You are on page 1of 14

IPTC 14944

Experimental Study on the Phenomenon of Barite Sag


Ali Parvizinia, Ramadan Ahmed, Samuel Osisanya, University of Oklahoma

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


Copyright 2011, International Petroleum Technology Conference

This paper was prepared for presentation at the International Petroleum Technology Conference held in Bangkok, Thailand, 7–9 February 2012.

This paper was selected for presentation by an IPTC Program Committee following review of information contained in an abstract submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the International Petroleum Technology Conference and are subject to correction by the author(s). The material, as presented, does not necessarily
reflect any position of the International Petroleum Technology Conference, its officers, or members. Papers presented at IPTC are subject to publication review by Sponsor Society
Committees of IPTC. Electronic reproduction, distribution, or storage of any part of this paper for commercial purposes without the written consent of the International Petroleum Technology
Conference is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous
acknowledgment of where and by whom the paper was presented. Write Librarian, IPTC, P.O. Box 833836, Richardson, TX 75083-3836, U.S.A., fax +1-972-952-9435

Abstract
Increasing global energy demand and diminishing petroleum reserves have continually elevated the significance of extended
reach deepwater drilling. The exorbitant cost of such specialized technologies necessitates efficient operations and the
shortest time to target possible. Settling of barite particles (barite sag) while drilling often results in undesirable equivalent
circulating density (ECD) and pressure fluctuations in directional wells. The sedimented particles of barite form a bed on the
low side of a wellbore, causing density differences in the cross section, which generate pressure imbalance and downhole
ECD fluctuations. Barite sag phenomenon can lead to a variety of drilling problems such as lost circulation, stuck pipe, poor
cement job and wellbore instability. The phenomenon is likely to occur under dynamic conditions, elevated temperatures and
inadequate annular flow velocities. Barite sag is also exacerbated in the absence of drillstring rotation and low annular
diameter ratios.
This article presents an experimental study conducted on barite sag behaviors of oil-based fluids. A cylindrical sag-
testing cell has been developed to measure the level of barite sag at different shear rates (from 0 to 0.82 1/s) and temperatures
(80°F and 120°F). The cell has a rotating dick at the top to create the shear field in the cylinder, which is filled with test fluid.
Pressure sensors were mounted on the wall of the cylinder to measure sag tendencies of the sample as a function of pressure
gradient. Several tests were conducted on OBM fluid at different temperatures and disk rotation speeds. Density and
rheological properties of the samples were measured before and after the test.
The experimental results indicate significant barite sag, especially in fluids subjected to high shear rates and elevated
temperature. For the fluids tested, the change in temperature had the greater influence over sag behavior in comparison to
shear rate. Results suggest that the viscosity of the oil-phase, which is very sensitive to temperature, has more pronounced
effect on sag than the rheology of the mud system. The outcomes of this investigation are very useful for wellbore pressure
management and drilling optimization.

1. Introduction
During drilling of oil and gas wells, mud is required to perform many tasks including hole cleaning, lubrication and cooling
of the drill bit, stabilization of the wellbore and bottom hole pressure control. Stability of the mud is critical for successful
completion of a drilling operation. Sedimentation of barite particles causes density variations in the cross section of a
wellbore. This generates pressure imbalance that induces secondary flow (i.e. downward sliding of heavy barite bed and
upward flow of lighter fluid layer). The secondary flow results in a convection current that in turn accelerates the settling
process. Barite sag can lead to a variety of drilling problems such as lost circulation, stuck pipe, poor cement job, excessive
fluid losses and wellbore instability (McLean & Addis, 1996). Although a number of studies have been conducted to
understand the mechanisms responsible for initiation and exacerbation of barite sag, this phenomenon is still not fully
understood.
The principal objective of this study is to examine the effects of fluid rheology, shear rate and temperature on static
and dynamic barite sag. Sedimentation theory suggests that the fluid temperature enhances the rate of barite sag by reducing
the viscosity of the fluid. The shear rate is also expected to increase the rate of barite sedimentation due to drilling fluid’s
shear thinning behavior and subsequent viscosity reduction. The shearing of the fluid could also encourage the separation of
heavy solids particles from the mud structure, which may in turn increase the sagging rate. Moreover, heavy components of
drilling mud have the tendency to aggregate and settle out of the fluid.
Barite sag in oil-based mud is a complex process, such as attempting to describe particle sedimentation in two-phase
fluid systems. As a result, theoretical analysis of barite sag is very challenging. This study predominately used experimental
methods to examine barite sag under static and dynamic conditions. Barite sag measurements were performed using a
2 IPTC 14944

cylindrical sedimentation cell containing a rotating disk to shear the mud sample. The cylindrical sedimentation cell
measurements provided pressure profiles in the sample as a function of time and depth from the surface of the sample. The
pressure profile measurements were used to determine the density variation in the sample and the level of barite sag. The
rheologies of the samples were measured at the test temperature to quantify viscosity reduction due to the change in
temperature and/or shear rate.

2. Literature Review
The first experiment conducted by Boycott (1920) on corpuscles settlements showed that the sedimentation rate of the

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


particles is a function of tubing inclination. Some studies (Hanson et al., 1990; Saasen et al., 1995; Bern, 1996; Bern, 1998;
Omland, 2004) suggested that a similar phenomenon could be involved in barite sag. Barite sag is often observed following
circulating bottoms-up operations, such as tripping pipe, in which the mud has not been circulated for an extended period of
time. It is more problematic in directional wells because fluid flow is skewed by the effects of drill pipe eccentricity, resulting
in low local shear rates beneath the eccentric pipe, creating conditions conducive to barite sag. In the past, barite sag was
associated with a static field environment. Consequently, test devices and rheological measurements were originally based on
static conditions. However, barite sedimentation and bed formation can take place during mud circulation, and bed sliding
occurs when circulation is stopped (Hanson et al., 1990). The bed begins to slide down when the critical thickness has been
reached causing undesirable bottom pressure fluctuations.
A recent study (Saasen, 2002) on barite sag found that both shear thinning and shear thickening behaviors exist in
oil-based fluids. At low shear rates, it has been observed that when the shear rate is increased, the viscosity of the fluid
increases and the fluid behaves like a shear thickening material. This behavior continues until a certain critical shear rate
value is reached, at which more increase in shear rate results in significant reduction in viscosity. As the shear rate increases,
there is a tendency of water droplets to group themselves in strings and clusters. This grouping results in the development of
large volumes of free oil between strings and clusters, resulting in a viscosity reduction and shear-thinning effect at high
shear rates (Saasen, 2002). In the low shear rate range, low-viscosity drilling mud can allow sagging if it’s composition does
not include sufficient amounts of organophilic clay, which is mostly used as primary viscosifier for oil-based mud (OBM)
and synthetic-based fluid (SBF).
The viscoelastic properties of drilling fluids are related to gel formation, which must be enhanced to keep barite in
suspension. Saasen et al. (1995) has experimentally investigated the relationship between viscoelasticity and barite sag under
static and dynamic conditions. The study indicated that a fragile gel structure does not prevent sagging. A fragile gel is one
that may exhibit high strength but be easily broken once the initial gel breaks. In such cases, the gel structure diminishes
rapidly after the peak gel strength is reached. Therefore, the fluid may have relatively high gel strength but still exhibit severe
dynamic sag when it is exposed to a low shear rate conditions.
Field and laboratory tests showed that maintaining barite in suspension in synthetic and oil based fluids is much
more difficult than in water-based mud (Omland et al., 2004). The ability to keep barite in suspension also depends on the
chemical composition of the water phase. Omland et al. (2004) evaluated and compared the sag tendencies of different
synthetic and oil-based muds. The results showed that the performance of oil-based mud was improved when the traditional
calcium chloride (CaCl2) salt was replaced with other salts. The best performance with respect to sag stability was observed
when an ammonium calcium nitrate (NH4CaNO3) was used as salt. Using NH4CaNO3 enables reduced viscosity coupled with
better sag stability.
Sagging tendency of drilling mud is affected by several factors, including its properties and drilling parameters. The
majority of earlier studies (Hanson et al., 1990; Bern et al., 1998; Jamison & Clement, 1990; Saasen et al., 1995; Kenny &
Hemphill, 1994; Zamora & Jefferson, 1994; Saasen et al., 1991; Dye et al., 1999) prior to 2000 have focused on drilling mud
as the primary cause of barite sag. However, recent studies indicate several additional factors and conditions present during
drilling operations, which influence barite sag. A number of studies (Jamison & Clements, 1990; Jefferson, 1991; Saasen et
al., 1991; Kenny & Hemphill, 1994; Zamora & Jefferson, 1994) indicate that the hole diameter, hole angle, wellbore length,
annular velocity, drillpipe rotation speed, flow regime, mud viscosity, mud gel strength, fluid density, weighting agent
density, particle size and shape, particle concentration and time influence barite sag. Bern (1998) presented a comprehensive
analysis of the effects of different drilling variables. Later, several important findings of Bern were verified by other studies
(Dye et al., 2001; Dye et al., 2002).
Annular fluid velocity has substantial influence on barite sag. Preventing barite sag is more difficult in flowing
fluids than fluid under static conditions (Hanson et al., 1990). Slumping (bed sliding) can initiate at an inclination angle in
the range of 30º and 60º. The highest barite sag rate occurs at inclination angles between 50º and 60º while the effect
diminishes as the inclination angle increases beyond this range. However, other researchers (Bern et al., 1996) reported a
different range of angles (60 to 75 degrees) in which sag is more dominant. Dynamic sag predominantly occurs at low
annular velocities, while its magnitude decreases as velocities increase above 100 ft/min (Dye et al., 2003). Results of flow
loop experiments showed that the effect of hole inclination angle on barite sag is less important than the effect of annular
velocity. A study conducted by Bern (1996) revealed a number of observances of barite sag behavior, including: i)
exacerbation of barite sag at low annular velocities with eccentric and stationary drillpipe; ii) unique characteristics of barite
beds; and iii) more responsiveness of barite beds to mud velocity and pipe rotation than most cuttings beds.
The rotational and axial movements of drillpipe have substantial impacts on barite sag. The rotation of the drillstring
IPTC 14944 3

agitates and moves barite beds into the flow stream, reducing bed height and size. On the other hand, axial movement of the
drillstring may trigger the sliding of the bed.

3. Theory
Particle movement in suspensions depends on fluid and particle properties, and particle concentration. Mechanical friction
and hydrodynamic interference between particles during the settling process affects the terminal velocity of the particle. Non-
Newtonian fluids containing clay (commercial bentonite) and polymer suspensions are often used to transport cuttings during
drilling operations. Often these fluids exhibit a yield stress (τo), which affects the settling behavior of the particles. Slattery

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


and Bird (1961) determined the drag coefficient (Cd) of a sphere by measuring its settling velocity in Carboxyl Methyl
Cellulose (CMC) solution. Valentik and Whitmore (1965) presented extensive data that was later analyzed by Dedegil (1987)
to establish a relationship between drag coefficient and the generalized particle Reynolds number:

vs2 .ρ f , ………………………………………………………………………………………………...… (1)


Re =
τ
The shear stress (τ) is calculated at the representative shear rate (vs / d). Hence, the Reynolds number for Yield Power Law
(YPL) fluid can be expressed as:

v s2 .ρ f , ……………………..……………………………….……………………………..……...… (2)
Re =
τ o + k (vs / d ) n

Drag coefficient of particles in viscous fluid is a function of the particle Reynolds number. Valentik and Whitmore (1965)
developed the following correlations to determine drag coefficient of a spherical particle in YPL fluids.

Laminar Re < 8 Cd = 24/Re, …………….………………….…………………………… (3a)

Transition 8 < Re < 150 Cd = 22/ Re + 0.25, ………………………………………...…………… (3b)

Turbulent Re > 150 Cd = 0.4, ……………………………….………………..……………… (3c)

When yield stress τo > 0, particles below a certain grain size do not settle and remain suspended (Dedegil 1987). In this case,
the stress and buoyancy forces balance the weight of the particle. Thus, for spherical particles, the momentum balance yields:

2
π .d 3 π 
(ρ s − ρ f )g =  d  τ o , ………………………………………..………………………………….............. (4)
6 2 

The critical diameter above which the particle settles in the fluid can be obtained from the above equation. Thus:

3π τ0 , ……………………………………………………………………………………………. (5)
d crit =
2 (ρ s − ρ f )g

Increasing the particle concentration in a fluid elevates the hydrodynamic interference and particle collision (Govier
& Aziz 1972). At low inter-particle distance, one might expect hydrodynamic interference, inter-particle collision, and
interaction to become significant. Rising and falling velocity of uniformly sized particles in laminar motion would be
uniform, and collisions would not be expected even as the concentration increases. Nonetheless, hydrodynamic interference
does exist at low volume fractions (1 to 2 percent); and without agglomeration, the settling velocity would reduce. Thomas
(1963) gives the following empirical correlation for the reduction in settling velocity with the increase in concentration. The
correlation is valid when the ratio Vc/Vo is between 0.08 and 1.0.

Vc
Ln = −5.9c , ……………………………………………………………...…………………………………...… (6)
Vo

where Vc is the terminal settling velocity of particles in a mixture of solids concentration (i.e. volume fraction) c, and Vo is the
terminal settling velocity of a single particle in pure fluid.
4 IPTC 14944

4. Experimental Study
Experiments were carried out to study the sedimentation behavior of barite particles in static and dynamic conditions using a
dynamic sedimentation cell (DSC) which consists of a transparent cylinder with a rotating disc at the top (Fig. 1).
Sedimentation tests were carried out on oil-based muds (OBMs) at different shear rates (rotating disc speeds), and different
temperatures (80°F and 120°F). Two sets of sedimentation tests were conducted varying the pressure sensor configurations.
Test matrices of the experiments are presented in Table 1. Sag experiments were conducted under static (stationary disc) and
dynamic (rotating disc) conditions. Rheologies of fluid samples were measured using a rotational viscometer (Fann 35).
Figure 2 shows flow curves of the test fluids at 80ºF and 120ºF. It can be seen from the figure that the rheology of the OBM

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


samples are roughly the same, except for OBM2-60/80, which has a relatively lower viscosity. This could be due to the
addition of small amount of water to compensate for evaporation of water phase during the experiments and maintain the
density of the sample at the desired value.

2
1

4 5

Fig. 1: Dynamic Sedimentation System (DSS): 1) Modified viscometer; 2) Water bath; 3) Sedimentation cell; 4) Copper coil;
5) Pressure sensors; 6) Rotating disc; and 7) Cell cover

Table 1: Test matrix


Disc Rotation Test Temperature
Configuration
(RPM) 80ºF 120ºF
0 OBM1-0/80 OBM1-0/120 #1
60 OBM1-60/80 OBM1-60/120 #1
0 OBM2-0/80 OBM2-0/120 #2
60 OBM2-60/80 OBM2-60/120 #2
100 OBM2-100/80 OBM2-100/120 #2
IPTC 14944 5

80 60
WBM-0/80 WBM-0/120
70 OBM1-0/80 OBM1-0/120
50
OBM2-0/80 OBM2-0/120

Shear stress, ( lb/100 ft^2)


Shear stress, ( lb/100 ft^2)

60 OBM2-60/80 OBM2-60/120
OBM2-100/80
40
50

40 30

30
20
20

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


10
10

0 0
0 200 400 600 0 200 400 600

Shear rate, (1/sec) Shear rate, (1/sec)

Fig. 2a: Rheologies of test fluids at 80˚F Fig. 2b: Rheologies of test fluids at 120˚F

4.1 Experimental Setup


The Dynamic Sedimentation System (DSS) was designed and developed to measure the pressure profile in the mud sample as
function of time. The system (Fig. 1) consists of: i) sedimentation cell with pressure sensors; ii) cell cover (lid); iii) rotating
disc to shear the mud; iv) water-bath with copper-coil; v) data acquisition system to record and display test measurements;
and vi) modified viscometer to rotate the disc at the required rotational speed. The sedimentation cell is made of a
transparent acrylic cylinder that has 6" height, 5.5" inner diameter (ID) and 6.5" outer diameter (OD). Five pressure sensors
were installed on the cell wall. Figure 3 shows two different sensor configurations, which were used during the
investigation. In the first configuration (Fig. 3a), sensors were placed in the left and right side of the cylinder with 0.75" gap
between two consecutive sensors. The sensors measure pressure up to 1 psig (27 inch of water column) with maximum error
of ±1% of reading. Configuration #1, requires 1.65 liters of fluid to conduct the tests. After completion of the first set of
experiment, the cell was redesigned and sensors were reconfigured (Fig. 3b) to measure the pressure at the bottom of the cell.
The bottom plate of the cell was moved up 1.75" from the original position to allow the installation of a sensor on the bottom
side as shown in Fig. 3b. This configuration required 1.17 liters of test fluid sample. The configuration change was
introduced to increase the shear rate without increasing the disc speed.

4 5 6 4 5 6

7 7

3 3
0.25" 0.25"
0.88"

Legend:
1.50" 1.63" 1 Test cell
2.25" P4
2 Pressure sensor
1.50"
P4 P3 3 Rotating disc
1.50" 1.50" 4 Test cell cover (Lid)
P3 P2
1.50"
5 Modified viscometer cup (Sleeve)
P2 2 P1 6 Circular hole on the Lid
1.50"
0.50" 7 Test Fluid
P1 1
P0
8 Pressure Sensor Installation Hole
8
2 P0

0.50" 5.50"
1
8
5.50"

(a) (b)

Fig. 3: Sensor configurations used during the study: a) Configuration #1; and b) Configuration #2

The cell cover (lid) is designed to minimize water evaporation and heat loss from the sample. The lid has a 2.5-inch
circular hole at the center, which is used as a passage for the rotating disc components. It allows free rotation of the disc,
which consists of a 5-inch diameter acrylic disc and a 2-inch diameter acrylic tube attached precisely to the center of disc.
The tube is connected to a rotational viscometer (OFITE Model 800 Viscometer). To avoid agitation while shearing, parts of
the disc are assembled and properly aligned. A ¼-inch copper-coil (heating tube) installed inside the cell circulates hot water
to control the mud temperature. During the experiments, pressure measurements were recorded using a data acquisition
system. Pressure sensor readings were calibrated before each experiments using water as a calibration fluid.
6 IPTC 14944

4.2 Test Procedure


The test procedure of for the dynamic sedimentation experiments involves a number of steps including preparation of the
mud sample, heating test sample, measuring rheology and density, calibrating the pressure sensors, running the sedimentation
test, and measuring final properties of test fluid. A detailed test procedure includes the following steps:

Step I. Preparation of the mud sample: Prior to running each test, the mud sample was mixed thoroughly using a low-speed
agitator for about 20 minutes. This is to sufficiently homogenize the fluid and distributed barite particles uniformly
in the sample.

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


Step II. Preheating the mud: Fluid samples were placed in an oven to heat the mud to the test temperatures (80°F and
120°F). During the heating process, the mud temperature was monitored using two thermometers, and the sample
was agitated periodically to maintain a uniform temperature distribution.
Step III. Measuring initial fluid properties: After preheating the mud, the rheology and density of the samples were measured.
Step IV. Calibration of pressure sensors: In order to ensure the reliability of measurements, all the pressure sensors were
calibrated before each test. First, the sensors were calibrated against the atmospheric pressure (i.e. with completely
empty sedimentation cell). Then, the cell was completely filled with tap water and calibration was repeated. In both
calibration steps a Visual Basic (VB) program was run to update calibration curve parameters of each sensor.
Step V. Running the test: Immediately after filling the cell with the sample, the test begins by selecting the desired disc
rotation speed and turning on the modified viscometer while the VB computer program displays and records test
measurements. The sample temperature was maintained at a constant level during the test by circulating hot water
through the copper coil that is place inside the cell.
Step VI. Measuring final sample properties: At this stage, the mud sample was removed from the cell and its density and
rheology were measured and recorded as final measurements.

4.3 Test Results


Figure 4 shows the raw data obtained from the data acquisition system. Pressure readings measured at different fluid levels
reduce with time due to barite sag. Attempts were made during the experiments to minimize the variations in fluid level
between tests. However, minor discrepancies in initial pressure measurements from different tests implicate slight fluid level
variations. To compare results of different tests, the measurements are presented in terms of normalized pressure (i.e. the
ratio of measured pressure at any time to the initial pressure measurement) and normalized density (i.e. the ratio of measured
density at any time to the initial density). Appendix A presents derivations of normalized density equations for fluid layers
(Fig. A-1) between to consecutive pressure sensors. The use of the normalized pressure minimizes the effects of initial
measurement errors in the data analysis. Finally, the change in density at different levels (pressure sensors positions) of test
samples during the experiment is calculated using the measured pressures.

P0 P1 P2 P3 P4

1600

1200
Pressure (Pa)

800

400

0
0 400 800 1200 1600 2000

Time (min.)

Fig. 4: Measured pressure vs. time at different levels in the sample at 120ºF and 60 rpm disc rotation speed

Results of the first set of OBM experiments are presented in Figs. 5 and 6 as normalized pressure versus time.
Pressure reduction at P4 sensor position is the highest. The rotation of the disc increased the pressure reduction and barite
sag. Particularly, pressure reduction (Fig. 6b) was severe when both temperature and disc rotation speed were increased
(120ºF and 60 rpm) simultaneously. After completion of the test, a clear oil phase was observed at the top of mud sample,
especially during high-speed and high-temperature tests (i.e. OBM1-60/120). Significant pressure reduction occurred due to
the barite sag.
IPTC 14944 7

Normalized P0 Normalized P2 Normalized P4 Normalized P0 Normalized P2 Normalized P4

1.1 1.1

N ormalized Pressure

N ormalized Pressure
1 1

0.9 0.9

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


0.8 0.8

0.7 0.7
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500 3000
Time (min.) Time (min.)

(a) (b)

Fig. 5: Normalized pressure vs. time for OBM1 at 80ºF: a) 0 rpm; and b) 60 rpm

Normalized P0 Normalized P2 Normalized P4 Normalized P0 Normalized P2 Normalized P4

1.1 1.1
Normalized Pressure

Normalized Pressure
1 1

0.9 0.9

0.8 0.8

0.7 0.7
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500 3000
Time (min.) Time (min.)

(a) (b)

Fig. 6: Normalized pressure vs. time for OBM1 at 120ºF: a) 0 rpm; and b) 60 rpm

Figures 5 and 6 show the effects of temperature and disc rotation speed on the pressure distribution in the cell. The
figures present the normalized pressure as a function of time for three different depths. The other two sensor readings are not
displayed here to avoid overlapping of data points. The measured pressures reduce with time due to the sedimentation of
barite particles that decreases the density of the fluid. Figures 5a and 5b present the results of two tests with different disc
rotation speeds (0 rpm, and 60 rpm) at 80ºF. The highest-pressure reduction is observed in the upper part of the mud sample
(i.e. above P4 sensor). Similar pattern of pressure variations are shown in Fig. 6; however, at 60 rpm disc speed, the pressure
reduces slightly at higher rate than during the test without disc rotation. Comparisons of different temperature measurements
show that the rate of pressure reduction (i.e. barite sag) increases when the temperature is raised under dynamic conditions.
In addition to the first batch of OBM tests, an additional six oil-base mud (OBM2) experiments were conducted after
the test cell was modified. Experiments were carried out under similar conditions as OBM1 tests. Two additional tests were
conducted at 100 rpm disc rotation to examine high. Figures 7 and 8 present normalized pressure for OBM2 at different disc
rotation speeds and temperatures. Like the results of OBM1, pressure reduction at P4 sensor location is higher than the
pressure reduction at P2 and P0 sensor locations. The pressure reduction at P2 position is greater than the pressure reduction at
P0 position. Figure 7 illustrates the effect of disc rotation on pressure depletion at 80ºF. Comparing Figs. 7a, 7b, and 7c, the
increase in disc speed facilitates pressure reduction in the top layers of the mud sample. In contrast, the disc speed does not
substantially affect the bottom part of mud sample, and the pressure at P0 sensor remains almost constant. Figure 8
demonstrates the effect of increasing the temperature on rate of pressure change at different disc rotation speeds. The result
shows that the increase in temperature facilitates pressure reduction. Particularly at P4 sensor, it can be seen that the pressure
reduction is significantly enhanced by the rotation of the disc. At the end of the tests, clear oil phase was found at the top of
the mud samples. This indicates complete sedimentation of the barite particles from the top layer (Layer 1). The final density
of mud sample in Layer 1 can be estimated using the pressure reading at the end of test.

P4
ρ1 = ρ mud −balance
P4∗ , …………….………….…………………………………………………………..………… (7)
8 IPTC 14944

where P4 and P4* are the final and initial pressure readings of P4 Table 2: Calculated final mud densities Layer 1
sensor. The pressure readings are converted to densities and presented Mud sa mple Density [lbm/gal]
in Table 2. The results show significant density reduction in the top OBM2- 0/80 13.3
layer of the mud samples. For high-temperature and high-shear tests OBM2- 60/80 13.6
(OBM2-100/120), the density reduced from 14.9 ppg to 9 ppg, which is
OBM 2- 100/80 13.2
slightly greater the unweighted base fluid density (7.7 ppg).
OBM2- 0/120 11.7
OBM 2- 60/120 9.9

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


OBM2- 100/120 9

Normalized P0 Normalized P2 Normalized P4 Normalized P0 Normalized P2 Normalized P4

1.1 1.1

Normalized Pressure
1
Normalized Pressure

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6
0 400 800 1200 1600 0 400 800 1200 1600

Time (min.) Time (min.)

(a) (a)
Normalized P0 Normalized P2 Normalized P4 Normalized P0 Normalized P2 Normalized P4

1.1 1.1

1
Normalized Pressure
N ormalized Pressure

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6
0 400 800 1200 1600 0 400 800 1200 1600
Time (min.) Time (min.)

(b) (b)
Normalized P0 Normalized P2 Normalized P4 Normalized P0 Normalized P2 Normalized P4

1.1 1.1
Normalized Pressure

1
Normalized Pressure

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6
0 400 800 1200 1600 0 400 800 1200 1600
Time (min.) Time (min.)

(c) (c)

Fig. 7: Normalized pressure vs. time plot for OBM2 at 80ºF: Fig. 8: Normalized pressure vs. time plot for OBM2 at 120ºF:
a) 0 rpm; b) 60 rpm; and c) 100 rpm a) 0 rpm; b) 60 rpm; and c) 100 rpm
IPTC 14944 9

5. Analysis of Results and Discussions


Sedimentation of barite particles in the drilling mud results in density variations in the vertical axis. Test results and
visual observations show the movement of barite particles from the top part of the sample to the bottom. To determine the
density change occurring in the mud sample, normalized density equations have been developed and presented in Appendix
A. Figure 9 shows measured normalized densities of different layers in the mud sample as a function of time for OBM2-
100/120 test. The highest density reduction was observed in the top layer (Layer 1), and the highest density increase was
observed in the bottom layer (Layer 5). Densities of intermediate layers (Layer 2, Layer 3, and Layer 4) are approximately
constant. Despite some exceptions, most of the test results follow the pattern show in Fig. 9 and meet the general

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


expectations that come from previous discussions. Increasing the shear rate and temperatures improved barite sedimentation.
However, temperature has more effect on density variation than disk rotation speed.

Normalized r1 Normalized r2 Normalized r3 Normalized r4 Normalized r5

1.2

1.1
Normalized density

1.0

0.9

0.8

0.7

0.6
0 200 400 600 800 1000 1200 1400 1600
Time (min.)

Fig. 9: Measured normalized density vs. time for OBM2 at 120ºF and 100 rpm

In order to compare the level of barite sag under different test conditions, normalized density of the top layer is
presented as a function of time for each set of experiments. Figure 10 shows the variation of normalized density of the first
layer during the experiments with OBM1 and OBM2. Significant density reduction was observed in the top layer.
Normalized density measurements were not stabilized, indicating the continuation of sedimentation process even after 1500
minutes. The OBM2-0/80 test has the least density reduction while OBM2-100/120 exhibits the highest. Increase in disc
rotation speed had less effect on the barite sag compared to the temperature increase. This could be due to significant oil-
phase viscosity reduction resulting from the temperature increase, which leads to a higher rate of particle sedimentation and
density decrease in the top layer.

OBM1- 0/80 OBM1- 60/80 OBM1- 0/120 OBM1- 60/120 OBM2- 0/80 OBM2- 60/80 OBM2-100/80
OBM2-0/120 OBM2-60/120 OBM2-100/120
1.1
1.1
Normalized density layer 1

1
Normalized density layer 1

1
0.9
0.9
0.8
0.8
0.7
0.7
0.6 0.6
0.5 0.5

0.4 0.4
0 500 1000 1500 2000 2500 3000 3500 0 500 1000 1500 2000
Time (min.) Time (min.)

(a) (b)

Fig. 10: Normalized density of the top layer vs. time at different temperature and disc rotation speeds: a) OBM1; and b) OBM2

Settling velocity of particles in suspension strongly relates to the viscosity of the fluid and hydrodynamic
interactions between particles. For non-Newtonian fluids, the viscosity is a function of the shear rate. In order to calculate the
effective viscosity (i.e. representative viscosity) of a sample in the test cell, an equation for the average shear rate has been
developed (see Appendix B). Table 3 presents the primary average shear rate in the test cell for different rotation speeds. The
level of deformation rate resulting from particle sedimentation is negligible compared to the primary share rate. Hence,
apparent viscosities (Table 4) of the samples during OBM2 test are evaluated based on the primary share rate. Without the
10 IPTC 14944

disc rotation, the apparent viscosity is a function of the settling shear rate (Vs/d). Therefore, the apparent viscosity is expected
to be very high as the settling velocity of barite is very small or close to zero.

Table 3: Average primary shear in the test cell


Rotation Speed Test Fluid
Unit
(rpm) OBM1 OBM2

60 0.35 0.49 1/s

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


100 0.58 0.82 1/s

Table 4: Apparent viscosity of OBM2 at 80ºF and 120ºF


Disc rotation (rpm) Shear rate (1/s) µapp at 80ºF (cp) µapp at 120ºF (cp)

0 - very high (>> 16.7) very high (>>13.1)

60 0.49 16.7 13.1

100 0.82 10.1 8.1

Figure 11 presents apparent viscosities of the samples at two different temperatures. The results indicate that the
apparent viscosities of test samples reduce with increasing the temperature and disc rotation speed. Apparent viscosities of
the mud during 60 rpm and 100 rpm tests reduced by 21.5% and 19.8%, respectively, when heated from 80ºF to 120ºF. The
increase in disc rotation speed also has a similar impact on mud viscosity. Increasing the disc speed from 60 rpm to 100 rpm
at 80ºF and 120ºF reduced the apparent viscosity of the fluid by 39.5% and 38.2 %, respectively. According to this analysis,
the increase in rotation speed is expected to have more influence in reducing the apparent viscosity than the raise in
temperature. However, experimental results showed dominant effect of temperature on barite sag. One possible explanation
for the unexpected sag results could be due to the large reduction in localized viscosity of the oil phase as the temperature
increases, although the bulk viscosities of the mud samples show only moderate reduction. Often the viscosities of structured
fluids such as foam, emulsion and invert-emulsion have less temperature sensitivity than the viscosities of their components.

OBM2-80 deg F OBM2-120 deg.F

18

16
60rpm
Apparent viscosity (cp)

14

12

10
100rpm
8

4
2

0
0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9
Primary shear rate(1/sec.)

Fig. 11: Apparent viscosity vs. primary shear rate of OBM2 tests

Test fluids (Fig. 2) show strong non-Newtonian behavior and rheology measurements indicate a yield stress at low
shear rates. The critical particle diameter is calculated according to the method presented in Section 3. Results show that the
particles with a diameter less than 1.62 mm are not expected to settle in the mud samples. The barite particles have diameter
ranging from 3 to 74 micro-meters, which is significantly less than the critical diameter. Therefore, barite particles are
theoretically expected to suspend indefinitely in the fluid. However, this opposes the experimental observations. This could
be again associated with the disparity between the localized and bulk viscosities of the mud system. For invert-emulsions
(oil-based muds) systems, localized viscosity of the fluid may be different from the bulk viscosity, which is measured directly
from a viscometer.
IPTC 14944 11

6. Conclusions
Barite sag behavior of oil-based mud was investigated experimentally under both static and dynamic conditions. The
following conclusions can be drawn from this investigation:

• Both temperature and shear rate have strong influences on the sedimentation of barite particles. Increases in
temperature and shear rate facilitate the sagging process by reducing the viscosities of the suspension and its
components. However, dynamic environments, which produce constant agitations such as extremely high shear rate
conditions can minimize barite sag. The turbulence created by the dynamic environment dominates the viscosity

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


reduction due to shearing.

• The increase in shear rate has more pronounced effect on the bulk viscosity of the mud than the raise in temperature.
However, results of sedimentation tests show that the increase in temperature has more pronounced impact than the
increase in shear rate. Therefore, the barite sag is more likely associated with the viscosity of each phase or
localized viscosity rather than the bulk viscosity.

Acknowledgments
We wish to express our appreciation to the University of Oklahoma and Mewbourne School of Petroleum and Geological
Engineering for their support. We also would like to thank M-I SWACO for providing the test fluids used during the
investigation.

Nomenclature
c = solid concentration Greek Symbols
d = Particle diameter γ& = Shear rate
dcrit = Critical diameter = Apparent viscosity
µ app
g = Acceleration due to gravity
h = Height of fluid, in ρ = Density
K = Consistency index ρ *
= Initial density
n = Fluid behavior index
p = Pressure ρ = Normalized density
P* = Initial pressure reading ρf = Fluid density
r = Radius = Liquid density
ρl
R = Radius of disc
Re = Particle Reynolds number ρs = Solid density
t = Time τ = Shear stress
T = Temperature, τo = Yield stress
Vc = Settling velocity of particles in suspensions φ = Volume fraction of dispersed phase
Vo = Settling velocity of particles in clear fluid ω = Angular velocity
vs = Terminal settling velocity
vt = Tangential velocity

References
Bern, P.A., Zamora, M., Slater, K.S., and Hearn, P. J. 1996. The Influence of Drilling Variables on Barite Sag, paper SPE
36670 presented at the Annual Technical Conference and Exhibition, Denver, Colorado, 6-9 October.
Bern, P.A., Oort, E., Neustadt, B., Ebeltoft, H., Zurdo, C., Zamora, M., and Slater, K.S.:1998. Barite Sag: Measurement,
Modeling, and Management, paper SPE 47784 presented at the IADC/SPE Asia Pacific drilling Technology
Conference held in Jakarta, September.
Boycott, A.E. 1920. Sedimentation of Blood Corpuscles, Nature, Vol. 104, 532.
Dedegil, M.Y. 1987. Drag coefficient and Settling Velocity of Particle in Non-Newtonian Suspensions. Institute fur
Fordertechnik.Abt Siromungsfordertechnik. University of Karlsruhe (TH), Federal Republic of Germany Journal of
Fluids Engineering, September, Vol. 109/319.
Dye, W., Hemphill, T., Gusler, W., and Mullen, G. 1999. Correlation of Ultra-low Shear Rate Viscosity and Dynamic
Barite sag in Invert-Emulsion Drilling fluids, paper SPE 56636 presented at SPE Annual Technical Conference and
Exhibition held in Houston, Texas, October.
Dye, W., Hemphill, T., Gusler, W., and Mullen, G. 2001. Correlation of Ultra-low Shear Rate Viscosity and Dynamic Barite
sag, SPE 70128, SPE Drilling and Completion, March.
Dye, W., and Mullen, G. 2002. New Technology to Manage Barite Sag, AADE-02-DFWM-HO-12, presented at the
American association of Drilling Engineers 2002 Technology Conference, Houston, Texas 2-3 April.
Dye, W., Mullen, G., and Gusler, W. 2003. Drilling Processes: The Other Half of the Barite Sag Equation paper SPE 80495
12 IPTC 14944

presented at the SPE Asia Pacific Oil and Gas Conference and Exhibition held in Jakarta, Indonesia, April.
Famularo, J., and Happel, J. 1965. A.I.Ch.E. Journal, 11, 981.
Govier, G.W., and Aziz, K. 1972. The Flow of Complex Mixtures in Pipe. Van Nostrand Reinhold Company, New York. pp.
13-19.
Hanson, P.M., Trigg, Jr. T.K., Rachal, G., and Zamora, M. 1990. Investigation of Barite Sag in Weighted Drilling Fluids in
Highly Devted well, paper SPE 20423 presented at the 65th Annual Technical Conference and Exhibition of the
Society of Petroleum Engineers held in New Orleans, LA, September.
Jamison, D.E., and Clements, W.R. 1990. A New Test Method to Characterize Settling/Sag tendencies of drilling Fluids

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


Used in Extended Reach Drilling Technology Symposium.
Kenny, P., and Hemphill, T. 1994. Unique Hole Cleaning Capabilities of Ester-Based Drilling Fluid Systems, SPE 28308
presented at the 1994 SPE Annual Technical Conference, New Orleans Sep. 25-28.
McLean, M.R., and Addis, M.A. 1996. Wellbore Stability: The Effect of Strength Criteria on Mud Weight
Recommendations, paper SPE 20405 presented at Annual Technical Conference and Exhibition, New Orleans, 23-
26 September.
Omland, T., Albertsen, T., Taugbol, K., Saasen, A., Svanes, K., and Amundsen, P. 2004. The Effect of the Synthetic-and
Oil- Based Drilling Fluid’s Internal Water Phase Composition on Barite sag, paper SPE 87135 presented at the
IADC/SPE Drilling Conference, Dallas, Texas, U.S.A. March.
Saasen, A., Marken, C., Sterri, N., and Jakobsen, J. 1991. Monitoring of Barite Sag Important in Deviated Drilling, OGJ,
Aug.
Saasen, A., Liu, D., Marken, C.D., Sterii, N., Halsey, G.W., and Isambourg, P. 1995. Prediction of Barite Sag Potential of
Drilling Fluids from Rheological Measurement, paper IADC/SPE 29410 presented at the SPE/IADC Drilling
Conference Held in Amsterdam, February.
Saasen, A. 2002. Sag of Weight Materials in oil Based Drilling Fluids, paper IADC/SPE 77190 presented at the IADC/SPE
Asia Pacific Drilling Technology held in Jakarta, Indonesia, September.
Slattery, I.C., and Bird, R.B. 1961, Non Newtonian Flow Past a sphere, Chem. Eng. Science, Vol. 16 1961, pp. 231-241.
Thomas, D.G. 1961-1964, A.I.Ch.E. Journal, 7, 423 (1961); 9, 310 (1963); 10, 303 (1964).
Valentik, L., and Whitmore, R.L. 1965. The Terminal Velocity of Spheres in Bingham Plastics, Brit. J. Appl. Phys., Vol. 16,
pp. 1197-1203.
Zamora, M., and Jefferson, D.T. 1994. Controlling Barite Sag can Reduce Drilling Problem, OGJ, Feb. 1.
IPTC 14944 13

Appendix A: Normalized Density Equations


This appendix presents the derivation of the normalized density equation for each layer in the sedimentation test cell (Fig. A-
1). In general, the gauge pressure in the fluid at any depth (h) can be determined as:

P = ρ .g .h ….……………………..………..... (A-1)

The pressure difference (∆P) between two consecutive sensors


can be calculated as:

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


∆h1
ρ1
∆P = ρ .g .∆h ………..…………….…………. (A-2) P4

ρ2 Layer 2 Height (h)


∆h2

where ∆P is the pressure difference; g is the gravitational P3

acceleration; and ∆h is the distance between two consecutive Layer 3 ρ3 ∆h3

sensors. The average density (ρ) of fluid between two P2

consecutive sensors is expressed as: ∆h4 ρ4 Layer 4


P1

1 Layer 5 ρ5
.∆P = k .∆P …………..…………….. (A-3)
∆h5
ρ=
g.∆h
P0

where k is a constant. The recorded pressures are gage


pressures, thus the pressure above first layer is 0 psig. Fig. A-1: Sedimentation test cell
Therefore, the density of first (top) layer can be determined as:

ρ1 = k ( P4 − 0) ………………………………...…………….……………………….…………………………. (A-4)

And initial density of first layer (at t = 0) would be:

ρ1* = k ( P4∗ ) ……………………….………………………………………………………….………………….. (A-5)

where: ρ1* and P4* are the initial density of Layer 1 and initial pressure reading of P4 sensor. Then, the normalized density of
first layer is expressed as:

_
ρ 1 P4
ρ1 = =
ρ 1* P4* ……………………...…………………………………………..…………………………...….. (A-6)

Similarly, the densities of other layers can be estimated:

_
ρ2 P − P4
ρ2 = = 3∗
ρ 2 P3 − P4*
*
…..……………………………………………………………………….……….….……. (A-7)
_ ρ P −P
ρ 3 = 3* = 2∗ 3*
ρ 3 P2 − P3 …..…………………………………………………………………………………...……. (A-8)
_
ρ P −P
ρ 4 = 4* = 1∗ 2*
ρ 4 P1 − P2 ….………………………………………………………………………………...………. (A-9)
_ ρ P −P
ρ 5 = 5* = ∗0 1*
ρ 5 P0 − P1 …..……………………………………………………………………..…………...……. (A-10)

After determining the normalized density of each layer, the density of any layer can be determined as:
_
ρ = ρ .ρ mud −balance
………………………………………………………………………..…………...……… (A-11)

where ρmud-balance is the initial mud density measured using a mud-balance.


14 IPTC 14944

Appendix B: Derivation of Primary Shear Rate Equation

Tangential velocity at any point on the surface of the rotating Side view of the rotating Top view of the rotating
disc
disc is given as:
ω
v
0.25”
vt = ω .r ……………………...…..……………… (B-1) dr

Then, the shear rate at any point is determined as: Height (h) R

Downloaded from http://onepetro.org/IPTCONF/proceedings-pdf/11IPTC/All-11IPTC/IPTC-14944-MS/1668209/iptc-14944-ms.pdf by Saint Francis University user on 02 March 2021


• du vt ………………...……….…..……… (B-2)
γ= =
dy h

Combining Eq. (B-1) and Eq. (B-2), the shear rate can be written
as a function of radial distance:
Fig. B-1: Top and side view of rotating disc
• ω .r ….……………………………………………………………………….……………….………… (B-3)
γ (r ) =
h

The share rate varies from the center of the cylinder to the wall. The average shear rate can be estimated as:

R
 ω.r  ω 2
• ∫  h dr (2πr ) 2π h ∫0 r dr …………………………………………....…………………..……........... (B-4)
γ ave = =
A π .R 2

After simplifying Eq. (B-4), the average shear rate is given as:

• 2 ω .R ...……………………………………..………………………………………...…….…………… (B-5)
γ ave =
3 h
where:
vt is the tangential velocity;
ω is the angular velocity (rpm);

γ is the shear rate (1/sec.);
r is the radial distance from the center of the disc.
R is the radius of disc, 2.5 in.
h is the distance between disc and the bottom of the cell.

You might also like