You are on page 1of 15

Journal of Petroleum Science and Engineering 175 (2019) 881–895

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Experimental and numerical study of oil drop motion within an ESP impeller T
a a,∗ b
Rodolfo Marcilli Perissinotto , William Monte Verde , Mariana Gallassi ,
Gabriel F.N. Gonçalvesb, Marcelo Souza de Castroa, João Carneirob, Jorge Luiz Biazussia,
Antonio Carlos Bannwarta
a
School of Mechanical Engineering, University of Campinas, Rua Mendeleyev, 200, Cidade Universitária, Campinas, SP, CEP 13083-860, Brazil
b
ISDB FlowTech, Rua Lauro Müller, 116/3006, Botafogo, Rio de Janeiro, RJ, CEP 22290-160, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: The Electrical Submersible Pump (ESP) is a multistage centrifugal pump used in the petroleum industry as an
Centrifugal pump artificial lift method. The ESP usually works with the presence of two-phase liquid-liquid flows that constitute
Two-phase liquid-liquid flow dispersions and emulsions, causing performance losses and operational problems. This research aims to in-
Flow visualization vestigate the behavior and evaluate the dynamics of individual oil drops in an oil-in-water dispersion within an
Oil drop dynamics
ESP impeller. The study adopts experimental and numerical approaches. Initially, experiments were performed
CFD
using an experimental facility with a high-speed camera and an ESP prototype working at 600 rpm and 900 rpm,
for water flows around the Best Efficiency Point (BEP) and with the injection of oil drops at a low flow rate. The
acquired images were processed, and a drop sample was tracked, enabling the analysis of the size, shape, path,
velocity, and acceleration of the oil drops. Numerical simulations were executed in ANSYS® software to define
relevant parameters related to water and oil drops, such as velocities, accelerations, forces, turbulent dissipation,
and residence time. The images reveal a unique flow pattern of dispersed drops in a continuous water phase. The
oil drops' diameters vary from tenths of a millimeter to around 3 mm. The drops' trajectories can be classified
into three different regions within the impeller channels. The drops’ velocities stay in the order of 1 m/s, while
accelerations can reach hundreds of m/s2. The velocity profiles show that the oil drops tend to decelerate during
their trajectory, while the acceleration profiles suggest peaks at the channel inlet and outlet. High intense tur-
bulence is present in the impeller entrance zone. The evaluation of the residence time and the particle Reynolds
number suggest that smaller oil drops follow the water streamlines, while larger oil drops tend to be affected by
external forces. The main forces that govern the oil drop motion are the drag, the pressure gradient, and the
virtual mass forces. The force from the pressure gradient is tenfold greater than the force from the drag. The
virtual mass effect is significant only in the impeller inlet. In general, in this research, numerical results show a
satisfactory agreement with the experimental data.

1. Introduction (ESP). ESPs are largely used in onshore and offshore wells, with low and
high liquid flow rates (Takacs, 2017). It is estimated that around the
The petroleum industry uses the term lifting to refer to the process world more than 100,000 wells operate with ESPs (Flatern, 2015).
of vertically transporting fluids extracted from a reservoir, which must Although ESPs are widely employed, several aspects related to their
flow from the bottom of the well to the production facilities. When the usage are still problematic. This is especially the case when it comes to
reservoir pressure is insufficiently high to overcome the weight of the operating in the presence of gas or viscous fluids, conditions quite
fluid column as well as the frictional pressure drop, producers must rely common in applications associated with the petroleum industry.
on external energy provided by artificial lifting methods. Examples of operation with high-viscosity fluids include heavy oil fields
One of the most consolidated techniques for lifting oil is the elec- and the formation of liquid-liquid mixtures when water is present. In
trical submersible pumping, in which the fluids are elevated by a this last case, the phase distribution can be arranged as a dispersion,
multiple-stage centrifugal pump, called Electrical Submersible Pump depending on the flow conditions.


Corresponding author.
E-mail addresses: rodolfomp@fem.unicamp.br (R.M. Perissinotto), wmv@unicamp.br (W. Monte Verde), mariana.gallassi@isdbflowtech.com (M. Gallassi),
gabriel.farah@isdbflowtech.com (G.F.N. Gonçalves), mcastro@fem.unicamp.br (M.S.d. Castro), joao.carneiro@isdbflowtech.com (J. Carneiro),
biazussi@unicamp.br (J.L. Biazussi), bannwart@fem.unicamp.br (A.C. Bannwart).

https://doi.org/10.1016/j.petrol.2019.01.025
Received 11 July 2018; Received in revised form 13 December 2018; Accepted 6 January 2019
Available online 09 January 2019
0920-4105/ © 2019 Elsevier B.V. All rights reserved.
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Nomenclature P Modified pressure


P1 Pressure at the inlet of ESP prototype
Abbreviations P2 Pressure at the outlet of ESP prototype
q Fraction of QBEP
BEP Best Efficiency Point Qw Water volume flow rate
BSL Reynolds Baseline Model QBEP Water flow rate at BEP
CFD Computational Fluid Dynamics Ri , Ro Impeller radius: from center to inlet and outlet
DPM Discrete Phase Model R Radial position of oil drop in polar reference
ESP Electrical Submersible Pump R∗ Normalized radial position of oil drop
FBRM Focused Beam Reflectance Measurement Ṙ First derivatives of radial position
FG Full Geometry R̈ Second derivatives of radial position
GVF Gas Volume Fraction Red Particle Reynolds number
HSC High-Speed Camera t t-distribution
MRF Moving Reference Frame t Time
PIV Particle Image Velocimetry tres Impeller residence time

PTV Particle Tracking Velocimetry tres Normalized residence time
RANS Reynolds-Averaged Navier-Stokes T Temperature of water

RG Reduced Geometry U Velocity of water
SMM Sliding Mesh Model Utip Blade velocity
SST Menter's Shear Stress Transport VR Radial velocity of an oil drop
VM Virtual Mass Vθ Swirl velocity of an oil drop
VOF Volume of Fluid →
V Velocity of oil drop in non-inertial system
VSD Variable Speed Driver →
VI Velocity of oil drop in inertial system
x, y Oil drop position in Cartesian reference
Latin Letters y+ Dimensionless distance from the wall
z Position perpendicular to xy or Rθ planes
A Constant in normalized oil drop position
AR Radial acceleration of an oil drop Greek Letters
Aθ Swirl acceleration of an oil drop

A Acceleration of drop in non-inertial system ΔP Difference between pressures P1 and P2
⎯→

AI Acceleration of drop in inertial system Δt Time between two consecutive images
CD Drag coefficient Δφ Confidence band of a generic variable
CVM Virtual mass effect δij Kronecker's delta
d Oil drop diameter ε Turbulent eddy dissipation
dij Deformation tensor θ0 Initial angle of an oil drop
D Impeller diameter θc Channel angle
f (t ) Generic function - position θ Swirl position of oil drop in polar reference
f ′ (t ) First derivative in finite-difference method θ∗ Normalized swirl position of oil drop
f ′′ (t ) Second derivative in finite-difference method θ̇ First derivatives of swirl position
f + (R∗) Central path position function θ̈ Second derivatives of swirl position
f − (R∗) Central path position function κ Turbulent kinetic energy
g (t ) Generic function - velocity or acceleration μw Dynamic viscosity of water
g (t )avg Correction by the moving average method μo Dynamic viscosity of oil
⎯→
⎯ μt Turbulent viscosity
FD Drag force
⎯→
⎯ μef Effective viscosity
FL Lift force
⎯→
⎯ νw Kinematic viscosity of water
FP Pressure gradient force
⎯→
⎯ ρw Water density
FR Resultant force ρo Oil density
⎯⎯⎯⎯⎯→
FVM Virtual mass force σ Standard deviation
g Gravity σa / o Oil-air surface tension
h Time step in finite-difference method σw / o Oil-water interfacial tension
K Constant in Saffman's lift force equation φ Mean of a generic variable
mo Mass of an oil drop Φ Dimensionless flow rate
mw Mass flow rate of water Ψ Dimensionless pump head
n Sample size ω Angular speed of impeller
N Rotational speed of impeller ∇s p Pressure gradient along a streamline
Nc Number of impeller channels
p Static pressure

Indeed, oil production is closely related to the occurrence of dis- the impeller, resulting in the break-up of phases into small droplets.
persions and emulsions (Abivin et al., 2009). Their formation depends The apparent viscosity of a dispersion may be higher than the pure
on mechanical processes that provide the energy necessary to dissipate oil viscosity. In the case of ESPs, the presence of water can increase the
one liquid into the other (Sjoblom, 2005). In this sense, the ESP can effective petroleum viscosity and consequently impair pump perfor-
promote the creation of dispersions, due to the fluids being agitated in mance, leading to a significant rise in operational costs. The behavior of

882
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

real flows is very different from that observed in single-phase flows. (2016), and Cavicchio et al. (2018). In the scope of centrifugal pumps,
Thus, the project of designing an ESP system calls for engineers to experimental studies on dispersions and emulsions are few, such as
possess a detailed comprehension of multiphase flow dynamics within Khalil et al. (2008), Morales et al. (2013), Bulgarelli et al. (2017a,
centrifugal pumps. 2017b), and Perissinotto et al. (2018). There are also some numerical
Considerable efforts have been made to investigate two-phase flows works related to ESPs, including those by Mundo (2014), Stel et al.
inside pumps. Computational Fluid Dynamics (CFD) methods are (2015), and Zhu et al. (2016), the last ones concerning multistage
commonly adopted to simulate flow in a computational environment pumps. However, no studies are available with focus on dynamics of
and may provide detailed insight into flow fields and dynamics. Flow- individual liquid drops.
visualization techniques are used to experimentally identify flow pat- Khalil et al. (2008) experimentally investigated the performance of
terns and analyze the interaction between the phases present in a centrifugal pumps working with oil-in-water emulsions. The authors
mixture. A powerful method for flow visualization is the high-speed concluded that the presence of emulsions degrades the pump head and
imaging, as it helps researchers to observe transient phenomena with a efficiency. The performance also suffers a reduction when the holdup
satisfactory spatial and temporal resolution (Mohammadi and Sharp, and temperature decrease.
2013). Morales et al. (2013) studied the dispersion formation in a pump
Regarding two-phase gas-liquid flows, initial studies were focused working with a dispersion of water and oil. The authors used a particle-
on the nuclear industry, where centrifugal pumps were employed in size analyzer to measure the drop-size distribution and concluded that
reactor cooling systems and there was a concern about radioactive the drop size mostly depends on the pump rotation. The drops become
fluids leaking. The pioneer in the petroleum industry was Estevam smaller as the rotation increases.
(2002), who developed the first ESP prototype for experimental vi- Mundo (2014) conducted experimental and numerical evaluations
sualization of air-water flows. for water-in-oil emulsions within an ESP. The author studied the effect
Two-phase gas-liquid flows within pumps were also examined by of water fraction and drop size on performance curves, using a Reynolds
such authors as Barrios and Prado (2011), Trevisan and Prado (2011), Baseline Model (BSL) for turbulence. For both oils tested, the author
Zhang et al. (2016), Monte Verde et al. (2017), and Cubas Cubas verified that there was a reduction in the head and flow rate capacity,
(2017). These authors, who used high-speed cameras to visualize flow and an increase of the emulsion effective viscosity with higher con-
in impellers, aimed to identify flow patterns; observe the phenomena as centrations.
accumulation, agglomeration, and coalescence; and investigate how the Stel et al. (2015) performed CFD simulations in order to investigate
behavior of bubbles was influenced by rotational speed, liquid flow the single-phase flow in a multistage ESP. They evaluated the influence
rate, intake pressure, viscosity, and surface tension. of the turbulence model and the number of stages on the pump per-
Caridad and Kenyery (2004), Caridad et al. (2008), Barrios et al. formance and compared some numerical and experimental results. The
(2009) and Pineda et al. (2016) studied gas-liquid flows inside ESPs authors verified that the Menter's Shear Stress Transport (SST) turbu-
using CFD simulations. Caridad and Kenyery (2004) adopted a two- lence model was more suitable for their system.
fluid model to predict pressure, velocity fields, and Gas Volume Frac- Zhu et al. (2016) were interested in the oil viscosity effect on
tion (GVF). To model turbulence, the authors used the κ-ε model. They multistage ESPs. They conducted experimental analysis and numerical
successfully predicted the presence of gas pockets for different GVFs simulations, varying oil viscosity, flow rate, and rotation speed. Steady-
inside the pump. Later, Caridad et al. (2008) investigated the ESP head, state simulations with SST turbulence model indicated that the pump
bubble diameter, flow angle at the impeller outlet and phase distribu- efficiency starts to decrease when the oil viscosity is higher than 200 cP.
tion within the channels. The results presented excellent agreement Bulgarelli et al. (2017a, 2017b) used an 8-stage ESP operating with
with experimental data. The authors predicted that, due to the low emulsions to study the phase inversion phenomena and the chord-
velocity of the liquid, the gas would agglomerate on the high-pressure length distribution. Tests were carried out at various viscosities and
side of the impeller blade. The κ-ε turbulence model was also used by rotation speeds. The phase inversion occurs at water fractions ranging
Barrios et al. (2009) to perform one-fluid and two-fluid simulations. between 10% and 30%. For low water fractions, oil is the continuous
They observed some vortices at the trailing edge on the suction side and phase, with a higher effective viscosity that degrades the pump per-
also noticed a recirculation zone inside the channels. Using a correla- formance. For high water fractions, water is the continuous phase, with
tion for the drag coefficient, the authors verified an accumulation of gas a lower effective viscosity that increases the pump performance. Using
at the intake. Finally, Pineda et al. (2016) used the Volume of Fluid a Focused Beam Reflectance Measurement (FBRM) technique, the au-
(VOF) model to obtain the void fraction and the pump head. In a thors observed that the chord length increased up to the phase inversion
comparison with experimental measurements, the CFD was capable, point and dropped after it.
under certain conditions, of determining the flow behavior. Fluctua- In short, as the literature review exposes, the two-phase liquid-li-
tions occurred only at high rotation speeds. quid flows with dispersions in ESPs are still insufficiently understood,
Minemura and Murakami (1980), Sabino (2015) and Ofuchi et al. despite their relevant application in the petroleum industry. Unlike
(2017) explored the air bubble dynamics inside impeller channels. their study of gas bubbles, researchers have not thoroughly investigated
According to them, the movement of the bubbles in the pump is con- liquid drops. Most studies in the literature examine the overall pump
trolled by five forces: a drag force due to the movement of the bubble performance, but there is a lack of studies exploring the dynamics of
relative to the water, a pressure force related to the pressure gradient in liquid phases.
the channel, a buoyancy force due to the difference of densities between The objective of this research is to study, through a joint study of
the fluids, an inertial force as a consequence of the water acceleration experiments and numerical simulations, the motion and the dynamics
and a force that considers the historical effect of the flow around the of individual liquid drops subject to a rotating field. First, oil drops are
bubble. However, the authors agree that, in the cases investigated, the visualized in a turbulent water flow inside the impeller of an ESP
most important forces acting on air bubbles were the drag force and the prototype. Images are captured and processed to obtain size, shape,
pressure gradient force. position, trajectory, velocity, and acceleration of oil drops. The flow is
Concerning two-phase liquid-liquid flows, pipelines are dealt with then simulated using a CFD method, which enables the evaluation of
in the vast majority of works, such as those by Brauner and Maron turbulent dissipation, residence time, velocities, accelerations, and
(1992), Alkaya et al. (2000), Angeli and Hewitt (2000), Lum et al. forces. Finally, numerical results are compared with the experimental
(2006), Rodriguez and Oliemans (2006), Vielma et al. (2007), Grassi ones.
et al. (2008), Wang et al. (2011), Castro et al. (2012), Rodriguez and The results can provide a better understanding on the motion of
Baldani (2012), Hanafizadeh et al. (2015), Loh and Premanadhan dispersed drops in pumps. In other words, this paper contributes to an

883
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

improved comprehension on the behavior of dispersions in two-phase Control valves regulate pressure levels and flow rates in the lines.
flows in ESPs. In the future, the results can be expanded to other cases Moreover, each pump is controlled by a Variable Speed Driver (VSD)
and associated with the efficiency and performance of different pumps. that sets the rotational speed and thus changes the flow rate. The ro-
In addition, the slip between both phases can be calculated when water tational speed of the peristaltic pump is used to quantify the oil flow
and oil velocities are evaluated. The long-term industrial application is rate, which has very low values when compared to the water flow rate.
the proposition of one-dimensional models in order to predict the best
type of ESP for each application, but also to find optimum operational 2.2. Test matrix
parameters for a given system.
Performance curves were determined for the ESP prototype oper-
2. Experimental methodology ating with a water single-phase flow and an oil-water two-phase flow.
The results were used as a reference for the visualization experiments.
An experimental facility with an ESP prototype was used to perform The visualization tests for oil-in-water dispersions were carried out
experiments at different operational conditions. Water and a mineral oil at two rotation speeds (N ) and three water volume flow rates (Q w )
were selected to compose the two-phase liquid-liquid flow. A High- around the best efficiency points (BEP), as presented in Table 1. In all
Speed Camera (HSC) was employed to capture images of the flow the tested conditions, the oil was injected at a constant flow rate of
within the prototype impeller. 2 ml/s, i.e., 0.007 m³/h.
The water flow rates corresponding to the best efficiency points
2.1. Experimental facility (QBEP ) were determined by the analysis of the single-phase performance
curves.
The tests were performed using an experimental setup that consists As already mentioned, the two-phase liquid-liquid flow is a mixture
of a water circuit, an oil injection system, and a visualization prototype of tap water and a mineral oil. Chemical tests were executed to char-
based on the Baker-Hughes ESP P23. The ESP prototype was designed acterize the oil properties at 25 °C; the tests yielded a density ( ρo ) of
and built by Monte Verde et al. (2017), who minutely present the pump 880 kg/m³, a dynamic viscosity ( μo ) of 220 cP, a surface tension (σa / o ) of
characteristics and the impeller geometry. The test bench is shown in 32 mN/m, and an interfacial tension (σw / o ) of 34 mN/m. All perfor-
Fig. 1. mance and visualization experiments were carried out at the tempera-
The ESP prototype has an impeller with a transparent top shroud ture of 25 °C.
made of an acrylic material, which enables the visual access to the flow The originally transparent oil was darkened with a black dye to
in all the channels. Monte Verde et al. (2017) used the prototype to enhance the contrast between the liquid and the white impeller. A
study gas-liquid two-phase flows; Perissinotto et al. (2018) then up- higher contrast improves the quality of the flow images. The dye does
dated the prototype to study liquid-liquid two-phase flows as well. not affect the oil chemical properties nor the water color.
A high-speed camera model VEO 640, manufactured by Phantom®, Even though the rotation speeds employed in real ESPs in the pet-
is responsible for the flow visualization. The equipment acquires images roleum industry can reach 3500 rpm, some restrictions in the experi-
with a maximum resolution of 2560 x 1600 pixel at a rate of 1400 ments limited the chosen test conditions to 600 rpm and 900 rpm. At
frames per second. Three reflectors are used as light sources, each with high rotational speeds and flow rates, the oil drops become so small that
84 Watts and 7700 lumens. the high-speed camera resolution is not sufficient to capture adequate
A schematic illustration of the experimental facility is presented in images. Moreover, the ESP acrylic shroud may suffer severe damages
Fig. 2. In the water line, a booster pump draws the liquid from a 0.5 m3 and may even break when subjected to high pressures.
separation tank. The water flows through a Coriolis meter and a shell-
and-tube heat exchanger, before finally reaching the prototype suction.
2.3. Experimental procedure
At the same time, in the oil line, a peristaltic pump drags the liquid out
of a 10-L reservoir. The oil is injected into the prototype near the im-
In the performance tests for single-phase flow, the prototype rota-
peller inlet through a capillary tube with 1/16-inch inner diameter. The
tion remains constant throughout the test. The water flow rate is first
oil becomes dispersed in the water flow. Both liquids enter, traverse and
adjusted to a maximum value that coincides with a zero-pressure gen-
exit the impeller channels, while the high-speed camera captures
erated by the ESP prototype. This condition is called open-flow.
images of the flow. The dispersion then goes to the separation tank
Afterward, data is acquired at 1 kHz frequency during 30 s and the
where, by differences in density, the water is gravitationally separated
average quantity is storage. Next, the discharge control valve is slightly
from the oil. The water flows back to its line, while the separated oil is
closed to reduce the water flow rate and data is acquired again. In this
dewatered and returns to the oil tank.
The Coriolis flowmeter model RHM12, fabricated by Metroval,
measures the water mass flow rate (m w ). It has a limit range of 6000 kg/
h and an accuracy of 0.2%. A resistance temperature detector type
PT100, manufactured by ECIL, with 1/10 DIN accuracy, quantifies the
dispersion temperature (T ) in the prototype inlet. Two capacitive
pressure transducers series 2088, manufactured by Emerson
Rosemount™, measure the intake (P1) and discharge (P2 ) pressures with
an accuracy of 0.05%. The analog output signals are acquired by a
National Instruments system. The data is monitored, processed, and
stored in a computer by a LabVIEW® code. In addition, the prototype
rotation speed (N ) is checked manually with a digital tachometer
model MDT-2238A, fabricated by Minipa, with an accuracy of 0.05%.
As the fluids are flowing in a closed loop, they tend to heat up
during the experiments. The temperature-control system, composed of
the heat exchanger and a thermo-chiller, is able to maintain the fluids at
a constant temperature. The system can also heat or cool the liquids; it
is thus possible to modify their properties during the tests, especially Fig. 1. Photograph of the experimental facility with the visualization proto-
the viscosity. type.

884
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Then, the oil injection and the image acquisition restart. The procedure
is repeated for all water flow rates. Finally, the test is performed again,
at a different prototype rotation speed.

3. Image analysis

The visualization tests for liquid-liquid two-phase flow resulted in


thousands of images. An observation of them indicates the occurrence
of a flow pattern characterized by dispersed oil drops in a water flow.
The image processing reveals the characteristics of the oil drops, as well
as allows the analysis of their dynamics, with an evaluation of velo-
cities, accelerations, and forces.

3.1. Image observation


Fig. 2. Schematic diagram of the experimental facility.
The flow images confirm the presence of dispersions within the
impeller. The dispersed oil drops in water occur at all operational points
Table 1 considered in the test matrix, constituting a unique flow pattern. In
Test matrix with six flow conditions.
addition, most drops have a spherical geometric shape, with only a few
N [rpm] 0.8 QBEP [m³/h] 1.0 QBEP [m³/h] 1.2 QBEP [m³/h] presenting irregular formats. The oil drops have equivalent diameters of
tenths of a millimeter to about 3 mm. Furthermore, the drop size di-
600 1.70 2.13 2.56
900 2.56 3.20 3.84
rectly depends on the ESP rotation and the water flow, as shown in
Fig. 3, Fig. 4, and Fig. 5, which compare the impeller under different
tested conditions. Such qualitative results are similar to the observa-
condition, the flow rate is below the maximum and the pump head is tions by Perissinotto et al. (2018).
above zero. The procedure is repeated until the water flow rate reaches Clearly, the oil drop size varies as a function of the flow char-
a minimum value that correlates with the ESP highest-pressure, a acteristics. The drops become smaller as pump rotation and water flow
condition called shut-off. Then, the prototype rotation is updated and rate increase. The phenomenon is a consequence of higher local
the test is carried out from the beginning. stresses, turbulence, and centrifugal field acting on the oil drops.
The procedure adopted for the two-phase flow performance tests is Breakage and coalescence are virtually not observed in the image
similar to the previous one. The only difference is the injection of oil at analysis. In fact, the oil drops tend to break before the impeller inlet, on
a constant flow rate. As the water flow decreases, the oil fraction in- the way from the injection point to the channel inlet, where flow re-
creases. strictions cause a high localized shear and the dispersion suffers a
However, the visualization experiments for two-phase oil-in-water sudden change of direction.
flow have a quite different procedure. Before the test, the position of the
light reflectors is carefully adjusted to avoid forming shadows; the 3.2. Image processing
camera is placed on a tripod and the lens is focused on the impeller. The
desired ESP rotation is set and remains constant throughout the test. The images were processed in a computational environment in
Initially, the water pump and the discharge control valve are operated order to enable the analysis of the oil drops trajectories and the eva-
to obtain the desired water flow rate. In the oil injection system, the luation of their dynamics. The procedure is based on that adopted by
control valves are opened and the peristaltic pump is turned on. The Perissinotto et al. (2018) for the investigation of oil drops velocity using
high-speed camera then captures flow images for 10 s. While the images a Particle Tracking Velocimetry (PTV) approach.
are saved in the computer hard disk, the peristaltic pump is turned off A MATLAB® routine firstly rotated each image counterclockwise. As
and the oil control valves are closed. After a few minutes, when the a result, the channels became static, as if the high-speed camera rotated
storage is complete, a new condition is established in the water loop. integrally with the impeller during the experiments. The reverse rota-
tion technique facilitates the analysis of the oil drops inside each

Fig. 3. Oil drops in the impeller at QBEP : (a) 600 rpm and (b) 900 rpm.

885
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Fig. 4. Oil drops in the impeller at 600 rpm: (a) 0.8 QBEP and (b) 1.2 QBEP .

channel without the impeller angular motion. Then, using the software
IDT Motion Studio, samples of oil drops were tracked from the images
of the ESP prototype operating at the test matrix conditions, as de-
scribed in Table 1. The selected drops have diameters approximately
from 1.0 mm to 2.5 mm. They execute well-behaved trajectories within
the channels.
The tracking process creates a table with the oil drop position as a
function of time, in a Cartesian coordinate system. The position, x and
y , is converted to a new position, R and θ , in the directions îR and îθ , in
polar coordinates with origin in the center of the impeller, as Fig. 6
illustrates. The frame of reference is non-inertial since it rotates along
with the impeller, due to the reverse rotation technique explained
above.
The R and θ values are used to calculate the radial and swirl velo-
cities, VR and Vθ , and the radial and swirl accelerations, AR and Aθ . In a
Lagrangian approach, they are obtained through the derivatives and the
chain rule. As Hibbeler (2010) explained, considering the oil drops as
rigid particles of negligible size, the velocity and acceleration are de- Fig. 6. Position of oil drop in a polar coordinate system.
termined as:

V = VR îR + Vθ îθ = R˙ îR + Rθ˙ îθ (3.1) to White (2011), the procedure consists of including effects such as the
→ →
tangential velocity due to the rotation, ω × R , and the centrifugal and
→ 2 → → → → →
A = AR îR + Aθ îθ = (R¨ − Rθ˙ ) îR + (Rθ¨ + 2Rθ
˙ ˙ ) îθ (3.2) Coriolis pseudo-accelerations, ω × ( ω × R ) and 2 ω × V . Therefore,
→ ⎯→

→ → the inertial velocity and acceleration, VI and AI , are evaluated as:
The particle velocity, V , and acceleration, A , are then mathemati-
cally converted from the rotating non-inertial frame of reference to a →
VI = R˙ îR + (Rθ˙ − ωR) îθ (3.3)
new fixed inertial one, with origin still in the impeller center. According

Fig. 5. Oil drops in the impeller at 900 rpm: (a) 0.8 QBEP and (b) 1.2 QBEP .

886
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

⎯→
⎯ Considering an oil drop with a spherical shape, the mass is esti-
AI = [R¨ − (ω + θ˙ )2R]îR + [Rθ¨ + 2(ω + θ˙ ) R˙ ] îθ (3.4)
mated as a function of the oil density, ρo , the drop diameter, d , and the
The number of dots on R and θ represent their first and second ⎯→
⎯ ⎯→

drop acceleration, AI . Hence, the resultant force, FR , in the inertial
derivatives in time, i.e., dR/ dt , dθ / dt , d ²R/ dt ² and d ²θ / dt ² . Since the oil
stationary frame of reference may be written as:
drop position is given in the form of a matrix, the derivatives should be
calculated via numerical derivation. Hence, the derivatives are ob- ⎯→
⎯ πd3 ⎯→

FR = ρo AI
tained by the central finite-difference method, with a step, h , in the 6 (4.2)
order of the time interval between two consecutive images, Δt . The ⎯→

finite-difference method is described by the equations below, where The drag force, FD , is a consequence of the relative velocity be-
f ′ (t ) and f ′′ (t ) are the derivatives of a generic function f (t ) . In this tween continuous and dispersed phases. The force is described as fol-
paper, the function f (t ) represents R and θ : lows:

f (t + h) − f (t − h) ⎯→
⎯ 1 πd 2 → → → →
f ' (t ) ≈ FD = ρw CD (U − V ) U − V
(3.5) 2 4 (4.3)
2h
f (t + h) − 2f (t ) + f (t − h) where ρw is the water density, CD is the drag coefficient, d is the oil drop
f '' (t ) ≈ → →
h2 (3.6) diameter, U and V are respectively the water and oil drop velocities.
There are many correlations to calculate the drag coefficient in the
The calculation of velocities and accelerations is related to un-
literature. The drag coefficient model applied in this work is the
certainties. As stated above, the tracking of an oil drop consists of de-
Spherical Drag Law by Morsi and Alexander (1972).
termining the position of its center, x and y , on each processed image. ⎯→

The force due to a pressure gradient, FP , is calculated as follows:
This position is obtained automatically by the software, but can be
corrected manually by the user. Specifying the position of an oil drop's ⎯→
⎯ πd3
FP = − ∇s p
center is a purely visual task; it thus may contain uncertainties in the 6 (4.4)
order of one pixel, which corresponds to one-tenth of a millimeter in the
where d is the oil drop diameter and ∇s p is the pressure gradient along
images captured during the experiments. This 0.1 mm uncertainty is
the oil drop path.
propagated in the calculation of R , θ , and their derivatives, as well as in
→ → In the case of gas-liquid systems, Minemura and Murakami (1980)
the determination of velocities, VR , Vθ , V , VI , and accelerations, AR , Aθ ,
→ ⎯→ ⎯ and Crowe et al. (2012) state that drag and pressure forces are the most
A , AI . relevant to the dispersed phase motion. This may be attributed to the
The uncertainties associated with these variables can be evaluated low-density ratio between dispersed and continuous fluids. However,
using the information available in Moffat (1988) and Feng et al. (2011). the same is not necessarily true for liquid-liquid systems, and other
It is noted that, in this current paper, the major cause of increased forces may not be neglected at first.
uncertainties is the time interval Δt adopted as the step h in the finite- ⎯⎯⎯⎯⎯→
The virtual mass force, FVM , generated by the acceleration of the
difference method of Equation (3.5) and Equation (3.6). This step has a
water near the oil drop, is defined as follows:
value around 0.001 s, as long as the images are captured at a 1000
frames per second rate, approximately. → →
⎯⎯⎯⎯⎯→ πd3 ⎛ DU dV ⎞
In order to reduce the uncertainties propagation, the velocities and FVM = ρw CVM ⎜ −
6 Dt dt ⎟ (4.5)
accelerations are corrected using a central moving average technique. ⎝ ⎠
This method, which consists of calculating a simple mean of previous where ρw is the water density, CVM = 0.5 is the virtual mass coefficient
and next data in the table, has the interesting advantage of reducing → →
for spherical drops, d is the oil drop diameter, U and V are respectively
fluctuations and oscillations in curves and graphs. In this paper, a
the water and oil drop velocities.
moving average method with five elements was adopted, i.e., the cor- ⎯→

Finally, the Saffman's lift force, FL , caused by a pressure distribu-
rected value g (t )avg depends on five original adjacent values. The gen-
eric function g (t ) represents a velocity or acceleration: tion developed on the oil drop in a velocity gradient, may be calculated
as follows:
g (t + 2h) + g (t + h) + g (t ) + g (t − h) + g (t − 2h)
g (t )avg = 1/2
5 (3.7) ⎯→
⎯ πd 2 2 K νw dij → →
FL = ρw (U − V )
6 (dlk dkl )1/4 (4.6)
Relative uncertainties were evaluated; the calculations yielded va-
lues around 10% in the case of velocities, and 60% in the case of ac- where ρw is the water density, d is the oil drop diameter, K = 2.594 is a
celerations. However, as stated above, the uncertainties decrease when constant, νw is the water kinematic viscosity, dij is the deformation
Δt increases. → →
tensor in the Einstein's notation, and U and V are respectively the
Finally, with the velocities and accelerations, a resultant force can ⎯→

water and oil drop velocities. This expression for FL is employed in
be determined. The next section describes the mathematical modeling
ANSYS® Fluent v16, the CFD software used to perform the numerical
of the main forces that act on oil drops in oil-in-water dispersions.
simulations in this current paper (Section 5).
Equation (4.6) was developed by Li and Ahmadi (1992) as a gen-
4. Drop motion
eralization of the original equation provided by Saffman (1965). It is
valid for small particle sizes, normally smaller than one μm, with the
The oil drop dynamics is investigated using a force balance on the
⎯→
⎯ particle Reynolds number much lower than the unit. In the present
fluid particle. The resultant force, FR , that governs the motion of an oil
→ work, the lift force is disregarded, since the oil drop diameters take
drop with a mass mo and a velocity VI , is considered here as a sum of values in the order of millimeters.
⎯→
⎯ ⎯→
⎯ ⎯→

drag force, FD , pressure force, FP , lift force, FL , and virtual mass force, To simplify the analysis, other forces described by Crowe et al.
⎯⎯⎯⎯⎯→
FVM . From Newton's second law, with a Lagrangian approach, and (2012) were neglected in this study, such as Magnus lift, Basset's term,
⎯→

using an inertial fixed frame of reference, the resultant force, FR , is and forces related to particle-particle and particle-wall interactions.
⎯→
⎯ ⎯→ ⎯
defined as: These forces usually have little importance when compared to FD , FP ,
⎯→
⎯ ⎯⎯⎯⎯⎯→
→ FL , FVM . If more forces were considered, the analysis would become
⎯→
⎯ dV ⎯→
⎯ ⎯→
⎯ ⎯→
⎯ ⎯⎯⎯⎯⎯→
FR = mo I = FD + FP + FL + FVM unnecessarily complex and there would not be sufficient information to
dt (4.1)
evaluate all the terms.

887
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

⎯→
⎯ ⎯→⎯
Unfortunately, it is not possible to calculate the values of FD , FP , and adjusted according to each case. The outlet was set to Pressure-
⎯→
⎯ ⎯⎯⎯⎯⎯→ → Outlet with gauge pressure 0 Pa. The SIMPLE scheme was used for
FL , FVM directly from the experiments, because the water velocity, U ,
and the pressure gradient, ∇s p, are both unknown. A CFD numerical pressure-velocity coupling and a second-order upwind scheme was
simulation is thus an interesting tool to further investigate the oil dy- chosen for spatial discretization. Convergence was attained when all
namics. In a future experimental approach, it would be a good choice to residuals were below 10−4.
use Particle Image Velocimetry (PIV), with the purpose of studying the Once the flow solution had been obtained, a Lagrangian particle
water flow. tracking was performed in the fixed velocity field. In this case, the one-
CFD simulations can assess the velocities, accelerations, and forces way coupling is justifiable on the basis of the very low volume fraction
involved in the movement of oil drops. Mass and momentum equations of the disperse phase. Thus, a mono-dispersed distribution of oil drops
are solved for the single-phase flow and a Discrete Phase Model (DPM) was injected at the inlet surface with the same velocity magnitude as
is responsible for reproducing the dispersed phase behavior. that of water. Equation (4.1) was integrated numerically to obtain the
Assuming steady-state, incompressible, and turbulent flow, with an position as a function of time of tracking. A trapezoidal scheme was
eddy viscosity model, the Reynolds-Averaged Navier-Stokes (RANS) employed, using an adaptive step such that a relative tolerance of 10−5
continuity and momentum equations take the following form in the was respected. The particles were tracked at a maximum of 5000 steps,
stationary reference frame, using the Einstein's notation: sufficient for at least 90% of oil drops to leave the system. The drops
were assumed to maintain a constant diameter (10 mm, 1 mm, 0.1 mm,
∂Ui or 0.01 mm) in the complete duration of the simulation, without
=0
∂x i (4.7) breaking up, which is consistent with experimental observations from
Section 3.1.
∂U ∂P ∂ ⎛ ⎡ ∂Ui ∂Uj ⎤ ⎞ An automated routine was finally implemented to acquire and
ρw ⎛⎜Uj i ⎟⎞ = − + μ +
⎝ ∂x j ⎠ ∂x i ∂x j ⎜ ef ⎢
⎣ ∂x j ∂x i ⎥
⎦⎠

(4.8) process data relative to particle ID, position, velocity, turbulent eddy

dissipation, and drag and pressure forces.
The water velocity, U , is evaluated in the inertial frame. The vari-
able t indicates the time and x represents the position. Water density,
6. Results
ρw , and effective viscosity, μef , characterize the continuous phase. The
modified pressure, P , which combines all contributions to normal
The experimental procedure provided quantitative results with re-
stresses and incorporates the dynamic pressure term, is defined as
2 spect to the pump performance and the oil drops motion. Similarly, the
P = p + 3 ρw κ δij , where p symbolizes the static pressure, ρw defines the
numerical simulations produced results concerning the prototype per-
water density, κ indicates the turbulent kinetic energy, and δij is the
formance, the water flow field, and the oil drops behavior. In the next
Kronecker's delta.
sections, trajectories, turbulent dissipations, residence times, velocities,
accelerations, and forces are presented and both methods are com-
5. Numerical methodology
pared.

This section describes the methodology adopted to perform the


6.1. Mesh sensitivity
numerical simulations. The ANSYS® Fluent v16 software was used to
simulate the same operational conditions as defined in the test matrix.
The prototype geometry is originally not axisymmetric, due to dif-
The position, velocity, acceleration, turbulent dissipation, and the
ferences between the number of impeller and diffuser blades. In the RG
forces acting on the oil drops were numerically evaluated.
model, the diffuser regions were modified to present the same number
of blades as the impeller, i.e., 7 instead of 8 blades, rendering the ESP
5.1. Computational model
an axisymmetric geometry. This was carried out to ensure an efficient
mesh-sensitivity analysis, in terms of computational costs, which is
Based on the original Baker-Hughes P23 ESP, two computational
important for verifying the mesh independence of results. Four meshes
models were designed for the visualization prototype: The Full
were generated, following the characteristics displayed in Fig. 8. The
Geometry (FG), which reproduces with fidelity the real pump-based
mesh sensitivity analysis was performed for a rotation N = 3500 rpm at
prototype, and the Reduced Geometry (RG), which is used for the mesh
a flow rate QBEP = 15.0 m3/ h .
sensitivity analysis with much lower computational costs. Both geo-
The result is presented in Fig. 9, which shows the pump head con-
metries, shown in Fig. 7, are composed of two main parts: the impeller
vergence plotted as a function of the number of cells. The solution
zone, which has a rotational movement, reproduced by a Moving Re-
becomes independent when the number of cells is approximately 530 k
ference Frame (MRF), and a stationary zone, corresponding to the dif-
and the pump head relative error is less than 3%, using as reference the
fuser. Although a Sliding Mesh Model (SMM) could provide higher fi-
solution obtained for a mesh with 890 k cells. Compared with experi-
delity results by taking into account the transient effects, the Moving
mental data (Biazussi, 2014), the pump head variation is around 12%.
Reference Frame (MRF) approach was found to provide acceptable re-
This mesh presents an average dimensionless distance from the wall of
sults for the present purposes at a much lower computational cost.
y+ ∼ 10 ; as mentioned in Section 5, this is acceptable for the chosen
The realizable κ-ε turbulence model was chosen with a scalable wall
function, which assured that results were consistent even in the case of
a dimensionless distance from the wall of y+ < 11. The model solves for
two additional transport equations to determine the turbulent kinetic
energy (κ ) and its dissipation (ε ), which are needed for calculating the
effective viscosity, μef . Further details about the model formulation may
be taken from ANSYS Inc (2015).

5.2. Simulations: water and oil drops

Six simulations were prepared and performed, in accordance with


the test matrix presented in Table 1. All cases were simulated in ANSYS® Fig. 7. Computational models for the ESP: (a) Reduced Geometry (RG) and (b)
Fluent v16 software. The inlet condition was set as Mass Flow Average Full Geometry (FG).

888
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Fig. 8. Mesh grid of (a) impeller zone and (b) details of prisms layers near the blades.

P23 ESP. Hence, the modifications applied to the ESP to obtain the
prototype for flow visualization did not significantly influence the ori-
ginal pump characteristics, thus the dynamic similarity was maintained
and the flow fields in the impeller are fully representative.

6.3. Flow field

Detailed insight into the flow field within the pump may be ob-
tained through CFD simulations. Velocity contour was plotted in a
plane parallel to the visualization area, z = −3 mm , i.e., situated 3 mm
below the top window. This plane contains the impeller zone and the
diffuser zone. Results are presented in Fig. 11 for the rotation speed of
N = 600 rpm . According to Monte Verde et al. (2017), the blade height,
Fig. 9. Influence of grid size on simulated ESP head. which corresponds to the distance between the transparent top window
and the white impeller bottom, is 6 mm.
wall function formulation. As can be seen, Fig. 11 presents the velocity contours in an inertial
frame, for 600 rpm, at the BEP. For this condition, the jets formed at the
6.2. Pump performance diffuser region are slightly asymmetric and are located mostly under the
diffuser blade. High-speed areas are seen at the impeller tips. The result
Numerical and experimental results for single-phase pump perfor- suggests that the impeller-discharged oil drops that follow through the
mance curves were obtained using the dimensionless flow rate (Φ ) and suction diffuser region may experience higher acceleration than other
head (Ψ ) coefficient definitions: oil drops.
Motivated by the importance of models that describe the oil drop
Qw
Φ= breakage events, a second analysis was conducted: the evaluation of the
ωD3 (6.1)
flow rate effect on the turbulent energy dissipation rate (ε ), normalized
ΔP by N 3D 2 , for both rotational speeds of 600 rpm and 900 rpm. Results
Ψ= can be found in Fig. 12, which introduces the volumetric average of the
ρω2D 2 (6.2)
dimensionless turbulent energy dissipation rate for the impeller and
where Q w represents the water flow rate, ΔP symbolizes the difference diffuser regions, as a function of the flow rate. It can be observed that
between the outlet and inlet pressures, ω designates the impeller an- the 600 rpm and 900 rpm curves for each region collapse, a fact that
gular speed, and D corresponds to the impeller diameter. suggests the normalization is appropriate and ε ∼ N 3D 2 indeed.
Curves for the dimensionless head (Ψ ), as a function of the di- The regions clearly present distinct behavior. The energy dissipated
mensionless flow rate (Φ ), are exposed in Fig. 10. The results were in the impeller region is influenced by the flow rate. A good fit for the
obtained from experimental measurements and calculations via nu- dependence is described below, where q is the fraction of QBEP :
merical simulations. As can be observed, CFD points and experimental
ε
curves present a satisfactory agreement, which indicates that the = 0.3567 e1.6822 q
N 3D 2 (6.3)
methodology adopted to carry out the simulations can predict very well
the real flow inside the prototype pump.
The experimental head curves of the prototype operating with the
oil-in-water dispersion suggest that the oil injection does not sig-
nificantly modify the ESP performance. For the operational conditions
investigated in this study, the curves of two-phase oil-water flow
showed excellent agreement with the curves of single-phase water flow,
with deviations less than 2% in the pressure increment. This difference
is smaller than the experimental uncertainties. Evidently, for larger oil
fractions, instabilities, phase inversion, and performance degradation
are expected; however, in this research, results indicate that the addi-
tion of 2 ml/s of oil does not cause performance losses to the prototype.
Indeed, the oil flow rate is very low, with relative values around 1%
when compared to the water flow rates at the best efficiency points.
Monte Verde et al. (2017) also executed single-phase flow experi-
ments and concluded that the performance curves for the visualization Fig. 10. Experimental and numerical performance of the ESP prototype oper-
prototype are quite similar to the curves for the original Baker-Hughes ating with a single-phase flow.

889
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Fig. 11. Velocity contour curves at a xy plane (z = −3 mm ) for 600 rpm simulation at 1.0 QBEP .

dissipation. Qualitative comparisons of this field are exposed in Fig. 13:


(a–b) for different rotation speeds, at a constant 1.0 QBEP condition, and
(c–d) for various flow rates, at a constant impeller rotational speed
N = 600 rpm .
In Fig. 13(a–b), flows are similar for both rotational speeds ana-
lyzed, which means not only the average of energy dissipation but also
the field itself scales as ε ∼ N 3D 2 . However, in Fig. 13(c–d), the high
dissipation zone, located near the impeller blades, tends to expand with
the flow rate rise, especially next to the leading edge. It explains the
exponential increase in the curve presented in Fig. 12.
It is expected that the oil drops crossing zones of high dissipation,
indicated in Fig. 13, are exposed to intense turbulence, which may re-
sult in a higher probability of breakage events and consequently smaller
drop sizes. Notwithstanding, the experimental observations (Section
3.1) suggest the occurrence of breakage events mostly before the im-
peller zone. This is better illustrated in Fig. 14, which presents the
contour of normalized turbulent energy dissipation rates plotted on an
axial plane, y = 0 .
Fig. 12. Dimensionless turbulent energy dissipation rate calculated as a func- As Fig. 14 reveals, the prototype geometry has a contraction fol-
tion of the flow rate. lowed by a flow direction change of 90° at the impeller zone entrance.
In this region, high turbulent energy dissipation rates are observed.
On the other hand, the energy dissipation in the diffuser region is Therefore, there are two main zones of high-intensity turbulence: in the
nearly constant, independent of the flow rate. These results can be impeller zone entrance (the most significant) and inside the impeller
explained by observing the contours of the normalized turbulent eddy channels. The result corroborates with the experimental observations

Fig. 13. Contour of normalized turbulent dissipation: at 1.0 QBEP condition (a) for 600 rpm and (b) 900 rpm; fixing the impeller rotational speed at 600 rpm (c) for
0.8 QBEP and (d) 1.2 QBEP .

890
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

channel with the largest radial and transverse displacements.


The oil drop trajectories are illustrated in Fig. 15: (a–c) with the
experimental paths observed in the processed flow images, and (d–f)
showing the numerical reproduction of an oil-tracking experiment. The
results embrace simulations at N = 600 rpm and Q w = 1.0 QBEP condi-
tion for oil drops with a diameter d = 1 mm .
To allow a systematic comparison between experimental and nu-
merical results, the characteristic paths are translated to a set of em-
pirical mathematical criteria. First, the oil drop polar position inside the
impeller, R and θ , is normalized according to new radial and transverse
Fig. 14. Contour of normalized turbulent dissipation at an axial plane ( y = 0 ).
and turbulent eddy dissipation ε calculated from numerical simulation. terms, R∗ and θ∗, as the following definitions:
R − Ri
R* =
and implies that, in the present cases, the oil drop breaks up before Ro − Ri (6.4)
entering the impeller channel. In the narrow gap region, turbulence
intensity is largely affected by the flow rate, implying an important
θ* =
θ + A ln ( )+θ
R
Ro 0

effect that controls the drop breakage process and the size distribution θc (6.5)
in the pump. These phenomena will be further analyzed in a future
work. In the equations, Ri and Ro are the impeller radius from the origin to
the inlet and to the outlet respectively, that is, Ri = 0.027 m and
Ro = 0.056 m , according to Monte Verde et al. (2017); A = −1.35 is a
6.4. Drop trajectory constant that depends on the blade geometry; θ0 is the drop's initial
angle in relation to the x axis, a parameter dependent on the impeller
An analysis of the processed images reveals that, in general, most oil position; θc is the channel angle, defined as:
drops follow random trajectories in the channels. Many drops undergo

lateral deviations in their paths, while others move in a parallel fashion θc =
Nc (6.6)
to the blades. Some oil drops remain close to a blade, while others move
away from it. where Nc = 7 represents the number of channels.
Despite the randomness, three trajectory patterns were easily Then, with R∗ and θ∗ values, three regions can be defined as de-
identified, and an oil drop can be classified into three categories, ac- scribed in Table 2, formulated to encompass the experimental classifi-
cording to its trajectory: (a) oil drop with a central path, (b) oil drop cation itself. In case of θ∗ < 0.4 , the region is classified as suction, while
near a suction blade, and (c) oil drop near a pressure blade. These re- θ∗ > 0.6 characterizes the pressure zone. Finally, the central path is
sults agree with Perissinotto et al. (2018). described from two position functions, f + (R∗) and f − (R∗) . A trajectory
Only a few drops follow the suction or pressure paths. Usually, these is classified as belonging to a determined path if the oil drop crosses the
patterns occur when an oil drop is very close to the blade, almost sliding entire impeller zone without leaving its defined region.
over it. In this situation, the particle-wall interaction is stronger than As already explained, the image observation from the experiments
the particle-fluid interaction. As a result, the turbulent flow, full of (Section 3) revealed that most oil drops present random trajectories,
vortices, jets, and recirculation, cannot influence the oil drop trajectory. preventing them from being classified into any of the regions. Around
On the other hand, many drops present the central path pattern. 25% of oil drops perform well-behaved trajectories within the channels.
They enter the impeller next to the suction blade but gradually deviate Regarding the numerical simulations (Section 5), from 103 particles
towards the pressure blade, where they finish the trajectory and exit the injected in the impeller, at least 30% passed through the classified re-
impeller. The result is an approximately curved diagonal track. The gions and could be accounted for. Therefore, the numerical procedures
central-path drops perform a quite long trajectory, since they cross the resulted in a satisfactory qualitative agreement with the experimental

Fig. 15. Experimental observation of three patterns for an oil drop path: (a) central, (b) on suction blade, (c) on pressure blade; and numerical reproduction of paths:
(d) central, (e) on suction blade, (f) on pressure blade.

891
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Table 2
Oil drop paths: description of criteria and regions.
Criteria Regions

Suction blade region


θ∗ < 0.4

Pressure blade region


θ∗ > 0.6

Central path
(θ∗ < f + (R∗)) ∧ (θ∗ > (f − (R∗)) f − (R∗) = 0.5R∗ f + (R∗) = 0.5R∗ + 0.5

visualizations. the mean of each variable, φ , is calculated with:


The next section discusses the experimental and numerical results σ
concerning velocities, accelerations, and forces related to the oil drop Δφ = t (0.95, n − 1)
n (6.7)
motion.
where n is the sample size at each radial position and t (0.95, n − 1) are the
6.5. Drop dynamics values for a t-distribution assuming a confidence interval of 95%. Both
analyses were performed using the non-inertial reference frame.
A numerical analysis of velocity profiles along the oil drops trajec- The velocity components of oil drops have magnitudes of up to
tories at the central region was conducted. The results can be now 2.0 m/s, while the acceleration components take very high values, in
compared to the experimental calculations (Section 3.2). The average of the order of hundreds of m/s2, from −150 to 200 m/s2. Both experi-
radial and swirl velocities, VR and Vθ , for 44 oil drops with a constant mental and numerical results are directly dependent on the flow con-
diameter of d = 1 mm were obtained at 600 rpm and 1.0 QBEP condition, ditions. Moreover, both the velocity and acceleration magnitudes pre-
using experimental and numerical approaches. The results are pre- sent their peak values around the inner radius. The effect of virtual
sented in Fig. 16. mass (VM) is small and most relevant in the entry region for the radial
Radial and swirl accelerations, AR and Aθ , were also calculated at component of acceleration.
600 rpm and 1.0 QBEP condition and the experimental and numerical In general, the numerical and experimental results present sa-
results are shown in Fig. 17. The acceleration plots also present the tisfactory agreement for all cases. Deviations are expected due to the
profiles obtained through a simulation that neglects the virtual mass use of the MRF approach, which is not capable of capturing inherent
term (VM). In total, 31 oil drops with a constant diameter of d = 1 mm unsteadiness. As a frozen-impeller approach, the MRF may give ac-
were analyzed. ceptable approximations when the interaction between moving and
Each variable (VR , Vθ , AR Aθ ) was interpolated to a fixed, fine, grid of stationary parts is not strong and the flow at the interface between
radial positions, and simple arithmetic averages with corresponding rotating and fixed regions does not present complicated features.
standard deviations, σ , were computed between individual values. Furthermore, it may be an interesting tool with which to evaluate the
These values correspond to individual drops, for the experiments, or influence of different parameters on the overall flow behavior, as per-
different injections, for the simulations. The confidence band, Δφ , for formed here (e.g., effect of flow rate and rotational speed on different

Fig. 16. Average velocity components at 600 rpm and 1.0 QBEP condition for 44 drops with d = 1 mm : (a) radial velocity and (b) swirl velocity.

892
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Fig. 17. Average acceleration components at 600 rpm and 1.0 QBEP condition for 31 drops with d = 1 mm : (a) radial acceleration and (b) swirl acceleration.

variables). Differences between the experimental and numerical results similar normalized residence time for all cases. This observation may be
may also be attributed to the assumption of a mono-dispersed dis- expected since small oil drops tend to follow the water streamlines.
tribution. In reality, there is a poly-dispersed distribution of drops and However, as oil drop sizes increase, residence times at the impeller zone
the interaction forces depend on the drop size, which in turn affect the also increase.
trajectories. These effects can be especially important when strong ac- Surprisingly, oil drops with d = 1 mm (Red = 2976.8) demonstrated
celerations are present, such as in the impeller inlet region, where higher residence times than oil drops with d = 10 mm (Red = 29768),
larger deviations are seen in Fig. 16 between the curves. except for in the 0.8 QBEP case. This fact can be explained by the ob-
As noted in the Section 4, the forces acting on the oil drops are due servation that there is a significant decrease in the number of drops
to drag, pressure gradient, and virtual mass. Vectors of pressure gra- with d = 10 mm captured at the central region in the numerical simu-
dient were plotted in a plane parallel to the visualization area, lations, as Table 3 shows. As this is the longest path out of the three
z = −3 mm , 3 mm below the top window. The results are presented in categories, a lower number of oil drops leads to a reduced residence
Fig. 18. Also mentioned before is the fact that the pressure force acting time. For the 0.8 QBEP condition, however, at both impeller rotation
on the drop is proportional to the pressure gradient, in the opposite speeds, this effect seems to be balanced out by the presence of re-
direction. circulation zones, which cause some oil drops to deviate and their re-
The vectors in Fig. 18 allow an assessment of the effect of the sidence time to be increased.
pressure gradient force on the drop trajectories in the impeller channel. The residence time reflects the time that the oil drops are exposed to
When an oil drop moves near the suction region, it comes across a high breakage and coalescence events; it is thus very relevant to understand
adverse pressure gradient that causes it to divert its trajectory toward the dynamics of the oil motion. In addition, this knowledge can be
the central region. On the other hand, when an oil drop moves near the important for scaling up pump operations.
pressure blade, it also suffers influence from an adverse pressure gra- To help illustrate how the different forces may impact the oil drop
dient and tends to deviate in the direction of the pressure blade wall. motion, Fig. 21 presents a water streamline and trajectories of oil drops
These numerical results explain why the oil drops preferably perform of various sizes. The injection was performed through a single point in
trajectories in the central region rather than close to the blades. the numerical simulations (Section 5.2). Confirming previous results,
Displayed in Fig. 19 is a comparison between the magnitudes of small drops tend to follow the streamline. On the other hand, oil drops
drag and pressure forces, along with the resulting acceleration in the with d > 0.1 mm show a deviation away from the same streamline. This
inertial frame. The analysis is restricted to the central region of the fact suggests that oil drops with typical sizes as the ones measured in
impeller, with the normalized radius, R∗, ranging from 0.3 to 0.7. In the the experiments are significantly influenced by external forces.
present case, the pressure force is the most important force acting on In future investigations, the measurement of the drop size dis-
the oil drops, being about 10 times larger than the drag effects. Yet, in tributions combining position, velocity, and acceleration would provide
the impeller entry region, other forces, such as drag and virtual mass, interesting insights, as well as the possibility to quantify the joint
may contribute towards decelerating the oil drops. It is important to probability distributions in the different regions of the pump, for ex-
note, however, that these effects scale differently with drop size, such ample. Size-velocity correlations could also support the understanding
that this trend may be reversed in the case of fine dispersions. of dominant physical mechanisms behind the oil drop motion in the
Results for the oil drop tracking at 600 rpm and 900 rpm for 1.0 QBEP impeller channels.
conditions are compared using two global parameters: the normalized

impeller residence time, tres , and the particle Reynolds number, Red ,
both defined as follows:

* Qw
tres = tres
D3 (6.8)

ρo d Utip
Red =
μw (6.9)

where tres is the impeller residence time, defined as the time spent by an
oil drop moving from the inlet to the outlet of the impeller zone. Red is
based on the blade velocity, Utip , and also depends on the water visc-
osity, μ w . As mentioned before, Q w represents the water flow rate, D
describes the impeller diameter, ρo corresponds to the oil density, and d
symbolizes the drop diameter.

Results for tres as a function of Red are shown in Fig. 20 for all the
conditions displayed in Table 1. Oil drops with d = 0.01 mm and Fig. 18. Vectors of pressure gradient plotted on xy plane (z = −3 mm ) for
d = 0.1 mm , corresponding to Red = 29.768 and Red = 297.68, have 600 rpm at 1.0 QBEP condition.

893
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

Fig. 19. Average drag force, pressure force, and acceleration in inertial frame for oil drops with d = 1 mm : (a) central region at 600 rpm and (b) central region at
900 rpm.

7. Conclusions

This work carried out experimental and numerical analyses on the


motion of individual oil drops dispersed in a water flow within an ESP
impeller.
Experimental tests were performed using a facility with an ESP
prototype, the transparent top shroud of which enabled visualization of
the flow in all the channels. The images revealed, in all operational
conditions, a unique flow pattern within the impeller: a dispersion of oil
drops in a continuous water phase. Most oil drops presented spherical
geometry and only a few had irregular shapes. Almost no breakage and
coalescence were observed within the channels. The oil drop diameter
varied from approximately tenths of a millimeter to 3 mm. The images
suggested a dependence between the oil drop size and the operational
conditions. As the impeller rotation speed and the water flow rate in-
creased, the oil drops became smaller.
∗ Oil drops that follow central trajectories in the channels were in-
Fig. 20. Normalized impeller residence time tres as function of Red for 600 rpm
and 900 rpm.
vestigated with respect to the dynamics of their motion. This motion is
mainly governed by phenomena classified as particle-fluid interaction,
in which three forces are in play: drag force, pressure gradient force,
Table 3 and virtual mass force. Flow images were processed for velocity and
Number of oil drops with different diameters captured in the central region.
acceleration calculations. The oil drops presented velocities with a
N [rpm] d [mm] Number of oil drops in central region (numerical simulations) magnitude of units of m/s, while accelerations reached hundreds of m/
s2. To verify the methodology, the measured pump performance was
at 0.8 QBEP at 1.0 QBEP at 1.2 QBEP
compared with previous single-phase experiments. The comparison
600 0.01 27 33 35 revealed that the low oil flow rate had very little effect on the pump
0.1 32 27 36 pressure gain.
1 34 36 34 A numerical procedure was also adopted to determine the drop
10 15 20 22
dynamics. The drop trajectory classes were defined in terms of nor-
900 0.01 23 32 37
0.1 28 27 37 malized coordinates inside the impeller, which allowed for a systematic
1 21 31 22 comparison between the results. The qualitative behavior of the paths
10 15 19 20 was shown to be well reproduced by the adopted methodology.
Comparisons of predicted and measured velocity and acceleration
components along the radial position were made, deeming the agree-
ment acceptable, given the limitations in the modeling.
An evaluation of the main forces involved in the oil drop motion
indicated that the dominant force was that of pressure gradient, and its
numerical values were about one order of magnitude larger than those
of drag force. Virtual mass forces were observed to have some influence
only at the impeller entrance. Finally, the evaluation of global para-
meters, such as impeller residence time, single oil drop trajectories, and
static pressure provided an overview of how external forces act on oil
drops. Small drops follow the water streamlines while oil drops with
large diameters are affected by external forces in different ways and
deviate from the continuous phase motion. In general, the oil drops
suffer the influence of an adverse pressure gradient and tend to deviate
Fig. 21. Water streamline and trajectories of single drops with different dia- from the suction blade to the pressure blade, a situation that facilitates
meters at 600 rpm. the occurrence of central trajectories.
This study aims to open the way for greater phenomenological un-
derstanding and should be able to facilitate the creation of simplified

894
R.M. Perissinotto et al. Journal of Petroleum Science and Engineering 175 (2019) 881–895

models that may help the design and operation of ESP devices in some Flow Operation. PhD dissertation. University of Campinas, Sao Paulo, Brazil.
challenging conditions. Such a framework could enable, for example, a Feng, Y., Goree, J., Liu, B., 2011. Errors in particle tracking velocimetry with high-speed
cameras. Rev. Sci. Instrum. 82, 053707. https://doi.org/10.1063/1.3589267.
parametric analysis of operating conditions, as well as a coupling to Flatern, R.V., 2015. The Defining Series - Electrical Submersible Pumps. Oilfield Review.
one-dimensional multiphase flow models. Grassi, B., Strazza, D., Poesio, P., 2008. Experimental validation of theoretical models in
two-phase high-viscosity ratio liquid-liquid flows in horizontal and slightly inclined
pipes. Int. J. Multiphas. Flow 34 (10), 950–965. https://doi.org/10.1016/j.
Acknowledgements ijmultiphaseflow.2008.03.006.
Hanafizadeh, P., Hojati, A., Karimi, A., 2015. Experimental investigation of oil-water two
The authors would like to thank Equinor Brazil, ANP phase flow regime in an inclined pipe. J. Petrol. Sci. Eng. 136, 12–22. https://doi.
org/10.1016/j.petrol.2015.10.031.
(“Compromisso de Investimentos com Pesquisa e Desenvolvimento”), Hibbeler, R.C., 2010. Engineering Mechanics: Dynamics, twelfth ed. Prentice Hall.
and PRH/ANP for providing financial support for this work. The ac- Khalil, M., Kassab, S., Ismail, A., Elazab, I., 2008. Centrifugal pump performance under
knowledgements are also extended to Center for Petroleum Studies stable and unstable oil-water emulsions flow. In: 12th International Water
Technology Conference, Alexandria, Egypt.
(CEPETRO), University of Campinas, Artificial Lift & Flow Assurance
Li, A., Ahmadi, G., 1992. Dispersion and deposition of spherical particles from point
Research Group (ALFA), ISDB FlowTech and SINTEF-NO. sources in a turbulent channel flow. Aerosol Sci. Technol. 16, 209–226.
Loh, W.L., Premanadhan, V.K., 2016. Experimental investigation of viscous oil-water
Appendix A. Supplementary data flows in pipeline. J. Petrol. Sci. Eng. 147, 87–97. https://doi.org/10.1016/j.petrol.
2016.05.010.
Lum, J.Y.L., Al-Wahaibi, T., Angeli, P., 2006. Upward and downward inclination oil-
Supplementary data to this article can be found online at https:// water flows. Int. J. Multiphas. Flow 32 (4), 413–435. https://doi.org/10.1016/j.
doi.org/10.1016/j.petrol.2019.01.025. ijmultiphaseflow.2006.01.001.
Minemura, K., Murakami, M., 1980. A theoretical study on air bubble motion in a cen-
trifugal pump impeller. J. Fluid Eng. 102, 446–455.
Funding sources Moffat, R.J., 1988. Describing the uncertainties in experimental results. Exp. Therm. Fluid
Sci. 1, 3–17.
Mohammadi, M., Sharp, K.V., 2013. Experimental techniques for bubble dynamics ana-
Equinor Brazil and Brazilian National Agency of Petroleum, Natural lysis in microchannels: a review. J. Fluid Eng. 135. https://doi.org/10.1115/1.
Gas and Biofuels (ANP) provided financial support for this work. 4023450.
Monte Verde, W., Biazussi, J.L., Sassim, N.A., Bannwart, A.C., 2017. Experimental study
of gas-liquid two-phase flow patterns within centrifugal pumps impellers. Exp.
References Therm. Fluid Sci. 85, 37–51. https://doi.org/10.1016/j.expthermflusci.2017.02.019.
Morales, R., Pereyra, E., Wang, S., Shoham, O., 2013. Droplet formation through cen-
Abivin, P., Henaut, I., Chaudemanche, C., Argillier, J.F., Chinesta, F., 2009. Dispersed trifugal pumps for oil-in-water dispersions. SPE J. 18, 172–178. https://doi.org/10.
systems in heavy crude oils. Oil Gas Sci. Technol. 64 (5), 557–570. https://doi.org/ 2118/163055-PA.
10.2516/ogst/2008045. Morsi, S.A., Alexander, A.J., 1972. An investigation of particle trajectories in two-phase
Alkaya, B., Jayawardena, S.S., Brill, J.P., 2000. Oil-water flow patterns in slightly inclined flow systems. J. Fluid Mech. 55 (2), 193–208.
pipes. Energy New Millenium 14 (17), 1–8. Mundo, D.D.C., 2014. Study of Oil/water Flow and Emulsion Formation in Electrical
Angeli, P., Hewitt, G.F., 2000. Flow structure in horizontal oil-water flow. Int. J. Submersible Pumps. Master’s Thesis. The University of Tulsa, USA.
Multiphas. Flow 26 (7), 1117–1140. https://doi.org/10.1016/S0301-9322(99) Ofuchi, E.M., Ancajima, F.C., Stel, H., Neto, M.A.N., Morales, R.E.M., 2017. Influence of
00081-6. operational conditions on the trajectory of an air bubble inside a centrifugal pump.
ANSYS Inc, 2015. ANSYS Fluent Theory Guide. Release v16.2. In: Proceedings of JEM, the IV Journeys in Multiphase Flow. Sao Paulo, Brazil.
Barrios, L., Prado, M.G., 2011. Experimental visualization of two-phase flow inside an Perissinotto, R.M., Monte Verde, W., Biazussi, J.L., Castro, M.S., Bannwart, A.C., 2018.
electrical submersible pump stage. J. Energy Resour. Technol. 133 (4). https://doi. Experimental analysis on the velocity of oil drops in oil-water two-phase flows in ESP
org/10.1115/1.4004966. impellers. J. Offshore Mech. Arctic Eng. https://doi.org/10.1115/1.4042000.
Barrios, L., Prado, M.G., Kenyery, F., 2009. CFD Modeling inside an Electrical Accepted Manuscript.
Submersible Pump in Two-phase Flow Condition. ASME 2009 Fluids Engineering Pineda, H., Biazussi, J.L., López, F., Oliveira, B., Carvalho, R., Bannwart, A.C., Ratkovich,
Division Summer Meeting. N., 2016. “Phase distribution analysis in an Electrical Submersible Pump (ESP) inlet
Biazussi, J.L., 2014. A Drift-Flux Model for Gas-Liquid Flow in Electrical Submersible handling water–air two-phase flow using Computational Fluid Dynamics (CFD). J.
Pump Operating with Low Viscous Liquid. PhD dissertation. University of Campinas. Petrol. Sci. Eng. 139, 49–61. https://doi.org/10.1016/j.petrol.2015.12.013.
Brauner, N., Maron, D.M., 1992. Flow pattern transitions in two-phase liquid-liquid flow Rodriguez, O.M.H., Baldani, L.S., 2012. Prediction of pressure gradient and holdup in
in horizontal tubes. Int. J. Multiphas. Flow 18 (1), 123–140. https://doi.org/10. wavy stratified liquid-liquid inclined flow. J. Petrol. Sci. Eng. 96, 140–151. https://
1016/0301-9322(92)90010-E. doi.org/10.1016/j.petrol.2012.09.007.
Bulgarelli, N.A.V., Biazussi, J.L., Castro, M.S., Monte Verde, W., Bannwart, A.C., 2017a. Rodriguez, O.M.H., Oliemans, R.V.A., 2006. Experimental study on oil-water flow in
Experimental study of phase inversion phenomena in electrical submersible pumps horizontal and slightly inclined pipes. Int. J. Multiphas. Flow 32 (3), 323–343.
under oil-water flow. In: Proceedings of OMAE, the 36th International Conference on https://doi.org/10.1016/j.ijmultiphaseflow.2005.11.001.
Ocean, Offshore & Arctic Engineering. Trondheim, Norway. https://doi.org/10. Sabino, R.H.G., 2015. Analysis of Dynamic of a Gas Bubble in a Centrifugal Pump.
1115/OMAE2017-61865. Master’s Thesis. Federal University of Technology, Paraná, Curitiba, Brazil.
Bulgarelli, N.A.V., Biazussi, J.L., Perles, C.E., Monte Verde, W., Castro, M.S., Bannwart, Saffman, P.G., 1965. The lift on a small sphere in a slow shear flow. J. Fluid Mech. 22,
A.C., 2017b. Analysis of chord length distribution in phase inversion of water-oil 385–400.
emulsions at electrical submersible pump outlet. In: Proceedings of ExHFT, the 9th Sjoblom, J., 2005. Emulsions and Emulsion Stability: Surfactant Science Series, second ed.
World Conference on Experimental Heat Transfer, Fluid Mechanics and CRC Press.
Thermodynamics. Foz do Iguaçu, Brazil. Stel, H., Sirino, T., Ponce, F.J., Chiva, S., Morales, R.E.M., 2015. Numerical investigation
Caridad, J., Asuaje, M., Kenyery, F., Tremante, A., Aguillón, O., 2008. Characterization of of the flow in a multistage electric submersible pump. J. Petrol. Sci. Eng. 136, 41–54.
a centrifugal pump impeller under two-phase flow conditions. J. Petrol. Sci. Eng. 63, https://doi.org/10.1016/j.petrol. 2015.10.038.
18–22. https://doi.org/10.1016/j. petrol.2008.06.005. Takacs, G., 2017. Electrical Submersible Pumps Manual, second ed. Gulf Professional
Caridad, J., Kenyery, F., 2004. CFD analysis of electric submersible pumps (ESP) handling Publishing.
two-phase mixtures. J. Energy Resour. Technol. 126, 99–104. https://doi.org/10. Trevisan, F.E., Prado, M.G., 2011. Experimental investigation of the viscous effect on two-
1115/1.1725156. phase flow patterns and hydraulic performance of electrical submersible pumps. J.
Castro, M.S., Pereira, C.C., Santos, J.N., Rodriguez, O.M.H., 2012. Geometrical and ki- Can. Petrol. Technol. 50, 45–52. https://doi.org/10.2118/134089-PA.
nematic properties of interfacial waves in stratified oil-water flow in inclined pipe. Vielma, M., Atmaca, S., Sarica, C., Zhang, H., 2007. Characterization of oil/water flows in
Exp. Therm. Fluid Sci. 37, 171–178. https://doi.org/10.1016/j.expthermflusci.2011. horizontal pipes. In: SPE Annual Technical Conference and Exhibition, Anaheim,
11.003. California, USA.
Cavicchio, C.A.M., Biazussi, J.L., Castro, M.S., Bannwart, A.C., Rodriguez, O.M.H., Wang, W., Gong, J., Angeli, P., 2011. Investigation on heavy crude water two phase flow
Carvalho, C.H.M., 2018. Experimental study of viscosity effects on heavy crude oil- and related flow characteristics. Int. J. Multiphas. Flow 37, 1156–1164. https://doi.
water core-annular flow pattern. Exp. Therm. Fluid Sci. 92, 270–285. https://doi. org/10.1016/j.ij multiphaseflow.2011.05.011.
org/10.1016/j.expthermflusci.2017.11.027. White, F.M., 2011. Fluid Mechanics, seventh ed. McGraw-Hill.
Crowe, C., Schwarzkopf, J., Sommerfeld, M., Tsuji, Y., 2012. Multiphase Flows with Zhang, J., Cai, S., Li, Y., Zhu, H., Zhang, Y., 2016. Visualization study of gas-liquid two-
Droplets and Particles. CRC Press. phase flow patterns inside a three-stage rotodynamic multiphase pump. Exp. Therm.
Cubas Cubas, J.M., 2017. Experimental Study of Biphasic Flow Air-Water in a Radial Fluid Sci. 70, 125–138. https://doi.org/10.1016/j.expthermflusci.2015.08. 013.
Centrifugal Pump. Master’s Thesis. Federal University of Technology, Paraná, Zhu, J., Banjar, H., Xia, Z., Zhang, H., 2016. CFD simulation and experimental study of oil
Curitiba, Brazil. viscosity effect on multi-stage electrical submersible pump (ESP) performance. J.
Estevam, V., 2002. A Phenomenological Analysis about Centrifugal Pump in Two-phase Petrol. Sci. Eng. 146, 735–745. https://doi.org/10.1016/j.petrol.2016.07.033.

895

You might also like