You are on page 1of 8

Rheological Behaviour of Oil and Water Emulsions and their Flow

Characterization in Horizontal Pipes

Jian Zhang and Jing-Yu Xu*


LMFS, Institute of Mechanics, Chinese Academy of Sciences, Beijing, 100190, China

In this work, the rheological behaviour of oil and water emulsions and their flow characterization were studied using a Haake RS6000 rheometer and
a lab scale flow loop. The rheological properties of the synthetic emulsions were investigated at oil volume fractions from 0.1–1.0 L/L and shear rates
from 0.001–1000 s1 at a system temperature of 20 8C. The emulsions flow, formed by the SMV static mixer installed before the test section, was
observed in the 2-m long horizontal pipes with 25 and 50 mm inner diameters at different mixture flow rates from 0.01–7.00 m3/h. A comparison
between rheological measurements and laminar flow data shows that the values of the Sauter mean droplet diameters formed by the SMV static
mixer are approximately one order of magnitude bigger than those measured by the rheometer. For oil-in-water emulsions with low oil content, the
stress-strain relationships obtained in the rheometer are more suitable for predicting the transport characteristics of the emulsions with a large pipe
diameter than those with a small pipe diameter. In addition, the phase inversion points in the pipe flow are closer to those measured at high shear
rates in the rheometer. Therefore, the apparent viscosity obtained by the rheology measurements at high shear rates could be introduced for
accurate calculations of the mixture Reynolds number and the friction factor in emulsions flow.

Keywords: emulsions, rheology, flow behaviour, droplet size, horizontal pipes

INTRODUCTION emulsion flow characteristics in large-diameter pipes. KeleSs oglu


et al.[17,18] investigated the rheological behaviour and pipe flow
n the oil industry, emulsion flow in pipes is very common,

I such as for oil-producing wells and crude oil transportation.


Typically, emulsions of oil and water can be classified into two
types: oil droplets as a dispersed phase in a continuous water
properties of water-in-oil emulsions using a stress-controlled
rheometer and an indoor flow loop, respectively. The apparent
viscosities of pipe flow emulsions calculated by the frictional
pressure drop were compared with those from the rheology
(o/w), and water droplets as a dispersed phase in a continuous oil measurement. The experimental data demonstrated that the
(w/o). In addition, there may also be an ambiguous region, where calculated apparent viscosities formed during the pipe flow
the dispersed phase becomes the continuous phase and vice were lower than those obtained by the rheometer. This
versa.[1] Generally, the rheological behaviour of emulsions is phenomenon was attributed to the unequal-size droplets and
determined by various interaction forces occurring in the system, polydispersity of both systems and a non-fully dispersed flow of
which in turn depend on the chemical composition of the the fluid mixtures in pipes.
continuous phase, dispersed phase holdup, interfacial tension of In practice, the emulsions flow of oil and water in pipes is
oil and water, and emulsifier properties. Other structural always unsteady and coarse. The system holds a fine stable
parameters important to the emulsion rheological characteristics structure only when the droplets are covered with a surfactant or
include the physical properties of the continuous phase, droplet emulsifier.[16] However, most of the published literature focuses
size, and droplet size distribution. However, until now these on a steady water-in-oil dispersed flow with an emulsifier.[19]
impacts on the transportation of oil and water emulsions have In these studies, the apparent viscosity is back-calculated from
not been studied thoroughly, and the rheology measurements the measured frictional pressure drop, which is dominated by
obtained in the rheometer for determining and calculating the hydrodynamics. In this way, the apparent viscosity from the
rheological properties of oil and water emulsions have not been rheology measurement introduced into the pipe flow conditions is
optimized for simulating the mixture flow in pipes, especially for difficult to use to represent the complex flow behaviour and
oil-in-water emulsions.[2,3] physical phenomena.[13] Therefore, to comprehensively under-
Accurate assessments of rheological relation, apparent viscos- stand the emulsion flow characteristics without an emulsifier and
ity, and pressure drop of the emulsion flow in pipes are very the mechanisms behind it, we start the present study as follows.
important for the operation of oil-producing wells, including flow Firstly, the rheological behaviour of synthetic emulsions was
assurance, wellbore modelling, and pipeline transportation of measured at a full range of oil volume fractions using a rheometer.
crude oil. To solve these problems, considerable theoretical and Secondly, the flow characterization of emulsions in pipes was
experimental studies have been carried out using two different
approaches: synthetic emulsions measured in a rheometer[3–7] and
emulsions induced in pipe flow.[8–15] However, a few works have
been carried out by simultaneously using two systems.[16–19] * Author to whom correspondence should be addressed.
Masalova et al.[16] studied the rheological and transport properties E-mail address: xujingyu@imech.ac.cn
of water-in-oil emulsions. A relationship between the estimation Can. J. Chem. Eng. 94:324–331, 2016
of rheology measurements and the predictions calculated with © 2015 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.22377
direct pipe flow data was presented. Results showed that Published online 8 January 2016 in Wiley Online Library
experimental data obtained in the rheometer could predict (wileyonlinelibrary.com).

324 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 94, FEBRUARY 2016
investigated in 2-m long horizontal pipes with 25 and 50 mm inner the CS model was defined as a controlled stress input and
diameters, respectively. Finally, a comparison of constitutive determining the resulting shear rate. In addition, the rheometer
relation, phase inversion, and apparent viscosity measured in a provided a shear rate range from 0.001–1500 s1 and a viscosity
laminar flow with results obtained from the rheometer was range from 0.001–1000 Pa  s.
performed. Reasonable calculation methods were suggested to According to previous work,[3] the rheological characteristics
solve the apparent viscosity and frictional pressure drop in of the oil and water emulsions can be studied by fitting the
emulsions flow. experimental data to the Ostwald-de Waele model (Power-law
model). This model is expressed by using two parameters (m and
n) in the following relationship:
EXPERIMENTAL
t ¼ mð: g Þn ð1Þ
Materials
In this work, tap water and industry white oil PS-40, manufactured where t and : g refer to the shear stress in Pa and the corresponding
by Yanshan Petrochemical Company in China, were used in shear rate in s1, respectively. m is the fluid consistency coefficient
preparing emulsions. The white oil is a refined mineral oil that in Pa . sn, and n the flow behaviour index.
consists of saturated hydrocarbons, and can be classified
according to its viscosity. At 20 8C, the density of white oil Experimental Setup and Procedure of Emulsion Flow in Horizontal
is 840 kg/m3 and the interfacial tension of oil and water is Pipes
0.0313 N/m. All experiments were conducted at 20 8C and The experimental setup used in emulsions pipe flow is shown in
atmospheric outlet pressure. Figure 1. The system consisted of a steel frame supporting
Perspex tubes. The flow loop was built up to perform a steady
Rheological Measurements flow of the fluid mixtures at a fixed temperature and to
Samples of synthetic emulsions were prepared through the entire simultaneously measure the flow rate, holdup, pressure drop,
range of volume fractions of oil phase from 0.1–1.0 L/L. The speed and system temperature. Experiments were carried out in two
of a homogenizer with a three-blade stirrer was kept at 1000 rpm, pipes with 25 and 50 mm inner diameters to establish the
and shearing was maintained at 300 or 600 s. After homogeniza- relationship between flow rate and pressure drop. The oil and
tion, by exploiting the performance of the rheometer, the apparent water phases were fed into the flow loop from their respective
viscosity and rheological characterization of emulsions were storage tanks. The volumetric flow rate of each phase was
measured. Additionally, a microscope (Olympus BX43) and CCD regulated independently and measured by two turbine flow-
imaging system (ProgRes C5 cool) were used to take photomicro- meters. The emulsion was formed through the static mixer
graphs of samples. Based on these photos, the droplet size and elements and then introduced into the 2-m long test sections,
droplet size distribution were calculated. where all parameters were measured. After the test sections, the
Rheological measurements were carried out on a Haake RS6000 emulsions returned to the storage tanks, from where they were
rheometer, which was equipped with a coaxial cylinder sensor pumped again. Several absolute pressure transducers were used
system having the following characteristics: Z38 DIN rotor, to measure the pressure drop in the test sections and the static
2.5 mm gap width, and 30.8 cm3 sample volume. A liquid mixers. The precise values were calculated by the difference
temperature-controlled system could make the sensor system between two transducers, installed at both ends of a 1.0-m long
reach a fixed temperature and keep this temperature throughout test section with 50 mm diameter, and of a 0.6-m long test
the experiment. In order to avoid emulsion stratification, section with 25 mm diameter. The sampling frequency of the
measurement began immediately after the sample was placed pressure signal was 500 Hz. A total of 60 000 data points were
on the sensor system. The CR mode, defined as a controlled shear collected, corresponding to the sampling time of 120 s.
rate input and determining the resulting shear stress, was used to A static mixer is a type of efficient pipeline mixing instrument
conduct most of the experiments. Each sample was measured at without moving elements which depends only on the pressure
least three times. If the repeatability of results was unacceptable, drop of fluid movement to realize hybrids. In this work, the
the CS mode would be used to perform the measurements. Here, emulsions pipe flow was formed by a SMV static mixer, provided

Figure 1. Schematic view of the flow loop.

VOLUME 94, FEBRUARY 2016 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 325
Table 1. Geometrical characteristics of the static mixers Table 2. Evolution of Ostwald de Waele (Power-law) model for
emulsions at 20 8C
D (mm) H/D Dh (mm) a
Emulsions Ostwald de Waele (Power) model
SMV-3.5/50 50 0.84 3.5 0.85
SMV-2.3/25 25 1.0 2.3 0.83 eo m n R2
D: the inner diameter of pipe; H: length of each element; Dh: hydraulic
0.1 0.0032 1.2192 0.9947
diameter of the static mixer; a: static mixer porosity
0.2 0.0065 1.1732 0.9951
0.3 0.0351 0.9585 0.9962
0.4 0.0232 1.0391 0.9999
by De-Yu Electrical Equipment (China). Each element of the 0.5 1.6436 0.4627 0.9142
SMV static mixer constituted a series of corrugated plates in 0.6 1.2611 0.5150 0.9134
different widths to obtain a cylindrical shape. The flow directions 0.7 0.2417 0.8499 0.9983
of adjacent plates were vertical to each other, and all of them 0.8 0.1214 0.9289 0.9994
were 458 compared to the axial pipe direction. In addition, the 0.9 0.0888 0.9420 0.9982
adjacent element of the SMV static mixer was 908 along the ring 1 0.0565 0.9728 0.9992
direction by the axial. The corresponding geometrical character-
istics are displayed in Table 1. For the static mixer, the theoretical
minimum droplet diameter is 10 mm, dependent on fluid The reason may be because shear stress is significantly reduced in
velocity. A detailed study of the static mixer can be found in regions of high shear rates. In addition, with increasing oil volume
Ghanem et al.[20] fraction, the flow behaviour index (n) first decreases and then
rises. However, the consistency index (m) shows the opposite
change to n. These findings reveal that oil and water emulsions
RESULTS AND DISCUSSION
with medium oil volume fractions have a stronger shear-thinning
rheological behaviour.
Rheology of Emulsions in the Rheometer
Figure 3 depicts the change of apparent viscosity with the oil
volume fraction at three different shear rates. Overall, the apparent
Apparent viscosity
viscosity increases sharply with increasing oil volume fraction
For an oil and water two-phase flow in pipeline transportation, and then reaches its maximum value. Afterward, the apparent
precise estimation of the apparent viscosity is very important viscosity decreases and ultimately reaches the value of oil phase
for accurately calculating the pressure drop as well as energy viscosity. The difference among the three curves is that at shear
requirements. In this work, 10 samples of emulsions are taken rates of 10 and 100 s1, the maximum value of apparent viscosity
during the rheological tests. occurs at ɛo ¼ 0.5 L/L, and at the high shear rate of 1000 s1 it
Figure 2 presents the rheograms of emulsions with oil volume appears at ɛo ¼ 0.7 L/L. The fluctuation of apparent viscosity as
fractions between 0.1–1.0 L/L at 20 8C. Generally, most samples a function of the oil volume fraction seems smaller at the shear
exhibit a certain degree of shear-thinning behaviour, especially rate of 1000 s1 compared to the shear rates of 10 and 100 s1.
those with medium oil volume fractions. Over the range of 0.5  ɛo  0.7 L/L the emulsions are very
According to Equation (1), the appropriate model parameters unstable, due to both oil and water coexisting as the continuous
can be identified by using the least squares method. The fitting or dispersed phases. In this process, the interaction between
results are presented in Table 2. As is shown, the power-law model two continuous phases is strongest.[1] In this work, the point of
gives a very high regression correlation coefficient (R2 > 0.99) for
emulsions with oil volume fractions between 0.1–0.4 L/L and 0.7–
1.0 L/L. However, for samples in the middle region (0.5–0.6 L/L),
the prediction results of the model are less accurate (R2 ¼ 0.91).

Figure 3. Apparent viscosity as a function of oil volume fraction for


Figure 2. Rheograms of the emulsions with different oil volume fractions. different shear rates.

326 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 94, FEBRUARY 2016
16
maximum value is defined as the phase inversion point. Thus, it f¼ ; in a laminar⁢ flow ð2Þ
can be concluded that the phase inversion point is dependent on ReMR
the shear rate for the synthetic emulsions.
0:079
f¼ ; in a turbulent flow ð3Þ
Droplet size distribution ReMR 0:25
Figures 4–5 present photomicrographs and droplet size distribu-
tions of samples at the phase inversion points for two different where the appropriate Reynolds number for non-Newtonian fluids
stirring times. Here, ɛo ¼ 0.5 and 0.7 L/L as in Figure 3. It can be is defined as follows:[21]
observed that the droplet size of the emulsions is smaller at 600 s
Dum rm
stirring time than at 300 s. The Sauter mean droplet size (D32) is ReMR ¼ ð4Þ
calculated as 22 mm for 300 s stirring time and 14 mm for 600 s. mapp
Clearly, the longer the stirring time, the smaller the droplet size.
Furthermore, the fine emulsions (600 s stirring) have a narrow where D and rm are the pipe diameter and the mixture density,
droplet size distribution and exhibits a lower tendency to respectively, and um is the mixture velocity of the fluids. The
flocculate. An interesting phenomenon is that droplet micro- apparent viscosity can be defined as follows:
structures are very similar at two different oil volume fractions  
when the stirring time is fixed. Namely, the emulsions of oil and 1 þ 3n n
mapp ¼ D1n 8n1 un1
m m ð5Þ
water possess similar droplet size distributions during the phase 4n
inversion region.
where m and n can be fitted by experimental data, and defined as
Flow Characterization of Emulsions in Pipes the fluid consistency coefficient in Pa  sn and the flow behaviour
index, respectively.
Pressure drop in horizontal flow The wall shear stress can be related to the frictional pressure
In a fully developed and steady flow of incompressible fluid in a drop through the following equation:[22]
pipe, the friction factor and Reynolds number are expressed by the  
following dimensionless forms of equations: D DP
tw ¼ ð6Þ
4 L

The friction factor is given as follows:


 
D DP
f 2
ð7Þ
2rm um L

The wall shear rate in a laminar power-law fluid flow can be


obtained as follows:[18,22]
 
 32Q 1  3n
gw ¼ ð8Þ
pD 4n

where Q refers to the mixture flow rate. Thus, in a laminar flow the
shear stress and shear rate can be acquired by measuring the
frictional pressure drop and flow rate.
On the basis of visual observations, the emulsion in the test
section formed by the SMV static mixer shows good dispersion,
Figure 4. Micrography for the effects of stirring time on emulsions at and neither droplet flotation nor stratification are evident. Figure 6
eo ¼ 0.5 and 0.7, respectively. indicates the pressure drop of emulsions in pipes with two
different diameters against oil volume fractions for different water
flow rates. As shown, the pressure drop rises with increasing oil
volume fraction. Moreover, at a fixed water flow rate the pressure
drop leads to a significant increase when reducing the pipe
diameter.
Figure 7 shows the change of pressure drop with the oil volume
fraction in two different pipe diameters at fixed mixture velocities.
The inversion route from oil-in-water emulsions to water-in-oil
emulsions is performed by maintaining a fixed mixture velocity
and increasing the volume fraction of the dispersed phase. The
graph indicates that an obvious peak appears in the pressure drop
curve when the oil volume fraction is around ɛo ¼ 0.8 L/L. This
phenomenon is associated with phase inversion. Here, it is
assumed to occur at or after the peak point of pressure drop. It can
be seen that the maximum pressure drop value is approximately
Figure 5. Droplet size distributions for emulsions with eo ¼ 0.5 and 0.7 at equal to twice the pressure drop of pure oil flow, and the peak
two different stirring times. point becomes sharper when increasing the mixture velocity.

VOLUME 94, FEBRUARY 2016 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 327
Figure 6. Frictional pressure drop of oil and water emulsions in pipes with Figure 7. Pressure drop at fixed mixture velocity against input oil volume
two different diameters at different oil volume fractions and flow rates fraction in two different pipe diameters.
(emulsions were formed in the pipe flow).

 
Interestingly, these points always appear at an oil volume fraction DP a2 D h
fh ¼ ð9Þ
of approximately 0.8 L/L, independent of the mixture velocity and L sm 2u2m rm
pipe diameter.
The wall shear stress measured in oil-in-water emulsions flow where (4P/L)sm refers to the frictional pressure drop in the static
as a function of the mixture flow rate is presented in Figure 8. As mixer, which is obtained with experimental measurement. The
expected, the shear stress increases when increasing the mixture ratio (um/a) is generally called “interstitial velocity,” and used as
flow rate, and the shear stress in the 25 mm pipe is greater than the characteristic velocity in porous media. The hydraulic
in the 50 mm pipe at a given mixture flow rate. Furthermore, a Reynolds number related to the friction factor is given as:
comparison between the shear stress measured and that calculated
by introducing single water viscosity (mw) into Equation (6) is also rm um Dh
Reh ¼ ð10Þ
presented. The mixture density is computed as the mean of oil and amapp
water phases. It is quite evident from the figure that the shear
stress using the water phase viscosity (mw), with a proper friction Here, the apparent viscosity (mapp) can be calculated using
factor defined by Equation (2), can predict the pressure drop in the Equation (5), in which the pipe diameter (D) is replaced by the
25 mm pipe fairly well. This phenomenon might be attributed to hydraulic diameter of the static mixer (Dh). In the present
there being significant slippage between a dispersed oil phase and experiments, the values of Reh are between 20–810. For the
a continuous water phase in the smaller pipe. Note that since the emulsions flow in the SMV static mixer, the work of Streiff
flow pattern of emulsions is an oil-in-water flow when the oil et al.[24,25] shows that the transient flow regime appears at Reh in a
volume fractions are between 0.1–0.6 L/L, oil droplets should not range from 20–2300. They proposed the following relationship
touch the pipe wall. Further, the pressure drop in a small pipe is to predict the Sauter mean droplet diameter (D32) in transient and
dominated by shear stress for the wetting water phase around the turbulent flow regimes:
pipe wall. Thus, the mixture viscosity can be taken as the viscosity
of the continuous phase.[23] When the pipe diameter is larger, the D32
¼ 0:21We0:5 Re0:15 ð11Þ
phenomena of oil droplet flotation and coalescence are more likely Dh h h

to occur, and small amounts of the wetting oil phase at the top of
the pipe wall also become important to the contribution of shear
stress.
Mean droplet diameter formed by the static mixer
As discussed above, the droplet size and droplet size distribution
are important for accurately identifying the transport properties of
emulsions in pipes. In pipe flow experiments, after the oil and
water mixtures flow through the SMV static mixer, the emulsions
are formed and immediately enter the test section. Here, the
droplet size distribution in pipe flow was not measured directly,
but was calculated by measuring the frictional pressure drop in the
static mixer.
There are a number of relationships in the literature for
predicting the mean droplet diameter in the static mixer based on
several dimensionless numbers. In the static mixer, the pressure
drop is expressed using the friction factor, taking into account Figure 8. Comparison between shear stress measured (symbols) and
some geometrical parameters: the porosity of the static mixer (a) computed by introducing single phase viscosity to Equation (7)
and the hydraulic diameter of the static mixer (Dh) as: (continuous line) in oil-in-water emulsions flow.

328 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 94, FEBRUARY 2016
Here, the Weber number can be expressed by taking into account
Table 3. Weber number and Capillary number in the SMV static mixer
the fluid mixture properties, hydraulic diameter, and interstitial at a fixed oil volume fraction
velocity:
No um eo D32 Weh Cap (104)
rm u2m Dh
Weh ¼ ð12Þ 1 0.28 0.5 336 12 9.91
sa2 2 0.42 0.5 245 26 11.95
3 0.57 0.5 197 46 13.65
In the present study, the relationship between Sauter mean droplet 4 0.71 0.5 166 72 15.14
diameter (D32) and corresponding mean energy dissipation rate 5 0.85 0.5 144 104 16.47
(z) is used to verify the calculation of emulsion flow droplet
size.[26,27] The mean energy dissipation rate is the source of energy
for forming the emulsion flow of oil and water, and is calculated Weber number and Capillary number in the SMV static mixer
by the frictional pressure drop, measured as follows: (3.5/50) at a fixed oil volume fraction (eo ¼ 0.5 L/L). When the
  mixture velocity increases from 0.28 to 0.85 m/s, the Sauter mean
DP um
z¼ ð13Þ droplet diameter is reduced. Combined with Equation (12), small
L sm arm and stable oil droplets in the static mixer can be obtained by
increasing the apparent viscosity or the mixture velocity. Namely,
The Sauter mean droplet diameters obtained in the present study, an increase in Weber number or Capillary number will result in a
and those in the available literature as the function of mean energy more stable droplet size distribution. Compared with Figure 5, the
dissipation in the SMV static mixer, are presented in Figure 9. values of the Sauter mean droplet diameters formed by the SMV
Experimental data are obtained for oil-in-water emulsion flow static mixer are approximately one order of magnitude larger than
with oil volume fractions between 0.1–0.6 L/L. For each data those measured by the rheometer. Therefore, the stability of the
series, the D32 is closely linearly related to z in a logarithmic emulsions in pipe flow is not as good as that in the rheometer,
coordinate system. Compared to results in the literature, the Sauter and these should be considered coarse emulsions.[3,8]
mean droplet diameters obtained in the present study are smaller
than those measured by Streiff.[24] This discrepancy might be Comparison between Rheological Measurements and Laminar
owing to the different dispersed phase volume fractions between Flow Data
two systems. In fact, the discrepancy is very important and should
affect the apparent viscosity and friction factor. Moreover, the Shear stress and friction factor
relationship of D32 and z in this work is very close to that of In a laminar flow of power-law fluid in pipes, the shear stress and
Lobry et al.[27] In their work, the emulsions flow in the SMV shear rate can be calculated using Equations (6,8). Figure 10
mixer is located in the turbulent regime so that the droplet size compares the experimental data of the shear stress with the shear
decreases with increasing mean energy dissipation. These results rate obtained in pipes and the data measured in the rheometer. It
demonstrate the accuracy of the method suggested for predicting can be seen that at low oil volume fractions, the shear stresses
Sauter mean droplet diameters in pipe flow. obtained by the two systems show close agreement, especially for
The stability of oil droplets formed by the SMV static mixer can the emulsions flow in the 50 mm pipe. With the increase of oil
be evaluated using the Capillary number in pore-scale fluid flow, volume fraction, the shear stresses of the emulsions in pipes are
which is suggested by Jamaloei et al.:[28–30] far lower than in the rheometer. This is due not only to the
different droplet size distributions, but to the different rheological
mapp um
Cap ¼ ð14Þ
sa2

where mm is the apparent viscosity of oil and water emulsions in


mPa  s. Table 3 presents the oil droplet size distribution versus

Figure 10. Comparison between experimental data of the shear stress


versus the shear rate obtained for emulsions of oil-in-water laminar flow in
Figure 9. Sauter mean diameters obtained in the present study and those pipes with those measured in the rheometer (solid lines ¼ rheograms of
in the available literature as a function of the mean energy dissipation in the emulsions measured in the rheometer; open squares and
SMV static mixer. circles ¼ experimental points in two different pipe diameters).

VOLUME 94, FEBRUARY 2016 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 329
Figure 11. Experimental friction factor versus Reynolds number in the Figure 12. Comparison of the apparent viscosities of the emulsions at
laminar flow of oil-in-water emulsions in two pipes with 25 and 50 mm different oil volume fractions by two systems (solid symbols ¼ data
diameters by using the Power-law fluid relationship in Table 2. measured with the rheometer; open symbols ¼ data formed in the pipe
flow and calculated from the frictional pressure drop).
behaviours around the phase inversion regions of both systems.
This result also validates the findings of Masalova et al.,[16] in pipe flow are closer to those measured at the high shear rates in
which the stress-strain relationships obtained in the rheometer the rheometer.
are more suitable for predicting the transport characteristics of In the literature, most existing viscosity prediction models for
emulsions in large pipes than in small pipes. emulsions are independent of shear rate;[3] only Pal and Rhodes[31]
For laminar-turbulent regime transition in a dispersed oil and developed a semi-theoretical model, which was based on the
water flow in pipes, the results of Pal[8] and Pouplin[12] show that method of Brinkman[32] to predict the apparent viscosity prepared
the Hagen-Poiseuille relationship is valid for Reynolds numbers in the rheometer. In Figure 12, the apparent viscosity of emulsions
up to 4000 in the case of coarse emulsions. Therefore, in the at shear rate 1000 s1 is closest to that in the pipe flow. In addition,
following study the emulsions flow in pipes is employed to identify this work does not build or validate any models for apparent
the laminar flow when the Reynolds number is < 4000. The viscosity prediction. Thus, the apparent viscosity of emulsions in
experimental friction factor versus Reynolds number in oil-in- the laminar flow was replaced by simply using values measured
water emulsions flow in the 25 and 50 mm pipes is shown in at shear rate 1000 s1 in the rheometer. Figure 13 shows the
Figure 11. Here, the experimental friction factor is back-calculated experimental friction factor versus Reynolds number in the
using the measured frictional pressure drop. In this process, laminar flow of emulsions calculated using the apparent viscosity
the consistency index (m) and the flow behaviour index (n) measured at shear rate 1000 s1. In combination with Figure 12,
are introduced into Equation (5) based on different oil volume it can be found that this method significantly improves the
fractions, as shown in Table 2. Note that an acceptable prediction prediction accuracy. In practice, for engineering applications, the
is obtained between theory and data for emulsion flows with low integral transport characterization is determined primarily by
oil fractions (0.1–0.2 L/L), but for high oil fractions (ɛo ¼ 0.3– the rheological properties of emulsions at high shear rates. In the
0.6 L/L), the experimental friction factors are scattered and far rheometer, the details of rheological behaviour of emulsions at
lower than the Hagen-Poiseuille relationship. This predicted result very low shear rates are not as important as those at high shear
is not consistent with the emulsions flow of water-in-oil in rates.[16] Therefore, the apparent viscosity obtained by rheology
the pipe;[18] previous findings show that the Hagen-Poiseuille
relationship can predict the emulsions flow in the pipe fairly well.
In the present experiments, although the static mixer was used
to generate emulsions before the test section, the emulsions
are still coarse and unstable in the laminar regime. In addition, it
should be stressed that the water-in-oil emulsion is more stable
than oil-in-water due to the high mean energy dissipation.
Apparent viscosity and phase inversion
For the oil and water mixture flow in pipes, the phase inversion
region can be defined as an ambiguous range of the phase volume
fraction in which both oil and water phases coexist as either
dispersed or continuous. Based on visual observations, the phase
inversion points in a horizontal flow always take place at an oil
volume fraction of 0.8 L/L, independent of the mixture velocity
and pipe diameter. Figure 12 compares the apparent viscosity
through the entire oil volume fraction range by two systems. One
can see that the apparent viscosities in the 25 mm pipe are slightly Figure 13. Experimental friction factor versus Reynolds number in the
higher than in the 50 mm pipe, and closer to the data from the laminar flow of emulsions calculated using the apparent viscosity measured
rheometer. It can also be seen that the phase inversion points in the at shear rate 1000 s1.

330 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 94, FEBRUARY 2016
measurements at high shear rates could be introduced for [4] N. Zaki, T. Butz, D. Kessel, Petrol. Sci. Technol. 2001, 19, 425.
accurately calculating the mixture Reynolds number and the [5] R. Pal, J. Rheol. 2001, 45, 509.
friction factor in emulsions flow. [6] R. Pal, Curr. Opin. Colloid Interface Sci. 2011, 16, 41.
[7] M. Meriem-Benziane, S. A. Abdul-Wahab, M. Benaicha, M.
CONCLUSIONS Belhadri, Fuel 2012, 95, 97.
Experimental studies of the flow curve, apparent viscosity, phase [8] R. Pal, AIChE J. 1993, 39, 1754.
inversion, and droplet size distribution for oil and water emulsions [9] M. N€adler, D. Mewes, Int. J. Multiphas. Flow 1997, 23, 55.
were made using a Haake RS6000 rheometer and a lab scale flow [10] A. Omer, R. Pal, Chem. Eng. Technol. 2010, 33, 983.
loop. The rheology measurements indicate that the synthetic [11] A. Omer, R. Pal, Ind. Eng. Chem. Res. 2013, 52, 9099.
emulsions exhibit shear-thinning behaviours, especially those
[12] A. Pouplin, O. Masbernat, S. Décarre, A. Liné, AIChE J. 2011,
close to the phase inversion region. In this region, the emulsions
57, 1119.
show similar droplet size distributions. In addition, the phase
inversion points are dependent on the shear rate for the synthetic [13] J. C. Vielma, O. Shoham, R. S. Mohan, L. E. Gomez, SPE J.
emulsions. 2011, March, 148.
The pipe flow measurements demonstrate that the phase [14] I. H. Rodriguez, H. K. B. Yamaguti, M. S. de Castro, M. J. Da
inversion points invariably occur at an oil volume fraction of Silva, O. M. H. Rodriguez, AIChE J. 2012, 58, 2900.
0.8 L/L independent of the mixture velocity and pipe diameter. [15] J. Plasencia, B. Pettersen, O. J. Nydal, J. Petrol. Sci. Eng.
In the process of phase inversion, the apparent viscosity of the 2013, 101, 35.
emulsions becomes very large, so that a high pressure drop or a [16] I. Masalova, A. Y. Malkin, P. Slatter, K. Wilson, J. Non-
low flow rate can appear in the pipes. Therefore, in oil- Newton. Fluid Mech. 2003, 112, 101.
producing wells, fluid mixtures with an oil volume fraction of lu, B. H. Pettersen, J. Sjo
€blom, J. Petrol. Sci. Eng.
[17] S. KeleSs og
0.80 L/L should be avoided, because the emulsion flow in this
2012, 100, 14.
region could reduce the efficiency of downhole pump and well
[18] lu, B. H. Pettersen, J. Sjo
S. KeleSs og €blom, J. Dispersion Sci.
productivity.[33]
In oil-in-water emulsions flow, the shear stress in small pipes Technol. 2012, 33, 536.
can be predicted fairly well by using the water phase viscosity, due [19] M. Al-Yaari, A. Al-Sarkhi, I. A. Hussein, F. Chang, M. Abbad,
to the significant slippage between the two phases. Consequently, Chem. Eng. Res. Des. 2014, 92, 405.
when the fluid mixtures are transported by a small-diameter [20] A. Ghanem, T. Lemenand, D. D. Valle, H. Peerhossaini,
pipeline, the apparent viscosity can be approximated by the Chem. Eng. Res. Des. 2014, 92, 205.
viscosity of the continuous phase. [21] A. B. Metzner, J. C. Reed, AIChE J. 1955, 1, 434.
Comparing rheological measurements and laminar flow data in [22] R. P. Chhabra, J. F. Richardson, Non-Newtonian Flow in the
the flow pattern of oil in water shows that the values of Sauter Process Industries, Butterworth-Heinemann, Oxford 1999.
mean droplet diameters formed by the SMV static mixer are
[23] T. Baron, C. S. Sterling, A. P. Schueler, “Viscosity of
approximately one order of magnitude greater than those in the
suspensions - review and applications of two-phase flow,”
rheometer. Therefore, the emulsions in the pipe flow should
Proceedings of the 3rd Midwestern Conference on Fluid
be considered coarse emulsions. Furthermore, the stress-strain
Mechanics, University of Minnesota, MN, USA, 23–25 March
relationships obtained in the rheometer are more suitable for
1953.
predicting the transport characteristics of emulsions in large pipes
than those in small pipes. [24] F. A. Streiff, Sulzer Technol. Rev. 1977, 3, 108.
Due to the phase inversion points in the pipe flow being closer [25] F. A. Streiff, P. Mathys, T. U. Fisher, Rec. Progr. Genie Proc.
to those measured at high shear rates in the rheometer, the 1997, 11, 307.
apparent viscosity obtained by the rheology measurements at [26] F. Theron, N. Le Sauze, Int. J. Multiphas. Flow 2011, 37, 488.
high shear rates could be introduced for accurately calculating the [27] E. Lobry, F. Theron, C. Gourdon, N. Le Sauze, C. Xuere, T.
mixture Reynolds number and the friction factor in the pipeline Lasuye, Chem. Eng. Sci. 2011, 66, 5762.
transportation of crude oil and water emulsions.
[28] B. Y. Jamaloei, R. Kharrat, J. Petrol. Sci. Eng. 2012, 92–93, 89.
[29] B. Y. Jamaloei, R. Kharrat, K. Asghari, F. Torabi, J. Petrol. Sci.
ACKNOWLEDGEMENTS Eng. 2011, 77, 121.
The financial support of the Special Development of National Key [30] B. Y. Jamaloeia, K. Asghari, R. Kharrat, Exp. Thermal Fluid
Scientific Instruments in China (2011YQ120048-02) and National Sci. 2011, 35, 253.
Natural Science Foundation of China (No. 51509235) is responsi- [31] R. Pal, E. Rhodes, J. Rheol. 1989, 33, 1021.
ble in part for the conduct of the present study. [32] H. C. Brinkman, J. Chem. Phys. 1952, 20, 571.
[33] F. Torabi, B. Y. Jamaloei, S. Y. Hong, N. K. Bakhsh, J. Petrol.
REFERENCES Explor. Prod. Technol. 2012, 2, 181.
[1] L. Y. Yeo, O. K. Matar, E. S. Perez de Ortiz, G. F. Hewitt,
Multiphase Sci. Technol. 2000, 12, 51.
[2] P. Partal, A. Guerrero, M. Berjano, C. Gallegos, J. Am. Oil Manuscript received December 5, 2014; revised manuscript received
Chem. Soc. 1997, 74, 1203. May 3, 2015; accepted for publication May 5, 2015.
[3] J. Zhang, J.-Y. Xu, M.-C. Gao, Y.-X. Wu, J. Dispersion Sci.
Technol. 2003, 34, 1148.

VOLUME 94, FEBRUARY 2016 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 331

You might also like