Journal of Petroleum Science and Engineering: A C A C B C B C B C

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Annular displacement in a highly inclined irregular wellbore: Experimental T


and three-dimensional numerical simulations
Hans Joakim Skadsema,c,∗, Steinar Kragseta,c, Bjørnar Lundb,c, Jan David Ytrehusb,c,
Ali Taghipourb,c
a
International Research Institute of Stavanger, P. O. Box 8046, 4068, Stavanger, Norway
b
SINTEF Industry, P. O. Box 4760 Torgarden, NO-7465, Trondheim, Norway
c
DrillWell, P. O. Box 8046, 4068, Stavanger, Norway

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

Keywords: Primary cementing of well casings is the operation where drilling fluid is displaced from the annulus between the
Displacement casing and the drilled rock formation and replaced by a cement slurry. The annulus to be cemented is normally
Annulus narrow and eccentric, and the formation wall may exhibit irregular geometric features such as local enlarge-
Viscoplastic fluids ments or washouts, originating from the drilling process and varying geological properties. Eccentricity and the
Geometry
presence of wellbore irregularities can lead to residual, non-displaced drilling fluid or mixing of the cement
Washout
slurry with other wellbore fluids, causing contamination and failure to hydraulically isolate the annulus outside
the casing. We study non-Newtonian fluid displacement in an irregular annulus experimentally and numerically
for concentric and eccentric configurations of the inner pipe, and for horizontal and slanted orientations of the
annulus. An overgauge section simulates a washed out zone in the wellbore. Eccentricity promotes fluid flow in
the wide sector of the annulus, leading to an axial elongation of the interface between the two fluids. Depending
on eccentricity and inclination, residual non-displaced fluid remains in different parts of the hole enlargement,
while inclination from the horizontal is found to improve the fluid displacement in the enlarged section. The
experiments and simulations agree well for the tests where the annulus is slanted, with simulations replicating
displacement trends and arrival times measured in experiments. The quantitative difference between experiment
and simulation is larger for the horizontal and concentric annulus tests. We discuss modelling assumptions and
experiment uncertainties that can explain the observed differences.

1. Introduction before the cement slurry (Nelson and Guillot, 2006).


The casing that is to be cemented in the well is generally not per-
Primary cementing of casings is a critical well construction opera- fectly centered in the wellbore. Casing eccentricity leads to non-uni-
tion where cement slurry is injected into the annular space between the form flow velocity over the annulus cross-section and an axially elon-
casing and rock formation. The hardened well cement is an important gated interface between fluids during displacement. Yield stress fluids
barrier element that should provide hydraulic isolation between for- such as drilling fluids can cease to flow altogether on the narrow side
mation layers and ensure mechanical support of the casing. Achieving (Walton and Bittleston, 1991; Szabo and Hassager, 1992) or in parts of
hydraulic isolation and a uniform annular cement sheath outside the the annulus where the fluid velocity is low, such as inside a washed out
casing is dependent on the mobilization and displacement of the dril- section of the well (Roustaei and Frigaard, 2013; Roustaei et al., 2015).
ling fluid that occupies the annular space initially before cementing. Possible consequences of incomplete drilling fluid displacement include
Drilling fluids are non-Newtonian fluids with rheological properties residual films of drilling fluid that can act as annular leakage paths past
designed to transport cuttings and rock fragments to the surface. the cement or chemical contamination and degraded cement quality.
Further, under static conditions most drilling fluids develop gel strength The effects of buoyancy, rheology, inertia and geometric effects such as
that needs to be exceeded in order to initiate flow. Displacement of such eccentricity and inner string rotation on annular displacement flows
fluids from a narrow annulus is a challenging task, normally attempted have been studied extensively in experiments and numerical simula-
by injecting a sequence of carefully designed preflushes or spacer fluids tions previously, in no small part due to the industrial importance of


Corresponding author. International Research Institute of Stavanger, Prof. Olav Hanssensvei 15, 4021 Stavanger, Norway.
E-mail address: hans.joakim.skadsem@norceresearch.no (H.J. Skadsem).

https://doi.org/10.1016/j.petrol.2018.09.007
Received 15 June 2018; Received in revised form 28 August 2018; Accepted 3 September 2018
Available online 07 September 2018
0920-4105/ © 2018 Elsevier B.V. All rights reserved.
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

such flows and the need to design and optimize mud displacement and The experiments reported by Malekmohammadi et al. focused on low
primary cementing. The purpose of this paper is to study the effect of a Reynolds number flows and their results largely support guidelines for
hole enlargement on the annular fluid displacement in a narrow an- primary cementing, namely that positive ratios of viscosity and mass
nulus, a topic that has not been extensively explored previously. density and smaller eccentricity promote stable vertical displacement.
The steady axial single phase flow of non-Newtonian yield stress The experiment results suggested that a narrow gap Hele-Shaw mod-
fluids has been studied analytically for concentric geometries elling of the displacement process is relevant, but also the existence of
(Fredrickson and Bird, 1958; Hanks, 1979), eccentric geometries dispersive effects not captured in a Hele-Shaw model
(Walton and Bittleston, 1991; Szabo and Hassager, 1992) and con- (Malekmohammadi et al., 2010). A more recent experimental in-
centric geometry with inner pipe rotation (Bittleston and Hassager, vestigation into annular non-Newtonian fluid displacements is reported
1992). In general, numerical methods have been used to investigate the by Ytrehus et al. (2017) who studied the effect of inclination, eccen-
effect of a rotating inner string on the steady single phase flow in an tricity and inner string rotation on annulus displacement. Their results
eccentric annulus (Meuric et al., 1998). A comprehensive bibliography indicate that inner string rotation improves displacement and that the
of analytical, numerical and experimental studies can be found in shape of the fluid interface is significantly affected by buoyancy.
Escudier et al. (2002). Such single phase flow results are particularly The works cited above have contributed to the current under-
relevant for circulation and conditioning of the drilling fluid before standing of non-Newtonian fluid displacement in regular and eccentric
displacement and cementing in a regular annular geometry. annuli. The drilling process itself may generate irregular geometric
Cement slurry is usually pumped down the well inside the casing, features along the wellbore, e.g. due to excessive lateral drill string
separated from the drilling fluid by a sequence of spacer fluids and vibrations (Plácido et al., 2002), the presence of weakly consolidated
mechanical plugs. After passing the casing shoe, the slurry enters the formation layers, or due to hydraulic erosion of the wellbore wall.
annulus to be cemented and begins flowing up toward the surface. The These washed out zones will generally represent an overgauge section
possibly density-unstable displacement flow inside the casing on the of the wellbore where the flow velocity decreases and there is a risk of
way down the well has been studied extensively by Taghavi et al. and static, residual drilling fluid in parts of the washout where the shear
Alba et al., see e.g. Refs. (Taghavi et al., 2012; Alba et al., 2013a, stress is low. It is only recently that the single phase flow of a Bingham
2013b). Annular displacement flows have also been studied extensively fluid in a channel with irregular geometrical features was investigated.
both numerically and experimentally. In most primary cementing op- Utilizing a numerical method based on variational principles, Roustaei
erations, the annulus to be displaced and cemented is narrow with an et al. (2015) studied the non-inertial flow of a Bingham fluid through a
annulus gap size of the order of centimetres. The azimuthal length is channel with different geometric enlargements, while Roustaei and
typically of the order of a meter, while several tens, even hundreds of Frigaard (2015) considered inertial, laminar flows in a channel with an
meters in the axial direction are to be displaced and cemented. These enlargement. These studies were motivated by the flow of yield stress
specific geometric features of the annulus have motivated displacement wellbore fluids past enlarged hole sections which is particularly re-
models that average physical quantities such as fluid concentration and levant for optimizing conditioning of drilling fluids ahead of displace-
velocity across the gap while retaining the full dynamics in the azi- ment and cementing. Their results show the existence of static fluid
muthal and axial dimensions. Bittleston et al. (2002) used scaling ar- inside deep washouts, and there is a nonlinear relation between the
guments to derive a two-dimensional gap-averaged model from the full Reynolds number and the circulating area of the washout. The simu-
three-dimensional (3D) equations. Pelipenko and Frigaard used this lations are conducted using a two dimensional computational domain
model to investigate the existence of steady, laminar displacements in for axisymmetric flows.
near-vertical wells (Pelipenko and Frigaard, 2004a, 2004b) and com- Relatively few studies have addressed displacement flows and ce-
pared model results to industry-accepted guidelines for effective la- menting of irregular annuli. Early work by Clark and Carter (1973) and
minar displacements (Pelipenko and Frigaard, 2004c). Carrasco-Teja Zuiderwijk (1974) as well as the more recent cementing experiments of
et al. investigated viscoplastic displacement flows in narrow, horizontal Kimura et al. (1999) did include washout geometries, but these studies
annuli with fixed casing wall (Carrasco-Teja et al., 2008) and in the are mainly oriented toward yard testing and provide limited insight into
presence of casing rotation (Carrasco-Teja and Frigaard, 2010). Com- the physics of fluid displacement in such geometries (Frigaard et al.,
putationally efficient implementations of the model in Ref (Bittleston 2017). Very recently however, experimental results have been reported
et al., 2002). can be used to simulate fluid displacement and cement by Lund et al. (2018) and combined experimental and simulation re-
placement over hundreds of meters along the annulus (Guillot et al., sults by Renteria et al. (2018).
2007; Bogaerts et al., 2015). Casing rotation and reciprocation was The purpose of the current paper is to extend the above work with
incorporated in the gap-averaged model by Tardy and Bittleston fully three-dimensional numerical simulations in order to study the
(2015), while casing centralization and fluid mixing inside the casing fluid displacement related to a model washout in detail. We present
were coupled to the annulus displacement model by Tardy et al. (2017). experimental measurements of the fluid displacement interface in the
Generalization of the model to transitional and fully turbulent flows proximity of a model washout in a medium-scale annulus, and compare
was performed by Maleki and Frigaard (2017). Recently, Tardy pre- these to 3D numerical simulations using an open source computational
sented a further development where the narrow annulus assumption is fluid dynamics software. Experiments have been carried out in the
utilized to maintain the two dimensionality of the pressure equation, annulus flow loop described in Refs (Ytrehus et al., 2017; Lund et al.,
but where now fluid volume fractions are resolved in all three dimen- 2018). where a model hole enlargement simulates a washed out seg-
sions (Tardy, 2018). ment of the wellbore. Within the simulated washout, the annulus is no
Laboratory measurements of non-Newtonian displacement flows in longer narrow, and the flow becomes fully three-dimensional. This
annuli were presented by Tehrani et al., 1992, 1993 Their experiments motivates a full 3D numerical solution of the flow for which we have
were carried out in a 3 m long laboratory flow loop and with an annulus used OpenFOAM (OpenFOAM). The numerical simulations are subse-
where eccentricity and inclination could be varied. The experiments quently used to extract the displacement efficiency and the location of
revealed the existence of stable azimuthal flows from the wide to the residual, non-displaced fluid in the hole enlargement for different ex-
narrow side also in a vertical annulus, driven by the density difference perimental configurations.
between the two fluids. Similar density-driven flows from the wide to The outline of the paper is as follows. The numerical and experi-
the narrow side of an eccentric annulus were also observed by Jakobsen mental methods along with the annulus geometry and fluid properties
et al. (1991). Experimental validation of the narrow annulus (Hele- are discussed in section 2. In eccentric annuli, yield stress fluids cease to
Shaw) displacement model developed by Bittleston et al. (2002) was flow on the narrow side of the annulus once the extent of the unsheared
partly the motivation for the work by Malekmohammadi et al. (2010). plug zone spans the gap between the inner and outer pipes of the

999
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

annulus. We address the possibility for having static fluid in the ec-
centric annulus tests in section 3. In section 4 we compare the experi-
mental measurements with corresponding numerical simulations. Ad-
ditional simulation results pertaining to fluid displacement in the hole
enlargement is discussed in section 5. The main dimensionless numbers
governing the flow are compared to a realistic field case in section 5.1.
We conclude in section 6.

2. Numerical and experimental methods

In the following subsections we define the annulus geometry,


properties of the two fluids, the experimental conditions as well as re-
levant details pertaining to experimental and numerical methods.

2.1. Annulus geometry

The annulus geometry under consideration is 10 m long in the axial


direction and composed of an inlet, a washout and an outlet section.
The inlet and the outlet sections are considered to be the regular part of
the annulus with nominal inner and outer diameter, whereas the
washout zone represents the irregularity. The physical dimensions of
the annulus geometry are tabulated in Table 1.
The outer radius of the annulus increases by 69% in the washout
zone compared to the adjacent regular inlet and outlet sections. In the
numerical model and in the experiments, the regular sections are con-
nected to the washout zone by linear tapers spanning axial lengths of
48 mm. See Fig. 3 for an impression of the system.
In this paper, we consider concentric and one eccentric configura-
tion of the inner pipe. Perfect centralization with azimuthally uniform
gap size between the inner and outer pipes correspond to the concentric
configuration. In a wellbore, the inner tubing will generally always be
configured eccentric relative to the outer casing or formation. We de-
fine the eccentricity as e = δ /(Ro − Ri ) , where δ denotes the offset be-
tween the axis of the two cylinders, and Ri and Ro denote the radii of the
inner and outer cylinders, respectively. The eccentric annulus discussed
in this paper corresponds to an eccentricity of e = 0.42 in the regular
parts, with the inner tubing shifted toward the low side of the outer
pipe. This eccentricity corresponds to an offset between the cylinder
centre lines of 8 mm. In the washout section, where the outer pipe ra-
dius is significantly larger than in the regular parts, the eccentricity is
only 0.105. The increase in hole size inside the washout is considered to Fig. 1. : Steady state Herschel-Bulkley flow curves for displaced and displacing
be realistic based on previous analyses of caliper logs from a 12 1/4” fluids, and fully developed single phase laminar friction pressure gradient for
hole section (Skadsem et al., 2017). each of the two fluids in the concentric annulus bounded by an inner pipe with
radius 0.0635 m and an outer pipe with radius 0.08255 m.
2.2. Fluid properties
fluids were viscosified with the synthetic clay laponite and with xan-
The mass density and the rheological properties of the two fluids than gum, respectively. These additives result in steady state behavior
were designed to achieve the desired density and viscosity hierarchy characterized by shear thinning and a non-zero yield stress. Fann 35
between the fluids, with the displacing fluid being the denser and ef- viscometer measurements of the two fluids are shown as the points in
fectively most viscous of the two. The mass density of the displaced are shown as Fig. 1a.
fluid was 1000 kg/m3 while the displacing fluid was slightly denser at We parametrize the steady state rheology with the constitutive
1100 kg/m3. To achieve desired rheology, the displaced and displacing Herschel-Bulkley model (Herschel and Bulkley, 1926) which describes a
medium that acts as a rigid body in regions where the shear stress τ is
Table 1 lower than the yield stress τy :
Physical dimensions of the annular geometry.
γ˙ = 0, where τ ≤ τy, (1a)
Dimension Value (mm)
τy
Outer radius, regular annulus 82.55 μ (γ˙ ) = + Kγ˙ n − 1, elsewhere
Outer radius, washout 139.7 γ˙ (1b)
Annulus inner radius 63.5
Vertical offset between cylinder axes 8.0
Here, the rate of strain tensor is denoted γ̇ and defined as
γ˙ = (∇v + (∇v)†)/2 with v the fluid velocity vector. The shear rate
Inlet axial length 6122 dependent viscosity is denoted μ and the Herschel-Bulkley model
Washout taper axial length 48 parameters that are fitted to viscometer measurements are the yield
Washout axial length 1660
stress τy , the consistency index K and the shear thinning index n < 1.
Washout taper axial length 48
Outlet axial length 2122 The shear rate is defined as the second invariant of the rate of strain
tensor:

1000
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Table 2 Table 3
Summary of fluid properties. Test matrix for the displacement experiments and simulations. Inclination is
measured from the vertical direction.
Fluid Mass density Rheology parametrization
Label Flow rate Eccentricity Inclination
3
Displaced 1000 kg/m τ = (6.42 + 0.55γ˙ 0.5) Pa
Displacing 1100 kg/m τ = (1.53 + 1.58γ˙ 0.5) Pa A2 4.36⋅10−3 m3/s 0.42 90∘
B2 4.36⋅10−3 m3/s 0.42 60∘
G2 4.36⋅10−3 m3/s 0 90∘
F2 4.36⋅10−3 3
m /s 0 60∘
γ˙ = 2(γ˙ : γ˙ ) = 2 ∑ γ˙ij γ˙ ji .
i, j (2)
2.3. Experiment and simulation matrix
Finally, the shear stress tensor is given by the product of the rate of
strain tensor and the viscosity function in Eq. (1b) (Bird et al., 1987):
We focus on the effect of eccentricity and inclination on the dis-
τ = 2μ (γ˙ ) γ˙ . (3) placement in the annulus test cell defined in Table 1, and define four
different parameter combinations listed in Table 3. The axial flow rate
The two solid lines in Fig. 1a correspond to the parametrization of is held constant at 4.36⋅10−3 m3/s (or equivalently 261.7 l/min) in these
the two fluids. A summary of fluid properties is provided in Table 2. tests, corresponding to a bulk velocity of 0.5 m/s in the regular part of
Both fluids are seen to be highly shear thinning, n = 0.5, and the the geometry. Referring to the friction pressure estimates in Fig. 1b, the
displaced fluid is characterized by a significant yield stress greater than pressure gradient is larger in the displacing fluid than the displaced
that of the displacing fluid. The higher consistency index of the dis- fluid at this flow rate. Tests G2 and F2 are repeats of A2 and B2, but
placing fluid makes this fluid the effectively more viscous of the two at with the inner pipe centered inside the outer pipe, giving a concentric
shear rates above 22.4 s−1. annulus. As mentioned above, the inner pipe is offset in the downward
In Fig. 1b we present relations between the fully developed laminar direction in tests A2 and B2 and consequently the narrow part of the
friction pressure gradient and constant axial flow rate for a concentric annulus is on the low side. For the slanted tests B2 and F2, the annulus
annulus bounded by an inner pipe with radius 0.0635 m and an outer is configured so that the outlet is elevated 5 m above the inlet, giving an
pipe with radius 0.08255 m. The pressure gradient and flow rate are inclination angle of 60∘ between the vertical and the axis of the annulus.
linked using the concentric annulus solution presented by Hanks In the numerical simulations, the displaced and displacing fluids
(1979). The friction pressure gradient is greater in the displacing fluid have mass density and rheology as defined in Table 2. Small variations
at flow rates greater than approximately 40 l/min. in rheology between consecutive experiments were observed, but this is
The fully developed axial velocity profile of the two fluids in a considered to be a minor effect that we neglect for the purpose of
concentric annulus is calculated based on the solution in Ref (Hanks, comparison with numerical results.
1979). and shown in Fig. 2 for an annulus bounded by inner pipe of
radius 0.0635 m and an outer pipe with radius 0.08255 m. For both 2.4. Experimental method
fluids, the velocity profile corresponds to a bulk axial velocity of 0.5 m/
s. The main difference between the two fluids is the larger central plug Experimental results presented in this paper have been obtained
zone of the displaced fluid compared to the displacing fluid. This results from displacement experiments in the flow loop described by Ytrehus
in a slightly higher wall shear rate in the displaced fluid, and a greater et al. (2017). The annulus test cell was constructed for the purpose of
plug velocity in the displacing fluid. investigating narrow annulus displacement flows with dimensions as
In sec. 3 we discuss the velocity profile of the displaced fluid in an listed in Table 1. The flow loop is instrumented by a flow meter, dif-
eccentric annulus, with particular focus on whether the plug region on ferential pressure transmitters and conductivity probes mounted azi-
the narrow side of the annulus is static or mobile at the flow rate used in muthally around the annulus, with the possibility to monitor four dif-
the experiments. ferent measurement stations A to D, as shown in Fig. 3. The
conductivity probes penetrate approximately 5 mm into the annulus
from the inner wall of the outer pipe. Sodium chloride was added to the
displacing fluid to achieve an electrical conductivity contrast to the
fluid to be displaced.
Experiments were carried out by first circulating the fluid to be
displaced through the flow loop and circulating the displacing fluid
through a bypass loop. The displacing fluid was then introduced from
the bypass loop into the annular test cell by operating a valve. The
displacement was then tracked by logging the conductivity probe sig-
nals, which indicate the presence of displaced or displacing fluids at

Fig. 2. Fully developed laminar axial velocity profiles for the displaced and the Fig. 3. Conductivity probes numbered 1 through 8 mounted around the an-
displacing fluid at bulk velocity of 0.5 m/s in the concentric annulus bounded nulus at the four measurement stations A through D. The axial distance from the
by an inner pipe with radius 0.0635 m and an outer pipe with radius 0.08255 m. inlet to station A and between consecutive stations is 2 m.

1001
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

different positions close to the outer wall of the annulus. We refer to the between accuracy and computational time was reached at 12 by 48 cells
paper (Ytrehus et al., 2017) for additional details of the experimental in the cross-section of the regular parts, as seen in Figs. 4, and 36 by
method. 48 cells in the cross-section of the model washout. A total of 224 cells
were used in the axial direction.
2.5. Numerical method
3. Mobilization on the narrow side of the regular annulus
The open source computational fluid dynamics software tool
OpenFOAM version 2.3.0 (OpenFOAM) has been used for the numerical The fully developed, non-inertial flow of a yield stress fluid in a
simulations of the displacement flows presented in this paper. The regular eccentric annulus is governed by the dimensionless Bingham
fluids are considered incompressible under the relevant pressures in the number, the shear thinning index, the annulus aspect ratio and the
annulus flow loop and we assume that the flow is laminar and iso- eccentricity. Depending on the specific combination of these para-
thermal. In this case, the governing equations are the continuity meters, the fluid may be static or mobilized and flowing on the narrow
equation side of the annulus. At low Bingham numbers, the applied pressure
∇⋅u = 0 (4) gradient is sufficient to yield the fluid throughout the annulus. At
higher Bingham number, the plug in the narrow sector spans the entire
and the momentum conservation equation
gap between the inner and outer walls of the annulus, resulting in static
∂ρ u fluid. In the following we investigate whether the fully developed,
+ ∇⋅(ρ uu) = −∇p + ∇⋅τ + ρ f.
∂t (5) single phase flow of the experiment fluids exhibits mobilized or static
fluid on the narrow side of the annulus at the axial flow rate used in the
The fluid mass density is denoted by ρ, u is the fluid velocity vector,
experiments. We limit the following discussion to the displaced fluid
p is the pressure, τ is the deviatoric stress tensor and f denotes any body
which has the largest yield stress of the two fluids.
forces acting on the fluids. Gravity is the only body force included in
In a narrow eccentric annulus with aspect ratio κ = Ri / Ro ap-
our simulations.
proaching unity, we may to leading order neglect the effect of curvature
In regions of the flow where the shear stress equals the yield stress
and approximate the minimum pressure gradient required to yield fluid
of the fluid, that is at the boundary of unyielded regions of the flow, the
on the narrow side by 2τy / h , where h is the radial gap between the inner
effective viscosity (1b) diverges to infinity. In this work, we have used
pipe and the outer pipe measured on the narrow side (Couturier et al.,
the regularized, built-in biviscosity model that comes with OpenFOAM.
1990). For the regular part of the eccentric annulus with dimensions as
In principle, this model limits the maximum attainable viscosity to a
in Table 1, h ≈ 0.011 m. We estimate that a minimum pressure gradient
large, numerical value, thereby allowing the fluid to flow everywhere,
of 1162 Pa/m is required to ensure mobilization of the displaced fluid
also in regions where τ ≤ τy . As we focus on transient flows where in-
on the narrow side of the annulus. The experiments and simulations
ertia and buoyancy should dominate over yield stress effects in most of
reported in this paper were carried out at an axial flow rate of
the geometric domain, we consider this to be a reasonable and prag-
4.36⋅10−3 m3/s, corresponding to an axial bulk velocity of 0.5 m/s.
matic choice also for the annulus geometry with an irregular, washout
Using the concentric annulus result obtained from the solution derived
section. The chosen limit for the viscosity was never reached in the
by Hanks (1979) and adjusting for eccentricity using the correlation
simulations. Additional discussion concerning the balance of different
developed by Haciislamoglu and Langlinais (1990), we predict a pres-
forces in these flows are provided in section 5.1, with dimensionless
sure gradient of 1384 Pa/m in the eccentric annulus. This exceeds the
numbers listed in Tables 4 and 5.
criterion of 1162 Pa/m by approximately 19%, and suggests mobiliza-
The governing equations (4) and (5) are discretized and solved
tion on the narrow side.
using the finite volume method in OpenFOAM (OpenFOAM). In order to
To further explore the flow regimes in the annulus for the displaced
simulate multiphase flow with different fluids, the interFoam solver
fluid, we construct a phase diagram for the flow according to the pro-
with the volume of fluid approach is used, where the volume fraction
cedure described by Szabo and Hassager (Bittleston and Hassager,
and interface of the phases are captured, but where the phases are not
1992). Although their analysis focused on Bingham fluids, it is
allowed to mix. In this solver, the two phases are treated as one in the
straightforward to generalize the analysis to Herschel-Bulkley fluids,
sense that weighted averages are used in the continuity and momentum
using the results in Ref. (Hanks, 1979). A phase diagram for the dis-
conservation equations (4) and (5) for the velocity, density and other
placed fluid in a κ = 0.77 annulus at varying eccentricities and Bingham
mixture properties (Rusche, 2003; Sawko, 2012; Deshpande et al.,
numbers is provided in Fig. 5. The Bingham number definition is here
2012; Friedmann et al., 2017; Schaer et al., 2018). The volume fraction
α works as an indicator function for the phases, and requires an extra
transport equation
∂α
+ ∇⋅(uα ) + ∇⋅(ur α (1 − α )) = 0
∂t (7)

to be solved. Here, ur is an artificial velocity field that is added to


compress the interface (Rusche, 2003; Friedmann et al., 2017).
Immiscibility of the phases is a simplification with respect to the
experiments, where the displaced and displacing fluids indeed were
miscible. However, for the current flow rate the diffusive time scale is
assumed to be large compared to the time scale of interest, and effects
of molecular diffusion on the displacement process are consequently
neglected. The interfacial tension is zero in the simulations.
To close the set of governing equations, we specify for the velocity Fig. 4. Mesh resolution is shown in a detailed view (a) and the full cross-section
Dirichlet conditions at solid boundaries and at the inlet (no slip and (b) of the regular parts. For comparison with the experimental conductivity
uniform, bulk inlet velocity, respectively) and Neumann conditions at measurements, the fluid volume fraction was averaged over spherical volumes
the outlet. For the pressure, Neumann inlet and Dirichlet outlet con- with radii 3 mm and each centered 4 mm from the outer wall, as indicated in
ditions are used. the figure. Due to symmetry, only positions on the left side when looking in the
Following a mesh dependency study, a reasonable compromise direction of the flow were used.

1002
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

achieved by extracting from the simulations the fluid volume con-


centrations at the location of the probe tip, however, since there is an
effective region of sensitivity for the probes (Kong et al., 2016), a
sphere with radius 3 mm is used. The sphere centers are positioned
4 mm into the annulus gap from the outer pipe wall as shown in Fig. 4,
and the minimum, maximum and average values of the displaced fluid
volume concentration inside the spheres are extracted.
As an alternative approach the distribution of displaced (red) and
displacing (blue) fluids are visualized in the cross-sections at stations C
and D at different times after the test started and can be compared to
the conductivity data time series. While the probe responses represent
point measurements of the displacement process, the cross-sectional
view provides information concerning fluid distributions throughout
the cross-section.
We begin by considering the two concentric annulus configurations
in sections 4.1 and 4.2, and turn to the eccentric configurations in
sections 4.3 and 4.4.
Fig. 5. Phase diagram for the displaced fluid. The notation is the same as in
Szabo and Hassager (Szabo and Hassager, 1992): (I) indicates no flow at all, (II) 4.1. Test G2: concentric and horizontal
is one moving plug zone in the wide sector and stationary fluid in the narrow
sector, (III) is two moving plug zones separated by a pseudoplug in the azi-
Station C is 6 m downstream of the inlet to the annulus geometry
muthal direction, and (IV) is one continuous moving plug zone. The blue dot
and with a bulk velocity of 0.5 m/s, a piston-like displacement front
indicates location in the phase diagram of the displaced fluid at estimated
would arrive at this position after 12 seconds from the start of the test.
pressure gradient 1384 Pa/m. (For interpretation of the references to color in
this figure legend, the reader is referred to the Web version of this article.) The experimental and simulated measurements from this station for test
G2 is shown in the graphs in Fig. 6, while cross-sectional fluid dis-
tributions from different times in the numerical simulation are shown
taken to be Bn = τy /(Ro (−dp / dz )) , as per Szabo and Hassager (1992).
below.
We denote the different flow regimes using the same notation as Szabo
As expected for the concentric annulus configuration, the denser
and Hassager (Bittleston and Hassager, 1992), i.e. (I) denotes no flow in
displacing fluid appears at the lowest probe at position 5 first, followed
the annulus, (II) denotes stationary fluid in the narrow sector and flow
by probes at position 6 and 3, then at position 7. The arrival time at
in the wide sector, (III) denotes plugs in the wide and in the narrow
probe position 5 is about 6–7 seconds in the experiment, i.e. much
sector moving at different axial velocities and (IV) indicates a con-
sooner than a piston-like displacement at this position. In all these cases
tinuous plug around the entire annulus and moving at a constant axial
the measured conductivity increases quickly toward unity. The probe
velocity.
signals acquired at positions 6 and 3 are very similar, indicative of
We estimate the Bingham number of the displaced fluid to be 0.056
symmetry about the vertical direction, as expected. Displacing fluid
for the experiments and the simulations based on the yield stress of
arrives later at the top positions 1 and 8, after about 12–13 seconds at
6.42 Pa, an estimated pressure gradient of 1384 Pa/m and Ro of
position 8, followed gradually by the probe at position 1. The two latter
0.08255 m. The location in the phase diagram is indicated by a blue
probe measurements fluctuate more than the measurements from the
dot, suggesting mobilized fluid also in the narrow part of the annulus as
lower probes, indicating more mixing and a more gradual displacement
per the analysis above.
from the upper part of the annulus or that the probe is close to a per-
sistent and wavy horizontal interface between the two fluids.
4. Experimental results and simulation comparison The numerical simulation of this test suggests arrival of displacing
fluid between approximately 10 and 12 seconds. The arrival time at
We now turn to the experimental results from the tests defined in probe position 5 in the simulation is particularly delayed compared to
Table 3 and comparison with the numerical simulations of the same the experimental measurement of arrival after about 6–7 seconds. The
tests. This is shown in Figs. 6–14. In the experiments, the fluid interface experimental measurements from probes at positions 3, 6 and 7 are also
was detected using conductivity probes spaced by 45° around the an- slightly ahead of the numerical simulations, while the top conductivity
nulus at the four measurement stations A to D as illustrated in Fig. 3. In probes at positions 1 and 8 agree quite well with the simulation which
this section, we focus on the conductivity probes mounted at station C indicate gradual arrival of displacing fluid. In short, the interface be-
and station D, i.e. just in front of and immediately after the washout tween the displaced and displacing fluids appear shorter and more
section. A total of 6 conductivity probes were installed at each of these piston-like in the numerical simulations compared to the experimen-
stations. The azimuthal probe positions relative to the vertical, upward tally observed interface. We remark that Renteria et al. (2018) also
direction are as follows: observed in their simulations that downward slumping of dense fluid
was significantly less than in the experiments. This seems to be in
• Probes at positions 1 and 8: ± 22. 5 ∘ qualitative agreement with our results presented in Fig. 6, where dis-
• Probe at position 7: − 67. 5 ∘
placing fluid arrives at all azimuthal probe positions within a narrow
• Probes at positions 3 and 6: ± 112. 5 ∘ time interval of a few seconds.
• Probe at position 5: 157. 5 ∘
Next, turning to measurements from station D after the washout in
Fig. 7, arrival of displacing fluid appears first at conductivity probe
Throughout, the raw data acquired from the conductivity probes position 5, after about 18 seconds in the experiment, followed shortly
have been normalized to the interval between 0 and 1, with 0 indicating by positions 3 and 6, then position 7 after about 20–21 seconds. Once
displaced fluid and 1 indicating displacing fluid at the location of the again these probe signals increase nearly monotonically indicating
probe at the given time measured from the start of the test. stable and efficient displacement of the original fluid. Arrival of dis-
Comparison with the numerical simulations is performed by esti- placing fluid at the top positions 1 and 8 is now significantly delayed
mating the conductivity probe signal from the simulations, as done very and signs of displacing fluid can be seen after about 40 seconds at probe
recently by Tardy (2018) and also by Renteria et al. (2018). This can be position 1. As shown for the simulation results, arrival of displacing

1003
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 6. Comparison between experimental measurements and simulations for Fig. 7. Comparison between experimental measurements and simulations for
test G2 (concentric and horizontal) at station C, before the washout. Solid lines test G2 (concentric and horizontal) at station D, after the washout. Labels as in
represent measurements from conductivity probes at positions indicated by Fig. 6.
azimuthal angle (and “left” and “right” for symmetric positions). The data have
been normalized and are interpreted as phase concentration, with 0 and 1 in-
between 10 and 12 seconds. The overall agreement between experiment
dicating displaced and displacing fluids, respectively. The dashed lines re-
present corresponding simulated results, estimated from the average volume
and simulation at this position is considered better than the corre-
fraction in a small sphere around the probe tip, and the colored region indicates sponding horizontal test in section 4.1.
the spread of the data. Snapshots of the cross-sectional distribution of phase The experimental results acquired at measurement station D, after
concentration are shown below for selected times. the washout, show a pronounced effect of inclination on arrival times of
displacing fluid at all conductivity probe locations. Once again, the
displacing fluid arrives first at the bottom position 5, at about 27–28
fluid at the bottom positions occur after about 25 seconds, and after
seconds. Arrival at the other probe locations follows within a few sec-
about 27 seconds at position 7. This is a significant delay compared to
onds, even at the top probe location 1 and 8. This is in clear contrast to
the experimental results, and consistent with the trend observed at
the results for the horizontal annulus shown in Fig. 7, where first arrival
station C, namely earlier arrival in the experiments than in the simu-
was some 10 seconds earlier at the lower probes, and significantly later
lations. Again, this is consistent with the observations of Renteria et al.
at the top probe locations 1 and 8. The observation that increased in-
(2018).
clination results in a shorter, more piston-like displacement compared
to the horizontal annulus is in accordance with expectations.
4.2. Test F2: concentric and slanted The numerical results in Fig. 9 show very good quantitative agree-
ment between simulation and experiment at this measurement station.
Turning to test F2, where the annulus is inclined by 60∘ from the
vertical, the vertical slumping of displacing fluid to the low side of the 4.3. Test A2: eccentric and horizontal
annulus should intuitively be smaller than the horizontal configuration.
This leads to a shorter axial extent of the interface between the two As we will see in this section, inner pipe eccentricity has an im-
fluids compared to the horizontal configuration considered in section portant effect on the fluid distributions in the now eccentric annulus.
4.1. Indeed, the experimental results acquired at station C and shown in Starting with the eccentric and horizontal test A2, the measurements
Fig. 8 are consistent with this expectation. Displacing fluid arrives first acquired at station C, shown in Fig. 10, show that arrival of the dis-
at probe position 5, as expected, followed shortly after by the other five placing fluid is now first at the upper probe locations, namely at posi-
probes. Arrival of displacing fluid occurs after about 7 seconds at po- tion 7, followed immediately by positions 1 and 8 at about 8–10 sec-
sition 5, about 9 seconds at positions 3, 6 and 7, and about 10–11 onds. Arrival is delayed to about 15 and 20 seconds at probe positions 6
seconds for positions 1 and 8. and 3, respectively.
The results of the numerical simulation are similar to the results for These observations agree well with the numerical simulation of test
the horizontal configuration in Fig. 6, with displacing fluid arriving A2, shown in Fig. 10. The denser displacing fluid flows fast in the wide,

1004
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 8. Comparison between experimental measurements and simulations for Fig. 9. Comparison between experimental measurements and simulations for
test F2 (concentric and inclined 60∘ from the vertical) at station C, before the test F2 (concentric and inclined 60∘ from the vertical) at station D, after the
washout. Labels as in Fig. 6, but the results of an extra repetition of the ex- washout. Labels as in Fig. 6, but the results of an extra repetition of the ex-
periment are added as thin lines. periment are added as thin lines.

top part of the annulus while buoyancy due to the mass density dif- location 1 and 8.
ference between the fluids leads to azimuthal flow of the displacing The simulation suggests arrival of displacing fluid at measurement
fluid from the top toward the low side. The net effect is arrival of dis- station D after about 17 seconds, and then primarily at probe locations
placing fluid nearly simultaneously at positions 1, 8 and 7 in both ex- 2 and 7. Some five to six seconds later do the simulation indicate dis-
periment and simulation. The simulation also indicates slow arrival of placing fluid at the other probe locations 3, 6 and 5. This is delayed
displacing fluid at positions 3 and 6, i.e. between 10 and 20 seconds, compared to the experimental result. As clearly indicated in the simu-
which again agrees well with the experimental results acquired at these lation results, there is a lingering film of displaced fluid close to the
two locations. We note that the conductivity probe at location 5 in upper part of the annulus, partly covering probe locations 1 and 8 for a
station C probably malfunctioned during this test. very long duration. This film is due to slow displacement of fluid within
Moving on to measurement station D, after the washout, we observe the washout, with the lighter displaced fluid exiting the washout in the
in Fig. 11 that arrival of displacing fluid occurs nearly simultaneously at upper part of the annulus. The probe at location 1 seems to detect a
all probe locations except the top probes at location 1 and 8. Arrival at small volume of displacing fluid starting at about 35 seconds resulting
the other positions occur after about 15 seconds and first at location 3. in a fluctuating sensor value for an extended period of time. This is
The arrival time at location 3 at station D is nearly the same as for this qualitatively consistent with the numerical results, however, small
position in station C, as seen in Fig. 10. This suggests that displacing variations in the position of the interface between the fluids may sig-
fluid that enters the washout in the wide, upper part of the annulus nificantly affect the probe signal, as shown in the visualizations of the
exits the washout at a lower azimuthal position. This is also consistent fluid distributions in the bottom panels in Fig. 11.
with the very late arrival of displacing fluid at conductivity probe

1005
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 10. Comparison between experimental measurements and simulations for Fig. 11. Comparison between experimental measurements and simulations for
test A2 (eccentric and horizontal) at station C, before the washout. Labels as in test A2 (eccentric and horizontal) at station D, after the washout. Labels as in
Fig. 6. Fig. 6.

4.4. Test B2: eccentric and slanted fluid occurs nearly simultaneously at probe locations 3, 5, 6 and 7 after
about 25 seconds. Particularly striking is the early arrival of displacing
Finally, the combination of eccentricity and inclination is studied in fluid at the top probe locations 1 and 8, where displacing fluid arrives
test B2. The measurements at station C, just in front of the washout, is after approximately 28 seconds. This suggests a more uniform fluid
shown in Fig. 12, and is qualitatively very similar to the measurements distribution over annulus cross-section compared to test A2, where the
acquired in from this station in the horizontal arrangement, Fig. 10. annulus was horizontal. In the B2 simulation, the film of non-displaced
Once again, probe location 7 sees the first arrival of displacing fluid, fluid still seems to be present, however, much less pronounced com-
followed shortly after by the probes at locations 1 and 8, while the pared to the horizontal test.
lower probes at position 3 and 6 detect displacing fluid somewhat later. It is apparent from these results that inclination improves fluid
The probe signals measured at position 3 and 6 are once again oscil- displacement from the top part of the washout and in the regular part
lating as function of time, indicating that the displacement here is not immediately after the washout. Agreement between experiment and
caused by a stable front but rather displacing fluid slumping down from simulation is very good also at this measurement station.
the wide, top part of the annulus in an unpredictable manner. The
numerical simulation shows results that are very similar compared to
4.5. Summary and discussion of the comparison
the eccentric, horizontal test discussed in section 4.3, and the agree-
ment between experiment and simulation is good with the exception of
To conclude this section on comparison between displacement ex-
probe 5, on the narrow side. The probe at location 5 is not detecting any
periments and the numerical simulations, we consider the simulations
injected, displacing fluid during the time interval shown in Fig. 12,
of the two eccentric tests, A2 and B2, to successfully capture the dy-
indicating either probe malfunction or the presence of static fluid on the
namic displacement process in the experiments, and there is a tendency
narrow side. Inspection of the entire time series acquired during the test
that the slanted tests are better replicated by the simulations than the
show that the probe indeed functions and detects displacing fluid at
horizontal are. The qualitative trends in these experiments are well
location 5 starting at about 150 seconds. This observation suggests
captured by the simulations, and even several arrival times of the dis-
mobilization issues in the experiments that are not captured by the
placing fluid are also matched well. The two horizontal tests, A2 and
numerical simulations.Whereas inclination did not alter the fluid dis-
G2, show larger differences between experiment and simulation, par-
placement in the regular part of the annulus leading up to station C, we
ticularly the concentric and horizontal test, where arrival times at
observe in Fig. 13 a significant effect of inclination on displacement at
several probe positions are considerably longer in the numerical si-
station D, after the washout. Compared to the horizontal annulus in
mulations compared to the experiments.
Fig. 11, arrival of displacing fluid at station D is delayed by approxi-
The observed differences can be due to a combination of simplifying
mately 10 seconds when the annulus is slanted. Arrival of displacing
modelling assumptions within the simulations, and uncertainties

1006
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 12. Comparison between experimental measurements and simulations for Fig. 13. Comparison between experimental measurements and simulations for
test B2 (eccentric and inclined 60∘ from the vertical) at station C, before the test B2 (eccentric and 60∘ from the vertical) at station D, after the washout.
washout. Labels as in Fig. 6, but the results of an extra repetition of the ex- Labels as in Fig. 6, but the results of an extra repetition of the experiment are
periment are added as thin lines. added as thin lines.

associated with the displacement experiments. The simulations do not the narrow side, we found that the steady state single phase flow of
account for fluid mixing, and treat the two fluids as inelastic, general- these fluids should result in fluid mobilization on the narrow side. This
ized non-Newtonian fluids with a time-independent rheology defined analysis does not account for inlet effects or e.g. thixotropic effects in
by the steady state flow curves in Fig. 1a. Further, simple Dirichlet the fluids. In Fig. 14, we show measurements acquired at station A, 2 m
boundary conditions are assumed at the inlet of the annulus geometry, downsteam the inlet position of the annulus for the eccentric, slanted
giving the same bulk velocity as in the experiments. These are ap- test B2.
proximations that, together with a finite spatial and temporal resolu- We observe significantly delayed arrival of displacing fluid on the
tion, introduce errors compared to the experiments. narrow side in this test compared to the other probe signals and com-
The positions and penetration depth of the conductivity probes into pared to the simulations. The arrival time at this station was even
the annulus are indicated in the figures in the preceding section. Minor longer for the eccentric, horizontal test A2. Although the relative vo-
variations in probe positioning or the actual fluid distribution in the lume in the narrow sector of the annulus is small, such effects result in
annulus can have significant effect on the interpretation of these mea- uncertainty and possible sources for the differences between experi-
surements. Especially in the start-up of the experiments, there will be a ment and simulation.
short transient period before the target flow rate is achieved, and this
introduces uncertainty related to the recorded arrival times.
5. Discussion of findings from simulations
Flow on the narrow side of the eccentric annulus is particularly
sensitive to geometry, fluid rheology and the applied pressure gradient
Based on the test results analyzed above, we observe that inclination
in the flow. Going back to the discussion in section 3 regarding flow on
has an important effect on the fluid displacement front at measurement

1007
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 15. Simulated ratio of displacing fluid volume in the washout to the total
volume of the washout as a function of injected fluid, specified in units of the
total annulus volume. At the current flow rate, one unit corresponds to a time of
approximately 36 s. The ratio is equivalent to the displacement efficiency in the
washout segment. Ideal, piston-like displacement is indicated by the black
dotted line.

stage not clear what is behind the difference between experiment and
simulation. Both numerical models are based on simplified rheological
descriptions of the fluids. A second possible source of uncertainty is the
existence of transient and/or inlet effects in the experiments that are
not captured in either of the simulations.
Fluid displacement from the washout segment of the annulus can be
assessed in terms of a displacement efficiency, which is the overall
volume fraction of the displacing fluid in the annulus (Tehrani et al.,
1992, 1993). In the following, we focus on the washout segment only,
so we define the displacement efficiency as the volume of displacing
fluid inside the washout relative to the total washout volume. Complete
displacement in the washout then corresponds to a displacement effi-
ciency of unity.
In Fig. 15 we plot the washout displacement efficiency as function of
injected volume of displacing fluid. The injected volume is measured
relative to the total annulus volume, counting both the regular sections
and the washout section. With the chosen flow rate, one unit injected
Fig. 14. Comparison between experimental measurements and simulations for volume corresponds to approximately 36 seconds. The washout dis-
test B2 (eccentric and inclined 60∘ from the vertical) at station A, 2 m down- placement efficiencies for the four tests considered above are plotted
stream the inlet. Labels as in Fig. 6, but the results of an extra repetition of the
along with a black dotted line which represents perfect, piston-like
experiment are added as thin lines. Fluid displacement on the narrow side is
displacement. The perfect washout displacement efficiency reaches
significantly delayed in the experiments compared to the simulations.
unity slightly before one full annulus volume has been injected due to
the regular annulus volume between the washout and the outlet of the
station D, after the washout. This is evident when comparing the ex- geometry. When comparing the displacement efficiency for the two
perimental results for the concentric annulus in Figs. 7 and 9, and for eccentric tests, A2 and B2, and the two concentric tests, G2 and F2, it is
the eccentric annulus in Figs. 11 and 13. Inclination delays the first evident that slanting improves the displacement efficiency of the
arrival of displacing fluid at station D for both the concentric and the washout. This is consistent with the experimental and simulation results
eccentric annulus compared to the horizontal configuration. On the discussed in section 4, suggesting that less inclination with respect to
other hand, inclination promotes a more even distribution of displacing the vertical direction leads to more even fluid distribution around the
fluid around the annulus at station D after the washout. This is seen annulus in the washout, and better overall displacement of the washout
from the conductivity traces in Figs. 9 and 13, where all the probes compared to the horizontal configurations.
around the circumference of the annulus begin to detect displacing fluid The washout displacement efficiency for the eccentric and hor-
within a short time interval. In the two horizontal tests, Figs. 7 and 11, izontal test A2 is particularly poor compared to the other three tests in
the probes mounted in the top positions 1 and 8 detect displacing fluid Fig. 15. We may investigate this further by extracting information on
significantly later than the probes mounted along the lower part of the the fluid distributions along the annulus, as in Fig. 16.
annulus. For each of the four simulations, the volume fraction of residual,
The simulations failed to reproduce the downward slumping and non-displaced fluid after injection of one annulus volume of displacing
rapid displacement on the low side in the concentric and horizontal test fluid is averaged over the cross-section and plotted as function of axial
F2. Our results agree, at least qualitatively, with the very recent si- position. Measurement stations C and D are located at axial position 6 m
mulation results of Renteria et al. (2018) in this respect. It is at this and 8 m, respectively, and the washout section is indicated by the

1008
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 16. Cross-sectional average of the volume fraction of residual, non-dis- Fig. 18. Volume fraction of residual, non-displaced fluid averaged across the
placed fluid after injecting one annulus volume displacing fluid, corresponding upper (wide) annulus gap and across the lower (narrow) annulus gap after
to approximately 36 s after the start of the test. injecting one annulus volume displacing fluid, corresponding to approximately
71 s after the start of the test.

colored region. Starting with the two concentric tests G2 and F2, we
observe nearly no residual fluid in front of the washout, but local ac- Once again, we evaluate the fraction residual fluid after one annulus
cumulation of residual fluid within the washout, close to the entrance volume displacing fluid has been injected.
region. The figure shows more residual, non-displaced fluid in the As expected, the residual non-displaced fluid in front of the washout
concentric horizontal (G2) compared to the concentric slanted (F2) test is located on the narrow side of the annulus for the two eccentric tests
toward the rear exit region of the washout. This is consistent with the A2 and B2. Less residual fluid remains in the narrow side for the slanted
displacement efficiency in Fig. 15. geometry in B2. Moving into the washout, we observe a large accu-
Turning to the two eccentric tests A2 and B2, we observe more re- mulation of residual fluid on the low side in the first part of the washout
sidual fluid in front of the washout, most likely located on the low, for the eccentric tests A2 and B2, but also significant residual fluid at
narrow side of the regular annulus. More striking is the presence of a the top, toward the upper wall of the washout. Further into the
large accumulation of residual, non-displaced fluid in the front half of washout, inclination has aided fluid displacement on the low side of the
the washout in test A2. There is a similar accumulation in the front of annulus in test B2. Comparing the two concentric tests G2 and F2, we
the washout also in the eccentric, slanted test B2, but the volume observe a similar accumulation of residual fluid both at the top and the
fraction is in this case significantly smaller than test A2. This indicates bottom in the first part of the washout, and a sign of the displacement of
that the poor washout displacement efficiency in test A2 is due to the fluid from the upper part of the washout to be improved in the slanted
accumulated residual fluid in the front of the washout. one. In both cases we note a film of residual, non-displaced fluid in the
In Fig. 17 we average the residual, non-displaced fluid volume bottom part.
fraction separately across the annulus gaps at the top and bottom sides We then evaluate the volume fraction of residual, non-displaced
of the system. Effectively, we evaluate the fraction of residual fluid in fluid in the upper and lower sides of the annulus after injection of two
the vertical–axial plane above and below the inner pipe separately. annulus volumes displacing fluid. The fluid distribution is now as
shown in Fig. 18. For the two concentric annulus tests G2 and F2, there
is still significant accumulation of residual fluid in upper and lower side
of the first part of the washout. Displacement of these volumes requires
injection of substantially more displacing fluid. The residual fluid to-
ward the top is slowly being displaced from the washout and leaves a
significant trail of residual fluid in the upper side of the annulus
downstream the washout.
For the eccentric, slanted test B2, Fig. 18 suggests the overall vo-
lume fraction residual fluid is similar to the two concentric tests G2 and
F2, but that the distribution of residual fluid is on the low side of the
washout in test B2. This impression is confirmed by Fig. 15, where the
washout displacement efficiency after injection of two annulus volumes
is nearly identical across the three tests. The eccentric, horizontal test
A2 shows significant residual non-displaced fluid especially at the low
side of the washout. This is also confirmed by the displacement effi-
ciency result for this test in Fig. 15. The local drop seen in the fraction
non-displaced fluid on the low side about 0.5 m into the washout is due
to dense, displacing fluid slumping from the wide, upper part of the
annulus and down toward the bottom of the washout. This is also re-
flected in the higher fraction of the lighter, residual fluid on the upper,
Fig. 17. Volume fraction of residual, non-displaced fluid averaged across the wide side toward rear end of the washout in test A2. There is a sig-
upper (wide) annulus gap and across the lower (narrow) annulus gap after nificant fraction of residual fluid in the wide, upper side of the annulus
injecting one annulus volume displacing fluid, corresponding to approximately in both of the two concentric tests G2 and F2, and for the eccentric,
36 s after the start of the test.

1009
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 19. Evolution of the displacement process inside the enlarged section of Fig. 20. Evolution of the displacement process inside the enlarged section of
the concentric and horizontal test G2. The red fluid is displaced by the blue the concentric and slanted test F2. The red fluid is displaced by the blue fluid
fluid entering from the left side (color scale as in Figs. 6–14). Time is indicated entering from the left side (color scale as in Figs. 6–14). Time is indicated from
from the start of the test. (For interpretation of the references to color in this the start of the test. (For interpretation of the references to color in this figure
figure legend, the reader is referred to the Web version of this article.) legend, the reader is referred to the Web version of this article.)

horizontal test A2. This is not as prominent in the eccentric, slanted test and the large volume of residual fluid on the low side in the first part of
B2. the washout that can be seen at 70 s in Fig. 21 is essentially absent in
Finally, we examine the three-dimensional evolution of the dis- Fig. 22.
placement process as predicted by the simulations through series of
visualizations of the washout, as shown in Figs. 19–22. Each image is 5.1. Dimensional analysis
composed of one slice in the vertical–axial plane and four cross-sections
to show the azimuthal variation of the fluids, where red and blue colors To investigate how the experiments and simulations discussed in
indicate displaced and displacing fluid, respectively. A comparison of this paper scale relative to a field case displacement flow, we introduce
Figs. 19 and 20 for the two concentric cases G2 and F2 confirms the the concentric annulus gap width d * = 2(Ro − Ri ) as characteristic
above discussed impression that less inclination with respect to the length scale. The velocity scale U * is taken to be the axial bulk velocity,
vertical direction improves the displacement and seems to prevent a while μi* is the characteristic viscosity scales for the fluids with i = 1
channel of residual fluid along the upper wall of the washout. denoting the displaced fluid and i = 2 the displacing fluid. The geo-
In the eccentric and horizontal case A2 shown in Fig. 21 the dis- metric scaling factors are the radius ratio κ and the eccentricity e of the
placing fluid is predicted to bypass a significant amount of residual fluid annulus, while the Reynolds numbers Rei , Bingham numbers Bni and
on the low side in the first half of the washout. This is also apparent in densimetric Froude number Fr express different force ratios.
both Figs. 17 and 18 and is likely due to the inertia of the displacing Specifically, the Reynolds number quantifies the ratio of inertia to
fluid upon entering the washout. Even after injection of two annulus viscous forces, the Bingham number quantifies the ratio of yield stress
volumes, corresponding to approximately 71 seconds of injection, re- to viscous stresses, while the Froude number is the ratio of inertia to
sidual fluid remains along most of the bottom part of the washout, as buoyancy forces. Finally, the Atwood number At characterizes the
evident in both Figs. 18 and 21. Inertia combined with downward density difference between the fluids.
slumping results in a pocket of residual fluid in the wide, upper side to We note that the definition of the Bingham number in Table 4 in-
be located primarily in the second half of the washout, closest to the volves the yield stress, the characteristic axial velocity and viscosity and
exit. The displacement evolves in qualitatively the same way for the the annulus gap width. This definition differs from that used in con-
slanted eccentric case B2, as can be seen in Fig. 22. However, the re- nection to Fig. 5, where the number is based on the friction pressure
sidual fluid is removed considerably faster than in the horizontal case, gradient and annulus radius in addition to the yield stress. It is con-
venient to use the latter definition in the eccentric annulus flow

1010
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Fig. 21. Evolution of the displacement process inside the enlarged section of Fig. 22. Evolution of the displacement process inside the enlarged section of
the eccentric and horizontal test A2. The red fluid is displaced by the blue fluid the eccentric and slanted test B2. The red fluid is displaced by the blue fluid
entering from the left side (color scale as in Figs. 6–14). Time is indicated from entering from the left side (color scale as in Figs. 6–14). Time is indicated from
the start of the test. (For interpretation of the references to color in this figure the start of the test. (For interpretation of the references to color in this figure
legend, the reader is referred to the Web version of this article.) legend, the reader is referred to the Web version of this article.)

diagram in Fig. 5, since the analysis is based on force balance con- Table 4
siderations that naturally involve the axial friction pressure gradient. In Dimensionless numbers for scaling analysis.
Table 4 we instead base the Bingham number on the characteristic bulk Number Definition Description
velocity and viscosity to ensure consistency with the definitions of the
Reynolds and Froude numbers. Finally, we note that since the effective κ Ri / R o Pipe radius ratio
viscosity of generalized Newtonian fluids is a function of the shear rate, e δ Eccentricity
Ro − Ri
different definitions of the Reynolds number are used in the literature. At ρ2 − ρ1 Atwood number
ρ1 + ρ2
An alternative definition of the Reynolds number to that in Table 4 is to
Rei ρi U*d* Reynolds number in fluid i = 1,2
require that the product of the Reynolds number and the Darcy friction
μi*
factor is equal to 64 for fully developed, laminar duct flow also for non- Bni Bingham number in fluid i = 1,2
τy , i d *
Newtonian fluids. For Herschel-Bulkley fluids in particular, this results μi*U*
in a Reynolds number (Madlener et al., 2009) Fr U* Froude number
Atgd*
U *(2 − ni) d *ni
ρi
ReDarcy, i = , ni Shear thinning index for fluid i = 1,2
(τy, i/8)(d */ U *)ni + Ki ((3mi + 1)/(4mi ))ni 8ni − 1 (8)

with casing. While there are no common casing centralization requirements,


ni Ki (8U */ d *)ni good cementing practices suggest that the eccentricity should generally
mi = be no greater than 0.25 (Nelson and Guillot, 2006). In the proximity of
τy, i + Ki (8U */ d *)ni
wellbore irregularities, the actual casing centralization can be less than
for fluid i. A recent discussion of viscoplastic dimensionless numbers is planned for (Skadsem et al., 2017). The eccentricity of 0.42 is therefore
provided by Thompson and Soares (2016). considered severe but not unrealistic.
Concerning the geometric scales of the annulus, namely radius ratio The rheology measurements shown in Fig. 1a and the Herschel-
and eccentricity, we find κ ≈ 0.77 for the physical dimensions in Bulkley parametrization suggested in Table 2 indicate highly shear
Table 1. This is comparable to the aspect ratio of a 9 5/8″ casing in a 12 thinning fluids. For comparison to a field case, we consider this to be
1/4″ open hole section, or the aspect ratio of a 7″ tubing inside a 9 5/8”

1011
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

Table 5 compare arrival times of the displacing fluid with predictions from the
Comparison between simulations/experiments and field case. numerical simulations of the experiments.
Number Simulations/experiments Field case We observe significant effects of both eccentricity and inclination on
the annulus fluid displacement. Reduced inclination improves the dis-
κ 0.77 0.73 placement efficiency within the hole enlargement for both concentric
At 0.048 0.048
and eccentric annuli. Reduced inclination is beneficial for displacement
Re1 96 945
Re2 85 1147
also in the regular part of the annulus in front of the enlargement. The
Bn1 2.5 2.0 most important effect of eccentricity is the variation in axial flow ve-
Bn2 0.5 1.1 locity over a cross-section of the annulus. The variation in flow velocity
Fr 3.8 6.5 leads to an elongated interface between the two fluids in the regular
n1 = n2 0.5 0.5
part of the annulus, and also to poor displacement in the enlargement in
the horizontal annulus.
most relevant for the situation of a drilling fluid displaced by a spacer The experiments and simulations agree well for the two slanted tests
fluid, as cement slurries are often less shear thinning. In the following included in the study, with simulations replicating both qualitative
we consider the field case of a 1500 kg/m3 drilling fluid displaced by a trends and matching several of the measured arrival times from the
1650 kg/m3 spacer fluid, and assume that both of these fluids have a experiments. There is a larger quantitative difference between experi-
shear thinning index of 0.5, thereby preserving the mass density ratio ment and simulation for the concentric tests than the eccentric tests in
and n compared to the laboratory experiments and numerical simula- this study, particularly in the concentric horizontal test the simulated
tions. We limit the following discussion to the geometry of a 7″ tubing displacement is delayed compared to the experiment. Differences be-
in a 9 5/8” casing, and assume displacement occurs at a constant flow tween experiment and simulation can be due to a number of simplifying
rate of 1500 l/min. For the drilling fluid, we assume a rheological modelling assumptions and certain experimental uncertainties dis-
profile of τ = (4 + 0.5γ˙ 0.5) Pa and for the spacer fluid τ = (2 + 0.65γ˙ 0.5) cussed above.
Pa. This choice ensures a higher friction pressure gradient in the spacer The specific pair of displaced and displacing fluids has been fixed
fluid compared to the drilling fluid at the assumed flow rate in a con- between the four tests considered in this paper. The two fluids have
centric annulus of the assumed field case diameters. We define the been designed to provide the desired mass density and viscosity hier-
characteristic shear rate γ˙ * = 4U */ d * and use the effective viscosity archies in the tests, namely that the displacing fluid is both denser and
τy / γ˙ * + Kγ˙ *n − 1 as characteristic viscosity scale for each of the fluids. effectively more viscous than the displaced fluid. The model irregularity
Values of the dimensionless numbers for the simulations and ex- is not replicating an actual hole enlargement, but we consider the
periments compared to the field case are summarized in Table 5. idealized shape to still capture the main effects caused by a more
Compared to the field case, the largest difference is in the Reynolds complex enlarged section. This includes the presence of a large, local
numbers for the displaced and displacing fluids. To maintain the same annulus volume where the axial flow decelerates, the eccentricity is
Reynolds when downscaling, one should increase the flow rate, increase reduced and where there can be accumulation of residual fluid that is
the fluid mass density and/or decrease the fluid viscosity compared to only slowly displaced. The tests were carried out under laminar con-
the full scale model. The Reynolds numbers for the simulations and ditions and modest flow rates, resulting in Reynolds numbers lower
experiments are an order lower than the assumed field case, since the than comparable field cases. Work remains to explore displacement
rheological profiles are quite similar, while the flow rate and fluid flows where inertia is more dominant than here and to investigate more
density are both lower in simulations and experiments. Consequently, realistic irregular geometries.
inertia is less dominant compared to the field case. The Bingham
numbers are of the same order, suggesting comparable ratios of yield Acknowledgements
stresses to viscous stresses in the fluids. The field case Froude number is
larger than that of the simulations and experiments, suggesting that the The Research Council of Norway, ConocoPhillips, AkerBP, Statoil
magnitude of inertia compared to buoyancy is lower in the experiments and Wintershall are acknowledged for financing the work through
and simulations than in the example field case. Finally, using the PETROMAKS2 project number 244577/E30 and the research centre
Reynolds number based on the Darcy friction factor definition, i.e. DrillWell - Drilling and Well Centre for Improved Recovery, a research
Eq. (7), we find Reynolds numbers of approximately 143 and 110 for cooperation between IRIS, SINTEF, the Norwegian University of Science
the laboratory displaced and the displacing fluids, respectively, and and Technology and the University of Stavanger. The authors would
1351 and 1548 for the displaced and displacing fluids in the field case. like to thank Prof. Arild Saasen, University of Stavanger, for valuable
Both definitions result in similar ratio of the laboratory and field case discussions and contributions leading up to the experimental work.
Reynolds numbers. The experiments were run at moderate to low
Reynolds numbers mainly due to practical reasons. As new numerical References
models of non-Newtonian turbulent single phase and displacement
Alba, K., Taghavi, S.M., de Bruyn, J.R., Frigaard, I.A., 2013a. Inclomplete fluid-fluid
flows have been developed (Maleki and Frigaard, 2017; Gavrilov and
displacement of yield stress fluids. Part 2: highly inclined pipes. J. Non-Newtonian
Rudyak, 2016, 2017; Gori and Boghi, 2011, 2012), and as there is no Fluid Mech. 201, 80–93.
general consensus concerning laminar vs. turbulent displacement effi- Alba, K., Taghavi, S.M., Frigaard, I.A., 2013b. Miscible density-unstable displacement
ciencies, future displacement experiments at higher Reynolds number flows in an inclined tube. Phys. Fluids 25, 067101.
Bird, R.B., Armstrong, R.C., Hassager, O., 1987. Dynamics of Polymeric Liquids, vol. 1:
are needed. Fluid Mechanics. John Wiley & Sons.
Bittleston, S.H., Hassager, O., 1992. Flow of viscoplastic fluids in a rotating concentric
annulus. J. Non-Newtonian Fluid Mech. 42 (1 - 2), 19–36.
6. Conclusions Bittleston, S.H., Ferguson, J., Frigaard, I.A., 2002. Mud removal and cement placement
during primary cementing of an oil well – laminar non-Newtonian displacements in
an eccentric annular Hele-Shaw cell. J. Eng. Math. 43 (2), 229–253.
Experiments and 3D numerical simulations have been performed to Bogaerts, M., Azwar, C., Bellabarba, M., Dooply, M., Salehpour, A., Cementing, Wellbore,
study the fluid displacement in an annulus geometry with a hole en- 2015. An integral part of well integrity. In: Offshore Technology Conference,
Houston, Texas, USA, 4-7 May 2015, pp. 1–11 OTC-25800-MS.
largement. The study is focused on the effect of eccentricity and in-
Carrasco-Teja, M., Frigaard, I.A., 2010. Non-Newtonian fluid displacements in horizontal
clination on the displacement process for a specific pair of displaced narrow eccentric annuli: effects of slow motion of the inner cylinder. J. Fluid Mech.
and displacing fluids. The fluid interface has been measured with dis- 653, 137–173.
tributed conductivity probes in the displacement experiments, and we Carrasco-Teja, M., Frigaard, I.A., Seymour, B.R., Storey, S., 2008. Viscoplastic fluid

1012
H.J. Skadsem et al. Journal of Petroleum Science and Engineering 172 (2019) 998–1013

displacements in horizontal narrow eccentric annuli: stratification and travelling Pelipenko, S., Frigaard, I.A., 2004a. Two-dimensional computational simulation of ec-
waves solutions. J. Fluid Mech. 605, 293–327. centric annular cementing displacements. IMA J. Appl. Math. 69 (6), 557–583.
Clark, C.R., Carter, G.L., 1973. Mud displacement with cement slurries. J. Petrol. Technol. Pelipenko, S., Frigaard, I.A., 2004b. Mud removal and cement placement during primary
25 (7), 775–783 SPE 4090. cementing of an oil well - Part 2; steady-state displacements. J. Eng. Math. 48 (1),
Couturier, M., Guillot, D., Hendriks, H., Callet, F., 1990. Design rules and associated 1–26.
spacer properties for optimal mud removal in eccentric annuli. In: Annual Technical Pelipenko, S., Frigaard, I.A., 2004c. Visco-plastic fluid displacements in near-vertical
Meeting. Petroleum Society of Canada, Calgary, Alberta, pp. 1–8 pETSOC-90-112, narrow eccentric annuli: prediction of travelling-wave solutions and interfacial in-
SPE 21594. stability. J. Fluid Mech. 520, 343–377.
Deshpande, S.S., Anumolu, L., Trujillo, M.F., 2012. Evaluating the performance of the Plácido, J.C.R., Santos, H.M.R., Galeano, Y.D., 2002. Drillstring vibration and wellbore
two-phase flow solver interfoam. Comput. Sci. Discov. 5, 014016. instability. J. Energy Resour. Technol. 124 (4), 217–222.
Escudier, M.P., Oliveira, P.J., Pinho, F.T., 2002. Fully developed laminar flow of purely Renteria, A., Maleki, A., Frigaard, I.A., Lund, B., Taghipour, A., Ytrehus, J.D., 28 April
viscous non-Newtonian liquids through annuli, including the effects of eccentricity 2018. Effects of irregularity on displacement flows in primary cementing of highly
and inner-cylinder rotation. Int. J. Heat Fluid Flow 23, 52–73. deviated wells. J. Petrol. Sci. Eng. https://doi.org/10.1016/j.petrol.2018.08.045. in
Fredrickson, A., Bird, R.B., 1958. Non-Newtonian flow in annuli. Ind. Eng. Chem. 50 (3), press.
347–352. Roustaei, A., Frigaard, I.A., 2013. The occurrence of fouling layers in the flow of a yield
Friedmann, C.J., Mortensen, M., Nossen, J., 2017. Multiphase flow simulation in an an- stress fluid along a wavy-walled channel. J. Non-Newtonian Fluid Mech. 198,
nulus configuration. In: MekIT’17 - Ninth National Conference on Computational 109–124.
Mechanics, Trondheim 11-12 May, International Center for Numerical Methods in Roustaei, A., Frigaard, I.A., 2015. Residual drilling mud during conditioning of uneven
Engineering (CIMNE), Gran Capitán S/n, 08034 Barcelona, Spain, pp. 151–168. boreholes in primary cementing. Part 2: steady laminar inertial flows. J. Non-
Frigaard, I.A., Paso, K.G., de Souza Mendes, P.R., 2017. Bingham's model in the oil and Newtonian Fluid Mech. 226, 1–15.
gas industry. Rheol. Acta 56, 259–282. Roustaei, A., Gosselin, A., Frigaard, I.A., 2015. Residual drilling mud during conditioning
Gavrilov, A.A., Rudyak, V.Y., 2016. Reynolds-averaged modeling of turbulent flows of of uneven boreholes in primary cementing. Part 1: rheology and geometry effects in
power-law fluids. J. Non-Newtonian Fluid Mech. 227, 45–55. non-inertial flows. J. Non-Newtonian Fluid Mech. 220, 87–98.
Gavrilov, A.A., Rudyak, V.Y., 2017. Direct numerical simulation of the turbulent energy Rusche, H., 2003. Computational Fluid Dynamics of Dispersed Two-phase Flows at High
balance and the shear stresses in power-law fluid flows in pipes. Fluid Dynam. 52, Phase Fractions. Ph.D. thesis. Imperial College London (University of London),
363–374. January.
Gori, F., Boghi, A., 2011. Two new differential equations of turbulent dissipation rate and Sawko, R., 2012. Mathematical and Computational Methods of Non-Newtonian,
apparent viscosity for non-Newtonian fluids. Int. Commun. Heat Mass Tran. 38, Multiphase Flows. Ph.D. thesis. Cranfield University.
696–703. Schaer, N., Vazquez, J., Dufresne, M., Isenmann, G., Wertel, J., 2018. On the determi-
Gori, F., Boghi, A., 2012. A three dimensional exact equation for the turbulent dissipation nation of the yield surface within the flow of yield stress fluids using computational
rate of generalised newtonian fluids. Int. Commun. Heat Mass Tran. 39, 477–485. fluid dynamics. J. Appl. Fluid Mech. 11, 971–982.
Guillot, D.J., Desroches, J., Frigaard, I.A., 2007. Are preflushes really contributing to mud Skadsem, H.J., Saasen, A., Håvardstein, S., 2017. Casing centralization in irregular
displacement during primary cementing? In: 2007 SPE/IADC Drilling Conference, wellbores. In: ASME 2017 36th International Conference on Ocean, Offshore and
Amsterdam, The Netherlands, 20-22 February 2007, pp. 1–8 SPE/IADC 105903. Arctic Engineering, Vol. 8. Polar and Arctic Sciences and Technology; Petroleum
Haciislamoglu, M., Langlinais, J., 1990. Non-Newtonian flow in eccentric annuli. J. Technology, pp. 1–10 OMAE2017-61106.
Energy Resour. Technol. 112, 163–169. Szabo, P., Hassager, O., 1992. Flow of viscoplastic fluids in eccentric annular geometries.
Hanks, R.W., 1979. The axial laminar flow of yield-pseudoplastic fluids in a concentric J. Non-Newtonian Fluid Mech. 45 (2), 149–169.
annulus. Ind. Eng. Chem. Process Des. Dev. 18 (3), 488–493. Taghavi, S.M., Alba, K., Moyers-Gonzalez, M., Frigaard, I.A., 2012. Inclomplete fluid-fluid
Herschel, W.H., Bulkley, R., 1926. Konsistenzmessungen von Gummi-benzollösungen. displacement of yield stress fluids in near-horizontal pipes: experiments and theory.
Kolloid Z. 39, 291–300. J. Non-Newtonian Fluid Mech. 167–168, 59–74.
Jakobsen, J., Sterri, N., Saasen, A., Aas, B., Kjosnes, I., Vigen, A., 1991. Displacements in Tardy, P.M.J., 2018. A 3D model for annular displacements of wellbore completion fluids
eccentric annuli during primary cementing in deviated wells. In: SPE Production with casing movement. J. Petrol. Sci. Eng. 162, 114–136.
Operations Symposium, 7-9 April. Society of Petroleum Engineers, Oklahoma City, Tardy, P.M.J., Bittleston, S.H., 2015. A model for annular displacements of wellbore
Oklahoma, pp. 509–517 SPE 21686. completion fluids involving casing movement. J. Petrol. Sci. Eng. 126, 105–123.
Kimura, K., Takase, K., Griffith, J.E., Gibson, R.A., Porter, D.S., Becker, T.E., 1999. Tardy, P.M.J., Flamant, N.C., Lac, E., Parry, A., Sri Sutama, C., Baggini Almagro, S.P.,
Custom-blending foamed cement for multiple challenges. In: SPE/IADC Middle East 2017. New generation 3D simulator predicts realistic mud displacement in highly
Drilling Technology Conference, 8-10 November. Society of Petroleum Engineers, deviated and horizontal wells. In: SPE/IADC Drilling Conference and Exhibition, 14-
Abu Dhabi, United Arab Emirates, pp. 1–10 SPE 57585. 16 March, the Hague, The Netherlands, pp. 1–18 SPE/IADC 184677.
Kong, W., Kong, L., Li, L., Liu, X., Xie, R., Li, J., Tang, H., 2016. The finite element Tehrani, A., Ferguson, J., Bittleston, S.H., 1992. Laminar displacement in annuli: a
analysis for a mini-conductance probe in horizontal oil-water two-phase flow. Sensors combined experimental and theoretical study. In: SPE Annual Technical Conference
16, 1352. and Exhibition, 4-7 October. Society of Petroleum Engineers, Washington, D.C., pp.
Lund, B., Ytrehus, J.D., Taghipour, A., Divyankar, S., Saasen, A., 2018. Fluid-fluid dis- 191–202 SPE 24569.
placement for primary cementing in deviated washout sections. In: Proceedings of the Tehrani, M.A., Bittleston, S.H., Long, P.J.G., 1993. Flow instabilities during annular
ASME 2018 37th International Conference on Ocean, Offshore and Arctic displacement of one non-Newtonian fluid by another. Exp. Fluid 14 (4), 246–256.
Engineering The American Society of Mechanical Engineers, Madrid, Spain, pp. 1–9 Thompson, R.L., Soares, E.J., 2016. Viscoplastic dimensionless numbers. J. Non-
OMAE2018-78707. Newtonian Fluid Mech. 238, 57–64.
Madlener, K., Frey, B., Ciezki, H.K., 2009. Generalized Reynolds number for non- Walton, I.C., Bittleston, S.H., 1991. The axial flow of a Bingham plastic in a narrow ec-
Newtonian fluids. Progress in Propulsion Physics 1, 237–250. centric annulus. J. Fluid Mech. 222, 39–60.
Maleki, A., Frigaard, I.A., 2017. Primary cementing of oil and gas wells in turbulent and Ytrehus, J.D., Lund, B., Taghipour, A., Divyankar, S., Saasen, A., 2017. Experimental
mixed regimes. J. Eng. Math. 107, 201–230. investigation of wellbore fluid displacement in concentric and eccentric annulus. In:
Malekmohammadi, S., Carrasco-Teja, M., Storey, S., Frigaard, I.A., Martinez, D.M., 2010. Proceedings of the ASME 2017 36th International Conference on Ocean, Offshore and
An experimental study of laminar displacement flows in narrow vertical eccentric Arctic Engineering, Vol. 8: Polar and Arctic Sciences and Technology; Petroleum
annuli. J. Fluid Mech. 649, 371–398. Technology. The American Society of Mechanical Engineers, pp. 1–7 OMAE2017-
Meuric, O.F.J., Wakeman, R.J., Chiu, T.W., Fisher, K.A., 1998. Numerical flow simulation 62028.
of viscoplastic fluids in annuli. Can. J. Chem. Eng. 76, 27–40. Zuiderwijk, J.J.M., 1974. Mud displacement in primary cementation. In: SPE European
Nelson, E.B., Guillot, D. (Eds.), 2006. Well Cementing, second ed. Schlumberger, Sugar Spring Meeting, 29-30 May. Society of Petroleum Engineers, Amsterdam,
Land, Texas, US. Netherlands, pp. 1–13 SPE 4830.
OpenFOAM, http://www.openfoam.org.

1013

You might also like