-Rivera, C., Bongers, P., Henson, M.A., 2010. Prediction
Incorporating emulsion drop coalescence into population of emulsion drop size distributions with population balance
balance equation models of high pressure homogenization. equation models of multiple drop breakage. Colloids Surf. A:
Colloids Surf. A: Physicochem. Eng. Aspects 396, 63–73. Physicochem. Eng. Aspects 361, 96–108.
Maniero, R., Masbernat, O., Climent, E., Risso, F., 2012. Modeling Raikar, N., Bhatia, S., Malone, M., McClements, D., Henson, M.,
and simulation of inertial drop break-up in a turbulent pipe 2011. Predicting the effect of the homogenization pressure on
flow downstream of a restriction. Int. J. Multiphase Flow 42, emulsion drop-size distributions. Ind. Eng. Chem. Res. 50,
1–8. 6089–6100.
Manning, F.S., Thompson, R.E., 1995. Oilfield Processing, Crude Ramkrishna, D., 2000. Population Balances – Theory and
Oil, vol. 2. Pennwell Books, Tulsa, Oklahoma. Applications to Particulate Systems in Engineering. Academic
Marchisio, D.L., Fox, R.O., 2005. Solution of population balance Press, New York.
equations using the direct quadrature method of moments. Ross, S.L., Verhoff, F.H., Curl, R.L., 1978. Droplet breakage and
Aeros. Sci. 36, 43–73. coalescence processes in an agitated dispersion. 2.
Marchisio, D.L., Vigil, R.D., Fox, R.O., 2003. Implementation of the measurement and interpretation of mixing experiments. Ind.
quadrature method of moments in CFD codes for Eng. Chem. Fundam. 17, 101–108.
aggregation-breakage problems. Chem. Eng. Sci. 58, Silva, L.F.L.R., Lage, P.L.C., 2011. Development and
3337–3351. implementation of a polydispersed multiphase flow model in
Marrucci, G., 1969. A theory of coalescence. Chem. Eng. Sci. 24, OpenFOAM. Comput. Chem. Eng. 35, 2653–2666.
975–985. Silva, L.F.L.R., Damian, R.B., Lage, P.L.C., 2008. Implementation
Martínez-Bazán, C., Montañés, J.L., Lasheras, J.C., 1999a. On the and analysis of numerical solution of the population balance
breakup of an air bubble injected into a fully developed equation in CFD packages. Comput. Chem. Eng. 32,
turbulent flow. Part 1. Breakup frequency. J. Fluid Mech. 401, 2933–2945.
157–182. Stewart, M., Arnold, K.E, 2008. Surface Production Operations,
Martínez-Bazán, C., Montañés, J.L., Lasheras, J.C., 1999b. On the Design of Oil Handling Systems and Facilities, vol. 1, 3rd ed.
breakup of an air bubble injected into a fully developed Gulf Professional Publishing, Oxford.
turbulent flow. Part 2. Size PDF of the resulting daughter Stewart, M., Arnold, K., 2009. Emulsions and Oil Treating
bubbles. J. Fluid Mech. 401, 183–207. Equipment: Selection. Sizing and Troubleshooting, Gulf
Mitre, J.F., Takahashi, C.P., Ribeiro Jr., C.P., Lage, P.L.C., 2010. Professional Publishing, Oxford.
Analysis of breakage and coalescence models for bubble Tsouris, C., Tavlarides, L.L., 1994. Breakage and coalescence
columns. Chem. Eng. Sci. 65, 6089–6100. models for drops in turbulent dispersions. AIChE J. 40,
Mitre, J.F., 2010. Breakage and Coalescence Models for Drops in 395–406.
Emulsion Flows. Universidade Federal do Rio de Janeiro, Vankova, N., Tcholakova, S., Denkov, N., Ivanov, I., Vulchev, V.,
PEQ/COPPE (Ph.D. thesis) (in Portuguese). Danner, T., 2007. Emulsification in turbulent flow. 1. Mean and
Moraes, A.O.S., 2011. Development and Implementation of maximum drop diameters in inertial and viscous regimes. J.
Parallel Versions of the Particle Swarm Algorithm on Clusters Colloid Interface Sci. 312, 363–380.
Using MPI. Universidade Federal do Rio de Janeiro, PEQ/COPPE, Wang, T., Wang, J., Jin, Y., 2003. A novel theoretical breakup kernel
RJ, Brazil (Master’s thesis) (in Portuguese). function for bubbles/droplets in a turbulent flow. Chem. Eng.
Noïk, C., Jiaqing, C., Dalmazzone, C., 2006. Electrostatic Sci. 58, 4629–4637.
Demulsification on Crude Oil: A State-of-the-art Review. In: Zwokak, J.W., Boggs, P.T., Watson, L.T., 2004. ODRPACK95: A
International Oil and Gas Conference and Exhibition, vol. 1, Weighted Orthogonal Distance Regression Code with Bound
5-7 December, Beijing, China, pp. 410–421. Constraints. Department of Computer Science, Virginia
Paiva, A.P., 2004. Estudo da minimização de erro das medições de Polytechnic Institute and State University, Blacksburg, VA,
concentração de emulsões por titração Karl–Fisher USA.
utilizando-se projetos de experimentos. Universidade Federal Zwolak, J.W., Boggs, P.T., Watson, L.T., 2007. Algorithm 869 –
de Itajubá, RJ, Brazil (Master’s thesis). odrpack95: a weighted orthogonal distance regression code
Pope, S.B., 2000. Turbulent Flows. Cambridge University Press, with bound constrains. ACM Trans. Math. Softw. 33,
Cambridge. 348–364.

You might also like