You are on page 1of 11

PO198900 DOI: 10.

2118/198900-PA Date: 25-October-19 Stage: Page: 1 Total Pages: 11

Study of Oil/Water Flow and Emulsion


Formation in Electrical Submersible Pumps
D. Croce, Colorado School of Mines, and E. Pereyra, University of Tulsa

Summary
Oil/water dispersions from two oils of different viscosities (16 and 220 cp), equal surface tensions (27 mN/m), and similar densities
(810 and 873 kg/m3) were circulated in a viscous loop using an electrical submersible pump (ESP). The viscosity of the formed disper-
sions was measured at the discharge of the pump using an in-line pipe viscometer. The objective was to evaluate the impact of the effec-
tive viscosity of the emulsions on the head and flow rate delivered by the ESP.

Introduction
In general, oil and water flow together from the reservoir toward the production system at any point in the productive life of the well.
Given the immiscibility between these two fluids, the shearing and turbulence involved in the process results in the dispersion (or emul-
sion) of the two phases. Regularly, the fluid phase with the lower volumetric fraction becomes the dispersed phase, whereas the one in
the larger volume proportion becomes the continuous phase. The dispersed phase distributes itself in the continuous phase in the form
of droplets. The size and concentration of these droplets are sensitive to the physical properties of the fluids, their volumetric fractions,
and the magnitude of the shearing forces and turbulence acting on them.
According to which fluid becomes the dispersed phase and which fluid becomes the continuous phase, emulsions and dispersions are
classified into three groups: water in oil (W/O), oil in water (O/W), and complex (or multiple) mixtures. W/O and O/W are dispersions
of water droplets in oil and oil droplets in water, respectively, whereas the complex-mixture (or multiple-mixture) group comprises
those phase distributions in which droplets of one fluid are dispersed inside droplets of the other, which are simultaneously dispersed in
a continuous phase.
Given the imbalance of forces between the fluids, dispersions are usually not stable systems. When held motionless, the phases tend
to separate because of the repulsive forces between them. Different mechanisms take place in the separation of the emulsions, such as
the coalescence of the dispersed droplets, Ostwald ripening, flocculation, and sedimentation/creaming. In all these processes, as the
films surrounding the droplets thin and collapse because of gravity, the walls separating the droplets disappear. This allows for the
rejoining of the parts of the dispersed fluid, which form larger conglomerates and are more prone to the effects of buoyancy. This results
in the final separation of the phases, evidenced by a layered distribution in which the less-dense fluid occupies the upper layer and the
heavier fluid occupies the lower layer. To maintain the balance in a dispersion, it can be with supplied energy, the surface tension
between the phases can be reduced, or the volumetric fraction of the fluids can be altered. For the first, the action of shearing forces
causes the dispersed phase to break into even smaller particles. This eases its distribution among the continuous phase. For the second
option, the presence of elements that reduce the repulsive forces between the fluids, such as salts, maintains the dispersed droplets in
place, reducing their tendency to coalesce and float. The substances capable of producing this result are called surfactants. Their action
is explained by the electromagnetic forces that their molecular distribution generates on the surrounding fluids. In the third case, as the
volumetric fractions are changed, the forces of the dispersed phase over the continuous phase are altered in such a way that the droplets
are less affected by the repulsion forces. This alteration can result in a phase inversion, where the dispersed droplets reorganize and
become the continuous phase, while the pre-existing continuous phase breaks, generating the new dispersed phase. This phase inversion
has been found to take place also by changes in temperature, given the modification of the fluid properties.
When phase inversion occurs, the equipment through which emulsions flow is affected by the change in frictional losses, caused by
a dramatic change in the effective viscosity of the dispersion. Because there is not an exact amount of energy or volumetric fraction
required but a range within which this phenomenon occurs, it is important to determine the possible range of operation in which the
phase inversion might occur and how the fluid properties affect it.
ESPs, which are one of the most widely used devices for artificial lift in the oil industry, are typically affected by the phase-
inversion phenomenon. The energy supplied by these devices across their stages as they rotate causes the shear and mixing necessary to
emulsify the oil and water coming from the formation and maintain the emulsion with a certain stability level. Understanding the effect
of the viscosity changes on the performance of facility equipment is key in the production-system design process. For this, here we pres-
ent an experimental analysis of the impact of the effective viscosity of different emulsions generated by an ESP. The effect is presented
in terms of the concentration of the phases and its effect on the head and flow rate delivered by the pump.

Experimental Setup
To measure the impact of the change of the effective emulsion viscosity, a series of experiments were held using different mixtures of
oil and water while operating the ESP.
The experiment was performed at the University of Tulsa Artificial Lift Project viscous loop. The facility consists of a 2-in. inner-
diameter steel pipe, a flowmeter, a heat-exchange system, an ESP, and a variable-speed driver as shown in Fig. 1.

ESP. The studied ESP consisted of seven stages of 4-in.-diameter impellers and diffusers with six and eight blades, respectively. The
best efficiency point of the pump was 1,750 B/D at 3,500 rev/min. A variable-speed drive was used to control the rotational speed of
the device. A cyclonic valve (installed at the pump outlet) was used to control the flow rate. Although the results in this work are pre-
sented only for this speed, tests were also performed at other rotating velocities for the net-positive-suction-head validation, as well as

Copyright V
C 2019 Society of Petroleum Engineers

Original SPE manuscript received for review 1 February 2019. Revised manuscript received for review 14 May 2019. Paper (SPE 198900) peer approved 8 July 2019.

2019 SPE Production & Operations 1

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 2 Total Pages: 11

for the data obtained from the in-line pipe viscometer. Pressure transducers were used to measure the pressure increase through the
third stage (from the entrance of the third impeller to the outlet of the third diffuser) and the entire ESP (from the inlet to the outlet of
the pump). Temperature detectors were also installed in the same port. As with the pipe and the rest of the operating components of the
loop, the ESP was insulated to minimize heat exchange with the surroundings, allowing a more accurate temperature control of the
circulating fluid.

TT TT

PT PT
TT TT In-line pipe viscometer

Variable-speed
PT PT

driver
PT PT
Pump

Feeding port TT
Flowmeter

Sampling valve
Heat-exchange system

PT Pressure transducer

TT Temperature transducer

Fig. 1—Configuration of the experimental facility.

In-Line Pipe Viscometer. The effective viscosity of the dispersions generated from the operation of the pump was measured with a
pipe viscometer at the discharge of the pump. It was composed of two pressure transducers along a horizontal, uninterrupted straight
section of a pipe. The location of the pressure transducers with respect to the pump outlet and the elbows of the loop was established
according to the required length for the full development of the flow pattern, according to the inner diameter of the pipe (d). Brito
(2012) successfully used the White (1998) recommended minimum length (L), which is dependent on the Reynolds (Re) number of the
circulating fluids, as indicated in Eqs. 1 and 2 for laminar and turbulent flow, respectively.
L
¼ 0:06  Re; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
d
and
L
¼ 4:4  Re1=6 : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ
d

Fluids. Two different commercial oils were used for the generation of the O/W dispersions, both of similar density and surface tension
but different in orders of magnitude in terms of viscosity. The first, ISOPARÔ-V, presented a viscosity of nearly 16 cp, and the second
oil, a general nondetergent oil (DN-20), had a viscosity of approximately 220 cp for the same temperature, as shown in Table 1. The
density and viscosity of the oils as a function of temperature were available from the manufacturer’s data (shown in Figs. 2 and 3,
respectively). These data were used later for measuring the head delivered by the pump during the single-phase and multiphase tests.

Properties at 60°F ISOPAR-V DN-20


3
Density (kg/m ) 810 873
°API value 40.3 30.5
Viscosity (cp) 15.7 222
Surface tension (mN/m) 27 27.5

Table 1—Main properties of the oils used for the experiments.

Heat Exchanger. Because of the sensitivity of the oil viscosity with respect to temperature, a heat exchanger was used to remove the
heat generated by the friction across the pump and along the loop, maintaining the temperature constant at approximately 95 F (63 F)
during the experiments.

Experimental Procedure
This section describes the experimental procedure considered during the data acquisition.

Single-Phase Tests. To validate the operation of the pipe viscometer, as well as the net positive suction head available for the pump,
the first step was to operate the loop with only one fluid (single-phase flow).

2 2019 SPE Production & Operations

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 3 Total Pages: 11

900
ISOPAR-V
DN-20
880

Density (kg/cm3)
860

840

820

800

780
0 20 40 60 80 100 120 140 160
Temperature (°F)

Fig. 2—Density of ISOPAR-V and DN-20 vs. temperature.

600
ISOPAR-V
DN-20
500

400
Viscosity (cp)

300

200

100

0
0 20 40 60 80 100 120 140 160
Temperature (°F)

Fig. 3—Viscosity of ISOPAR-V and DN-20 vs. temperature.

To obtain the effective viscosity of the oil circulating inside the pipe, the definition of the Reynolds number was used as indicated in
Eq. 3, where q is the density and u is the velocity of the fluid
qud
leff ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
Re
The density of the fluid was measured using a commercial Coriolis mass flowmeter installed in the loop and validated against the
data of the oil supplier for each operating temperature. The velocity was obtained from the flow rate measured from the same flowmeter
and the known internal diameter of the pipe.
Considering different operational flow rates from the pump, the Reynolds number was obtained for laminar flow as a function of the
Fanning friction factor ( fF), as indicated in Eq. 4
16
Re ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ
fF
or for turbulent flow from the Colebrook equation (Colebrook 1939) as indicated in Eq. 5
 
1 e=d 2:51
pffiffiffiffi ¼ 2  log þ pffiffiffiffi ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð5Þ
fF 3:7 Re  fF
where e is the roughness of the pipe.
From a momentum balance over a control volume of fluid inside the horizontal pipe, the friction factor can be obtained as
d dP
fF ¼  ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð6Þ
2  ðq  u2 Þ dL
where dP/dL is the pressure drop per unit length in the pipe section.

Two-Phase Tests. To study the effect of the water fraction over the head delivered by the pump, the fraction of water was varied in
intervals of 5%, starting from 0% and increasing to as close to 100% as possible (the total removal of the oil could not be expected
because of the wettability of the inner walls of the pipe). For every change in water fraction, the fluids were circulated and allowed to
mix before measuring the head delivered by the pump. We discuss the steps in more detail in the following subsections.

2019 SPE Production & Operations 3

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 4 Total Pages: 11

Step 1: Water-Fraction Change. By bleeding the sampling valve in the loop, the equivalent volume for a 5% variation of circulat-
ing mixture was extracted. This volume was replaced with the studied oil through the feeding port.
Step 2: Water-Fraction Validation. Knowing the densities of both the oil (qoil) and water (qw) as a function of temperature (as
shown in Fig. 2) and having measured the temperature of the mixture through the loop, the intended water fraction ( fw) was validated
with the mixture density (qM) obtained from the Coriolis flowmeter reading as

qM jTemp ¼ qoil jTemp  ð1  fw Þ þ qw jTemp  ð f w Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð7Þ

If the mixture-density reading from the flowmeter did not match the expected value for the volume of oil introduced, then Steps 1
and 2 were repeated until a match was obtained.
Step 3: Liquid Sampling. After an arbitrary amount of time, a sample was withdrawn from the loop. The objectives were to validate
the water fraction computed through the Coriolis flowmeter (see Eq. 7) and to record the separation time of the emulsion.
Step 4: Head Reading. The effect of the water-fraction change on the performance of the pump was measured through the head
delivered by the stage (DHstage). For this, the difference between the pressure readings at the entrance and the outlet of the stage was
used as
DP
DHstage ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8Þ
qM  g
where g is the gravitational constant.
For each water fraction, the cyclonic valve at the discharge of the pump was used to choke the mixture flow to build the head curve.

Step 5: Compute the Viscosity of the Emulsion. The effective viscosity of the emulsion was determined using the same procedure
described for the single-phase tests, as shown in Eq. 7. A corrected Reynolds number (Re0 ) was used to calculate the effective viscosity
(leff) resulting from the presence of the droplets created by the shear of the blades on the circulating fluid, as shown in Eq. 9, where uM
is the velocity of the mixture.
qM  uM  d
leff ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð9Þ
Re0
This adjusted Reynolds number (Re0 ) was obtained from the solution to the turbulent-flow regime, including a modification factor B0 ,
proposed by Vielma (2007), related to the droplet diameter of the dispersed phase,
pffiffiffi ( " pffiffiffiffi # ) !
1 2 1 Re0  fF0
pffiffiffiffi0 ¼  0  log   pffiffiffiffi0  2:54 þ B0 : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð10Þ
fF 2 k 1 þ 0:2  e=D  Re0  fF

The corrected Fanning friction factor fF0 was calculated from the pressure decline across the pipe viscometer,

d dP
fF0 ¼  : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð11Þ
2  ðqM  u2M Þ dL

The constant B0 is related to the droplet minimum diameter (dmin) and the volume fraction of the dispersed phase (u) after experi-
mental results (Laflin and Oglesby 1976; Trallero 1995; Angeli and Hewitt 1999; Soleimani 1999),
h b4
i
B0 ¼ B  bu1  ð1  uÞb2 þ b3  d~min ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð12Þ

where B ¼ 5.0, b1 ¼ 6, b2 ¼ 1.5, b3 ¼ –12, b4 ¼ 1.1, and d~min is obtained as

8r
d~min ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð13Þ
q  u2  d

k0 is a modified von Karman constant, obtained as


h c4
i
k0 ¼ k  cu1  ð1  uÞc2 þ c3  d~max ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð14Þ

where k ¼ 0.41, c1 ¼ 0.0041, c2 ¼ –5.1, c3 ¼ 4.0, and c4 ¼ 1.8 for W/O dispersions, whereas for O/W dispersions, c1 ¼ 1, c2 ¼ 0.12,
c3 ¼ 4.0, and c4 ¼ 1.8.
As for d~max , it can be obtained as a function of the Weber number (We) using the Blasius equation with a homogeneous no-slip
Reynolds number (Torres-Monzon 2006),
 
2:221 u 0:6
d~max ¼  : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð15Þ
We0:6  fF0:4 1u

Results
From the single-phase tests, the in-line pipe viscometer was validated against a commercial rheometer, as shown in Fig. 4. The viscosity
calculated from the pressure decline in the pipe viscometer presented the same trend and stayed within a relative discrepancy of less
than 8%, suggesting a fair reliability of the in-line pipe viscometer.
For the two-phase tests, from a qualitative perspective, extremely stable dispersions were generated for each of the oils. The samples
collected during the tests presented separation times on the order of days, as can be seen in Figs. 5 and 6.

4 2019 SPE Production & Operations

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 5 Total Pages: 11

10.0

9.5

9.0

Viscosity (cp)
8.5

8.0

7.5
3,500 rev/min, pipe
7.0 3,000 rev/min, pipe
2,500 rev/min, pipe
6.5
Rheometer
6.0
87 89 91 93 95 97
Temperature (°F)

Fig. 4—ISOPAR-V viscosity vs. temperature (pipe viscometer and rheometer).

(a) t = 0 (b) 1 minute (c) 4 hours (d) 24 days

Fig. 5—Time lapse of a 30:70 W/O of a DN-20 and water emulsion.

(a) t = 0 (b) 1 minute (c) 6 days (d) 1 month

Fig. 6—Time lapse of a 35:65 W/O sample of a DN-20 and water emulsion.

The required time for all the samples to achieve full separation for both oils was larger than the typical residence time of most of the
production equipment (3 minutes); therefore, the oil/water mixtures generated can be classified as emulsions. For the two oils, even
when the emulsion showed a layering time in the range of minutes, it took hours and even days (in the case of the sample with the
water cut closer to the phase-inversion point) to separate. Still, after the separation, a small and dense layer of droplets was clearly visi-
ble at the interface.
Fig. 7 shows the viscosity of the dispersions obtained from each of the two-phase tests (water and ISOPAR-V and water and
DN-20) vs. the water fraction. It can be seen how the effective viscosity (shown as a ratio of the viscosity of the dispersion to that of the
pure oil) presented a sustained increase as the water fraction was increased, showing a sudden interruption of the trend-phase-inversion
point for both cases. This is a typical behavior of dispersed-oil/water systems, as presented in the works of Arirachakaran et al. (1989)
and Vielma (2006). Before and after the inversion point, the behavior of the circulating emulsion will more closely resemble that of the
continuous phase than that of the dispersed phase. We observed an inversion point of 40% of water cut for the water/ISOPAR-V disper-
sion and 35% for the water/DN-20 dispersion. On the right of the inversion point (water cut > 40% and water cut > 35%), the water
was the continuous phase, and on the left (water cut < 40% and water cut < 35%), the oil was the continuous phase.
Considering the mixtures as emulsions, it is important to observe the performance of the pump with respect to the change in the
water fraction. Figs. 8 and 9 show the differential pressure across a stage vs. the mixture flow rate for different water fractions. For
the water-continuous region, the largest DP corresponds to the 100% water cut. As the water fraction decreased, the DP delivered by the
pump decreased, whereas the effective viscosity of the mixture increased. The reduction of the differential pressure with respect to the

2019 SPE Production & Operations 5

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 6 Total Pages: 11

water cut is larger in the water-continuous region compared with the oil-continuous side. For the oil-continuous region, the maximum
DP corresponds to 0% water cut, corroborating the effect of the effective mixture viscosity over the pump curve. In the oil-continuous
region, the variations of the pump-pressure difference are considerably smaller compared with those in the water-continuous region. This
phenomenon was observed for all flow rates.

7 ISOPAR-V DN-20
40%, 6.5
6

μeff /μoil
4
35%, 2.4
3

0
0 10 20 30 40 50 60 70 80
fw (%)

Fig. 7—Relative effective viscosity in the pipe viscometer (ISOPAR-V and DN-20 O/W dispersions).

14
fw = 0%
12 fw = 5%
fw = 10%
fw = 15%
10
fw = 20%
fw = 25%
ΔP (psi)

8
fw = 30%
fw = 35%
6
fw = 40%
fw = 45%
4 fw = 55%
fw = 75%
2 fw = 100%

0
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200
Flow Rate (B/D)

Fig. 8—Pressure increase delivered by the third stage vs. flow rate for different water fractions (ISOPAR-V and water at 3,500 rev/min).

14
fw = 5%
12 fw = 10%
fw = 15%
10 fw = 20%
fw = 25%
ΔP (psi)

8 fw = 35%
fw = 40%
6 fw = 45%
Water
4

0
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200
Flow Rate (B/D)

Fig. 9—Pressure increase delivered by the third stage vs. flow rate for different water fractions (DN-20 and water at 3,500 rev/min).

As shown in Figs. 10 and 11, the effect of the viscosity increase caused by water fraction change was registered in the third stage
and across the entire pump. When analyzing Fig. 10 and Fig. 12, it can be seen that for the emulsions generated with the two oils used,
the increase in viscosity resulting from water fraction change is the same. From this, it is clear that the head requires correction not only
for density but also for viscosity. The head curves, when the water fractions were less than the inversion point (on an oil-continuous

6 2019 SPE Production & Operations

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 7 Total Pages: 11

emulsion), stayed within the same values of the oil curve (100% oil), showing a slightly increasing trend as the water fraction
approached the inversion point. Once past the inversion point (40 and 35%, respectively, for ISOPAR-V and DN-20), the head curves
tended to the values computed when operating with 100% water, which were the highest head curves computed for the stage. Noting
that the density of the mixtures basically did not change as the water fraction was taken from 0 to 100% water, for the W/O emulsions
handled, it can be seen then that there is a clear relationship between the performance of the pump and the change in viscosity, with the
latter being a function of the water fraction.

9
fw = 0%

840 B/D
fw = 5%
8
fw = 10%
fw = 15%
7 fw = 20%
ΔH (m) fw = 25%
6 fw = 30%
fw = 35%
fw = 40%
5
fw = 45%
fw = 55%
4 fw = 75%
Water

3
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200
Flow Rate (B/D)

Fig. 10—Head delivered by the third stage vs. flow rate for different water fractions (ISOPAR-V and water at 3,500 rev/min).

60
fw = 0%
55 fw = 5%
fw = 10%
50 fw = 15%
fw = 20%
45
ΔH Pump (psi)

fw = 25%
fw = 30%
40
fw = 35%
fw = 40%
35
fw = 45%
fw = 55%
30
fw = 75%
25 fw = 100%

20
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200
Flow Rate (B/D)

Fig. 11—Head delivered by the pump vs. flow rate for different water fractions (ISOPAR-V and water at 3,500 rev/min).

9
fw = 5%
fw = 10%
840 B/P

8
fw = 15%
fw = 20%
7 fw = 25%
fw = 35%
ΔH (m)

6 fw = 40%
fw = 45%
Water
5

3
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200
Flow Rate (B/D)

Fig. 12—Head delivered by the third stage vs. flow rate for different water fractions (DN-20 and water at 3,500 rev/min).

2019 SPE Production & Operations 7

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 8 Total Pages: 11

This behavior is easier to observe when analyzing the head values for a fixed flow rate as a function of the water fraction. As pre-
sented in Fig. 13, for 840 B/D for both emulsions, clearly when at less than the inversion point, as the water fraction increased in the
oil-continuous region, the head dropped until the phase-inversion point was reached. When operating at higher than the phase-inversion
point, as the water fraction increased, the head stayed at those values registered when handling 100% water. This behavior is clearly
related to the change in viscosity of the emulsion presented in Fig. 7. As shown in Fig. 10, the same behavior for the head and flow rate
delivered was observed for the entire pump. This means that most probably the emulsion properties were not affected by the stages
after the third stage.

7.0

6.5

6.0

H (m)
5.5

5.0

Experimental
4.5 Turzo model

4.0
0 20 40 60 80 100
fw (%)

Fig. 13—Head delivered at 840 B/D by the third stage for different water fractions (ISOPAR-V and water at 3,500 rev/min). Turzo
model from Takacs and Turzo (2006).

The previously discussed correction of the head for the change in viscosity was approached using the model of Takacs and Turzo
(2006), which is dependent on the dynamic viscosity of the mixture, but the model fails to replicate the behavior when below the
inversion point.
Using dimensionless analysis to reflect the effect of the water-fraction change provides a better look at how the variation in the
effective viscosity of the emulsion affects the head delivered by the pump. We used the definition of the head coefficient (CH),
gH
CH ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð16Þ
N 2 D2
the flow coefficient (CQ),
q
CQ ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .ð17Þ
ND3
and the specific speed (NS),
pffiffiffiffi
N Q
NS ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð18Þ
ðgHÞ3=4
Figs. 14 and 15 show that as the water fraction is increased from 0% toward the inversion point (in a W/O emulsion), the head coef-
ficient drops. However, after the phase-inversion point, as the water fraction is increased (in an O/W emulsion), the head coefficient
increases toward the values registered for water in single-phase tests.

0.045

0.040

0.035 fw = 0%, NS = 5
fw = 10%, NS = 5
CH

fw = 20%, NS = 5
0.030 fw = 35%, NS = 5
fw = 45%, NS = 5
fw = 55%, NS = 5
0.025
fw = 75%, NS = 5
fw = 100%, NS = 5
0.020
0 20 40 60 80 100
fw (%)

Fig. 14—Head coefficient (CH) vs. water fraction at a constant specific speed (NS) (DN-20 and water at 3,500 rev/min).

8 2019 SPE Production & Operations

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 9 Total Pages: 11

0.060

0.055

0.050
fw = 5%, NS = 2.5
0.045 fw = 10%, NS = 2.5

CH
fw = 15%, NS = 2.5
0.040 fw = 20%, NS = 2.5
fw = 25%, NS = 2.5
0.035 fw = 35%, NS = 2.5
fw = 40%, NS = 2.5
0.030 fw = 45%, NS = 2.5
Water, NS = 2.5
0.025
0 20 40 60 80 100
fw (%)

Fig. 15—Head coefficient vs. water fraction at a constant specific speed (DN-20 and water at 3,500 rev/min).

Similar to the decline of the head coefficient, the flow coefficient shows a decrease before the inversion point as the water fraction
grows, but increases after the phase-inversion point as the water cut grew, as shown in Figs. 16 and 17. Given that a variable-speed
drive was used to control the pump, the rotating velocity was maintained constant; in theory, this should have resulted in the same exit
velocity of the fluids as they left the end of the blades of the propeller. There should not have been significant differences in terms of
the flow coefficient between the two dispersions if only the density was considered, given the proximity of the two values for each (less
than 7% difference). Given that a major difference was still found between the two dispersions studied, the only variable to which this
difference can be attributed is the increase in viscosity. In the same way, then, it can be concluded that the increase in the viscosity
diminishes the flow capacity of the stage.

0.0065

0.0060

0.0055
fw = 0%, NS = 5
fw = 10%, NS = 5
0.0050 fw = 20%, NS = 5
CQ

fw = 35%, NS = 5
0.0045 fw = 45%, NS = 5
fw = 55%, NS = 5
0.0040 fw = 75%, NS = 5
fw = 100%, NS = 5
0.0035
0 20 40 60 80 100
fw (%)

Fig. 16—Flow coefficient vs. water fraction at a constant specific speed (ISOPAR-V and water at 3,500 rev/min).

0.0030

0.0025

0.0020 fw = 5%, NS = 2.5


fw = 10%, NS = 2.5
fw = 15%, NS = 2.5
0.0015
CQ

fw = 20%, NS = 2.5
fw = 25%, NS = 2.5
0.0010 fw = 35%, NS = 2.5
fw = 40%, NS = 2.5
0.0005 fw = 45%, NS = 2.5
Water, NS = 2.5

0
0 20 40 60 80 100
fw (%)

Fig. 17—Flow coefficient vs. water fraction at a constant specific speed (DN-20 and water at 3,500 rev/min).

2019 SPE Production & Operations 9

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 10 Total Pages: 11

Conclusions and Recommendations


A clear relation was observed between the performance of an ESP and the change in the viscosity of an O/W emulsion. For an oil-
continuous emulsion, the increase in the viscosity as the water fraction increased to less than the phase-inversion point reflected directly
in a decrement of the head delivered by the pump. Once higher than the phase-inversion point, the head values computed were closer to
those when handling 100% water. It is important to consider that the changes in viscosity, flow-rate capacity, and delivered head are
sensitive to the number of stages, geometry of the blades, and rotating velocity of the ESP, as well as to the chemical properties of the
oils used. This should be considered for further studies.
The model of Turzo and Takacs (2006) fell short when trying to replicate the behavior of a centrifugal pump handling emulsions at
less than the phase-inversion point. The models failed to adjust to the marked increase in the viscosity of the emulsion as the water fraction
increased before the phase-inversion point. We believe that because the Turzo and Takacs (2006) model uses kinematic viscosity of the
fluid, it cannot reflect the effect of the disproportionate change between the effective dynamic viscosity and the density of the emulsion.
Additional studies considering ESPs with different blade geometries are recommended to develop a general correction factor to the
head and flow rate delivered by the pump when handling oil/water dispersions. This factor should be expressed as a function of the
water fraction.

Nomenclature
CH ¼ head coefficient, dimensionless
CQ ¼ flow coefficient, dimensionless
d ¼ pipe inner diameter, m
dP/dL ¼ pressure decline per unit length, Pa/m
d~max ¼ maximum droplet diameter, dimensionless
d~min ¼ minimum droplet diameter, dimensionless
fF ¼ Fanning friction factor
fFm ¼ mixture Fanning friction factor
fw ¼ water fraction
Fr ¼ Froude number
g ¼ acceleration of gravity, m/s2
L ¼ length of the pipe, m
N ¼ impeller rotating speed, rev/min
NS ¼ specific speed
Re ¼ Reynolds number
uc ¼ velocity of the continuous phase, m/s
um ¼ mixture velocity, m/s
Veo ¼ volume of oil extracted
Vm ¼ volume of mixture
Vme ¼ volume of mixture extracted
We ¼ Weber number
e ¼ roughness of the pipe
l ¼ viscosity, cp [centipoise ¼ 103 Pas]
leff ¼ effective viscosity, cp [centipoise ¼ 103 Pas]
lM ¼ viscosity of the mixture
qc ¼ density of the continuous phase, kg/m3
qm ¼ mixture density, kg/m3
qoil ¼ oil density, kg/m3
qw ¼ water density, kg/m3
r ¼ surface tension, N/m
u ¼ dispersed-phase concentration

References
Angeli, P. and Hewitt, G. F. 1999. Pressure Gradient in Horizontal Liquid-Liquid Flows. Int Jour Multiphase Flow 24 (7): 1183–1203. https://
doi.org/10.1016/S0301-9322(98)00006-8.
Arirachakaran, S., Oglesby, K. D., Malinowsky, M. S. et al. 1989. An Analysis of Oil/Water Flow Phenomena in Horizontal Pipes. Paper presented at
the SPE Production Operations Symposium, Oklahoma City, Oklahoma, USA, 13–14 March. SPE-18836-MS. https://doi.org/10.2118/18836-MS.
Brito, R. 2012. Effect of Medium Oil Viscosity on Two-Phase Oil-Gas Flow Behavior in Horizontal Pipes. MS thesis. University of Tulsa, Tulsa,
Oklahoma, USA.
Colebrook, C. F. 1939. Turbulent Flow in Pipes with Particular Reference to the Transition Region Between the Smooth and Rough Pipe Laws. Jour Inst
Civil Eng (London) 11 (4): 133–156. https://doi.org/10.1680/ijoti.1939.13150.
ISOPAR-V is a trademark of ExxonMobil Corporation, 5959 Las Colinas Blvd., Irving, Texas, USA 75039.
Laflin, G. C. and Oglesby, K. D. 1976. An Experimental Study on the Effects of Flow Rate, Water Fraction and Gas Liquid Ratio on Air-Oil-Water Flow
in Horizontal Pipes. BSc thesis, University of Tulsa, Tulsa, Oklahoma, USA.
Soleimani, A. 1999. Phase Distribution and Associated Phenomena in Oil-Water Flows in Horizontal Tubes. PhD dissertation, Imperial College, Univer-
sity of London, London, UK.
Takacs, G. and Turzo, Z. 2000. Equations Correct Centrifugal Pump Curves for Viscosity. Oil & Gas J. 98 (22): 57–61.
Torres-Monzon, C. F. 2006. Modeling of Oil-Water Flow in Horizontal and Near Horizontal Pipes. PhD dissertation, University of Tulsa, Tulsa,
Oklahoma, USA.
Vielma, J. C. 2006. Rheological Behavior of Oil-Water Dispersion Flow in Horizontal Pipes. MS thesis, University of Tulsa, Tulsa, Oklahoma, USA
(August 2006).
Vielma, C. 2007. Characterization of Oil-Water Flows in Horizontal Pipes. MS thesis, University of Tulsa, Tulsa, Oklahoma, USA.
White, F. M. 1998. Fluid Mechanics. Boston, Massachusetts, USA: McGraw-Hill.

10 2019 SPE Production & Operations

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037


PO198900 DOI: 10.2118/198900-PA Date: 25-October-19 Stage: Page: 11 Total Pages: 11

Daniel Croce is a research and teaching assistant at the Petroleum Engineering Department of the Colorado School of Mines.
His research interests include multiphase flow, production optimization, data analysis, computational fluid dynamics, flow assur-
ance, and artificial lift. He holds a bachelor’s degree in mechanical engineering from Simon Bolivar University, Venezuela, and a
master of science degree in petroleum engineering from the University of Tulsa.
Eduardo Pereyra is an associate professor at the McDougall School of Petroleum Engineering and associate director of the Fluid
Flow Project and Horizontal Wells Artificial Lift Project at the University of Tulsa. His research interests include multiphase flow and
its applications to pipelines, surface oil and gas facilities, artificial lift, separation, and metering systems. Pereyra holds bachelor’s
degrees in mechanical engineering and systems engineering from the University of Los Andes, Venezuela, and master of sci-
ence and PhD degrees in petroleum engineering from the University of Tulsa.

2019 SPE Production & Operations 11

ID: jaganm Time: 19:18 I Path: S:/PO##/Vol00000/190037/Comp/APPFile/SA-PO##190037

You might also like