Вы находитесь на странице: 1из 18

Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Contents lists available at ScienceDirect

Journal of Non-Newtonian Fluid Mechanics


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

Viscoplastic fluid displacement flows in horizontal channels: Numerical MARK


simulations
⁎,a
A. Eslamia, I.A. Frigaardb, S.M. Taghavi
a
Department of Chemical Engineering, Laval University, Quebec, QC, G1V 0A6, Canada
b
Departments of Mathematics and Mechanical Engineering, University of British Columbia, 1984 Mathematics Road, Vancouver, BC, V6T 1Z2, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: We study numerically isoviscous displacement flows of a Bingham fluid by a Newtonian fluid in a horizontal
Displacement flow plane channel. We consider a heavy-light displacement flow with a laminar imposed flow. We analyze the flow
Yield stress from various perspectives, using the three dimensionless parameters that largely describe the flow: the Bingham
Buoyancy number (Bn), the densimetric Froude number (Fr) and the Reynolds number (Re). In particular, we are able to
Inertia
categorize three classes for the main flow regimes, viz., center-type and slump-type, no-back-flow and temporary-
back-flow, and stable and unstable displacements, for which we also quantify their transition boundaries versus
the aforementioned dimensionless numbers. We show that the appearance of unstable flows can eventually lead
to a number of different flow patterns, such as sinusoidal shaped wave and periodic-detachment forms. In addition,
we study in depth an interesting feature exclusively associated to viscoplastic fluid displacements, i.e., the
formation of static residual layers on the upper/lower walls of the channel. We also study the evolution of the
key displacement flow features, i.e., the trailing and leading fronts, and classify certain secondary flow features
related to the leading front, such as plug-like, inertial tip, semi-detached and fully-detached fronts. Finally, we
quantify the velocity of the leading front at long times.

1. Introduction in alveoli [24]); cleaning of equipments, environmental surfaces and


stubborn soil [8]; food processing (rinsing pipes and cleaning proces-
Displacement flow in a confined geometry is one of the most sing machineries [17,62]); waxy crude oil pipeline restarts (cleaning
common processes appearing in diverse physical, chemical, biological, the paraffin of crude oil from pipelines [59]); biofilms applications
geophysical, and engineering systems. These fascinating flows are stu- (biological and biotechnological polymer gels [11]). The fluids involved
died from various perspectives thanks to their numerous applications. in these processes can have different properties, e.g., different densities.
Examples may includes: sugar refining (displacement of sugar liquors Displacement flows between miscible liquids frequently occur. Typical
by water in pipe [20]); printing devices (displacement of lubrication oil flow geometries include pipe, annulus, channel, etc. In this work, we
in printing rollers [58]); carbon sequestration (injection of carbon di- numerically consider buoyant miscible displacement flows of a visco-
oxide into saline aquifer [10]); chromatographic separations (dis- plastic fluid by a Newtonian one, where the imposed flow is laminar.
placement of fluids with different viscosities in packed chromato- A miscible displacement flow, i.e., the focus of our work, is funda-
graphic columns [46]); enhanced oil recovery (displacement of oil by mentally different from an immiscible one. When the displacing and
injection of gas or water into reservoir [26,38]); coating (spreading and displaced fluids are immiscible (i.e., σ ̂ > 0, with σ ̂ being the surface
fingering of silicone oil in spin coating [21]); adhesives (penetrating air tension coefficient), the density and viscosity remain constant in each
into oil during lifting a circular cell [36]). Displacement flows some- phase and a pressure jump at the interface is considered. However,
times occur in industrial processes where it is essential to remove a when the displacing and displaced fluids are miscible, there is no in-
gelled-like material (i.e., typically viscoplastic) from interior geome- terface, per se, between the phases so that diffusion and mixing between
tries. Examples include: oil & gas well cementing (displacement of them need to be considered. In this case, a relevant dimensionless flow
drilling mud by cement slurry in annulus [6,37]); gas-assisted injection parameter is therefore the Péclet number, Pe, (see Table 1 for the de-
molding (plastic manufacturing [40]); biomedical applications (pro- finition), which represents the ratio between advective and molecular
pagation of a liquid bolus along a tube [23] and clearing of mucus plugs diffusive transports. The diffusion between the displacement flow


Corresponding author.
E-mail address: seyed-mohammad.taghavi@gch.ulaval.ca (S.M. Taghavi).

http://dx.doi.org/10.1016/j.jnnfm.2017.10.001
Received 28 July 2017; Received in revised form 27 September 2017; Accepted 4 October 2017
Available online 05 October 2017
0377-0257/ © 2017 Elsevier B.V. All rights reserved.
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Table 1 viscoelasticity has been shown to create peculiar patterns and affect the
0
Definitions and ranges of the dimensionless parameters used in this manuscript. Here, V efficiency of displacement flows. Interfacial instabilities for viscoplastic
is the mean imposed velocity. In the definition of the Reynolds number, ν ̂ = ηĤ / ρ ̂ with ηĤ
displacements have been studied first by Pascal [41–43] in the context
being the viscosity of the heavy fluid and ρ ̂ = (ρL̂ + ρĤ )/2 . In the definition of the Péclet
number, D m is the molecular diffusivity. m is 1 everywhere, unless otherwise stated. of porous media. The Hele-Shaw cell viscoplastic displacement regimes
Hereafter, the densimetric Froude number, denoted by Fr, is called the Froude number for have been classified by Lindner et al. [27,28], who have shown that at
convenience. low imposed flow rates yield stress effects are dominant and ramified
structure patterns are formed. At high imposed flow rates, viscous ef-
Parameter Name Definition Range or value
fects are dominant and smooth displacing fingers appear [27,30]. At
Bn Bingham number 
τŷ D 0 − 200 extremely high imposed flow rates, yield stress effects are completely
0
μL̂ V eliminated and side-branching patterns are exhibited [14,27,30].
Fr Densimetric Froude number 0
V 0.1 − 1000 Non-Newtonian displacement flows in geometries other than the

̂
Atg D
Hele-Shaw cell have been less explored. de Sousa et al. [50] numeri-
m Viscosity ratio ηL̂ 1 cally analyzed displacement flows of viscoplastic materials in capillary
ηĤ tubes. They found that the value of residual layers of viscoplastic fluids
Re Reynolds number 0 D
V  50 − 500 on the wall is decreased by increasing the dimensionless yield stress
ν̂ value. Freitas et al. [16] studied immiscible viscoplastic-viscoplastic
δ Aspect ratio L̂ 100 displacement flows in a capillary plane channel, using numerical ap-

D proaches. They showed that the dimensionless yield stress number of
Pe Péclet number 0 D
 ≫1
V the displaced fluid (with respect to the one of the displacing fluid) has a
Dm
significant impact on the residual layer thickness. Gabard & Hulin [18]
studied the effects of non-Newtonian rheology on vertical tube dis-
placements. They showed that the transient residual film thickness for
phases becomes important when Pe is small, while at large Pe the de-
shear-thinning fluids is larger than that for viscoplastic fluids. Pa-
gree of molecular diffusive transport compared to advective transport is
paioannou et al. [39] studied displacement of a viscoplastic fluid by air
very small, implying that the two fluids do not have sufficient time to
in a pipe, demonstrating that a detachment of the viscoplastic material
mix in a typical displacement flow process. Previous studies [9,44] have
from the solid wall can take place at sufficiently large Bingham num-
shown that, in the absence of flow instabilities, the flow at high-Pe
bers (see Table 1). Swain et al. [51] studied computationally visco-
approaches its immiscible counterpart at the zero surface tension limit
plastic displacement flows in 2D channels. They showed that interfacial
(σ ̂ → 0 ) and consequently the interface between the two fluids remains
instabilities are decreased by increasing the yield stress and the flow
quite sharp. Knowing that many of displacement flow applications
index (which characterizes the shear-thinning tendency) of the dis-
occur at large Pe, from a modeling perspective, Pe → ∞ provides a good
placed fluid. In addition, they mentioned that increments in the flow
approximation to the flow behaviors. Note that the fluids can still mix
index lead to increasing the size of the unyielded region. Dimakopoulos
due to hydrodynamic effects (e.g., instabilities, secondary flows). In a
& Tsamopoulos [13] considered the displacement of a viscoplastic
numerical framework, mixing at the grid level is dominated by nu-
material by a Newtonian fluid in straight tubes (symmetric) and sud-
merical diffusion.
denly constricted tubes (asymmetric). They found that unyielded re-
Buoyant miscible displacement flows for Newtonian fluids have
gions arise in front of the displacing fluid and near the recirculation
been studied in detail, for various flow geometries such as circular
corners for straight and constricted tubes, respectively. Moisés et al.
pipes, 2D channels and Hele-Shaw cells. Here, we briefly review a few
[33] experimentally investigated the residual layer variation amplitude
of the relevant studies. For example, Seon et al. [49] have studied these
in a horizontal pipe when a Newtonian fluid displaces a viscoplastic
flows at the limit of zero imposed flow velocity (i.e., an exchange flow)
fluid. They found that the competition between the ratio of inertia to
in inclined pipes, revealing the competition between buoyant, viscous
viscous forces of the viscoplastic fluid leads to three different flow re-
and inertial forces that result in the formation of various flow regimes
gimes, with smooth, wavy or corrugated interface, respectively. Ta-
(e.g., a viscous-dominated or an inertial-dominated regime). In a series
ghavi et al. [55] studied semi-analytically heavy-light & light-heavy
of studies, Taghavi et al. [54,56,57] have considered numerically,
displacements of viscoplastic fluids in near-horizontal 2D channels.
analytically and experimentally buoyant miscible displacement flows in
They demonstrated that a yield stress in the displacing fluid enhances
near-horizontal pipes and 2D channels, for a wide range of flow para-
the displacement flow efficiency and a yield stress in the displaced fluid
meters (e.g., Re, Fr and pipe/channel inclination), and they have
leads to decreasing the displacement flow efficiency. Taghavi et al. [53]
quantified the appearance of different flow regimes when an imposed
and Alba et al. [2] performed experiments for yield stress fluid dis-
flow is added to an exchange flow. Earlier studies such as Chen &
placement flows, at near-horizontal and highly-inclined pipes, respec-
Meiburg [9] have analyzed displacement flows in capillary tubes at the
tively. They analyzed displacement flow exotic patterns at the condition
limit of high-Pe, using numerical approaches. Sahu et al. [48] have
when the yield stress in the displacing fluid is much larger than the
considered these flows along inclined channel, finding that mixing and
typical viscous stress. They identified two different patterns for heavy-
displacement rates are improved with increasing Fr and the density
light displacement flows: (1) a center-type and (2) a slump-type dis-
ratio between the two fluids. Amiri et al. [5] have recently focused on
placement flow regime, based on the density ratio between the two
vertical pipe flows, observing a stabilizing effect of the imposed flow for
fluids. The experimental-theoretical approach of Alba & Frigaard [1]
small density differences.
confirmed the formation of the two aforementioned flow patterns for
Due to complex rheological behaviors of non-Newtonian fluids, the
these fluids. In addition, they found that the front velocity in the slump-
literature of non-Newtonian displacement flows is less developed
type case is larger than the center-type case.
compared to that for Newtonian fluids. For these flows, a majority of
A crucial aspect associated of viscoplastic displacements is the for-
studies concern the well-known Saffman–Taylor instability (or the vis-
mation of static residual layers of the displaced fluid on the walls of the
cous fingering instability) and the associated pattern formations in the
flow geometry. These layers are hard-to-remove especially from interior
Hele-Shaw geometry (see review by McCloud [32]). The effects of
sections. The phenomenon of static residual layers usually occurs when
various non-Newtonian parameters, such as yield stress [12,27,30,42],
the yield stress of the displacing fluid is smaller than that of the dis-
shear-thinning [7,15,60], shear-thickening [7,15,35], elasticity [25,31]
placed fluid [4]. A relevant early work in this context is that of Poslinski
and elaso-visco-plastic properties [14], on displacement flows have
et al. [45], who experimentally studied the displacement of a visco-
been investigated numerically, experimentally and analytically. Fluid
plastic fluid by a Newtonian fluid (air) in a tube, finding that the

80
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

residual layer thickness is much larger in comparison with Newtonian


fluids (increasing up to 0.35 of the tube radius at high imposed flow
rates). Allouche et al. [4] showed that, unlike Newtonian fluids, the
residual wall layers for viscoplastic fluids can be completely static due
to their yield stress. Using a combined theoretical and computational
approach, they quantified the residual wall layer thickness for a sym-
metric displacement flow of two Bingham fluids flowing in a 2D
channel. Wielage–Burchard & Frigaard [61] studied computationally
the effects of the Reynolds number (Re), the Bingham number (Bn) and
the viscosity ratio (m) on the static layer thickness for isodense dis-
placements. They mentioned that the static residual wall layers are
quasi-uniform, but they did not provide information about the location/
time of the appearance of these layers. Mollaabbasi and Taghavi [34]
analytically studied viscoplastic fluid displacement flows in non-uni-
form 2D channels, showing that the channel convergence/divergence
decreases/increases the static residual wall layer thickness.
The outline of this manuscript is as follows. Below in Section 2, the
problem setting is discussed. The details of the computational metho-
dology and the benchmarking of our code are explained in Section 3.
Our main results are presented in Section 4. The paper ends with a brief
summary in Section 5.

2. Problem setting

Our computational study was performed in a plane channel, as re-


Fig. 1. Schematic view of the numerical domain a) at the initial flow configuration and b)
presented in Fig. 1. Fig. 1a shows a schematic of the flow geometry,
after the onset of displacement flow.
with the initial flow configuration and the notation used in this
manuscript. The displacement of a Bingham fluid (fluid L in Fig. 1) by a
Newtonian one (fluid H in Fig. 1) along a channel of width D  is con- as τ (u) = γ˙ (u) with γ˙ = ∇u + (∇u)T . For the displaced Bingham fluid,
sidered.1 The viscosity ratio (m), defined as the ratio of the plastic the constitutive law includes a yield stress and it is written as
viscosity ( μL̂ ) to the Newtonian fluid’s viscosity ( μĤ ), is one. The fluids
Bn ⎤
have different densities: the displacing fluid is denser than the displaced τ2 (u) = m ⎡1 + γ˙ (u) ⇔ τ2 (u) > mBn,
⎢ ˙
γ (u) ⎥
fluid. After the onset of displacement flow, the displacement flow ⎣ ⎦ (4)
configuration has two fronts: a leading front and a trailing front, as de-
picted in Fig. 1b. Cartesian coordinates (x ,̂ y )̂ are considered with x ̂ γ˙ (u) = 0 ⇔ τ2 (u) ≤ mBn, (5)
representing the stream-wise direction. The dimensionless equations of
where the second invariants, γ˙ (u) and τ2(u), are defined by
motion are
1/2
2
∂ ϕRe ⎡1 ⎤
[1 + ϕAt ] Re ⎡ u + (u . ∇) u⎤ = − ∇p + ∇ . τ + eg , γ˙ (u) = ⎢ ∑ [γ˙ij (u)]2 ⎥ ,
⎣ ∂t ⎦ Fr 2 (1) 2 i, j = 1
⎣ ⎦
1/2
2
∇ . u = 0, (2) ⎡1 ⎤
τ2 (u) = ⎢
2
∑ [τ2, ij (u)]2 ⎥ ,
1 2 ⎣ i, j = 1 ⎦ (6)
Ct + u·∇C = ∇ C,
Pe (3)
where m is the viscosity ratio. Table 1 shows the ranges of the di-
where u, p and τ the velocity, the pressure and the deviatoric stress, mensionless parameters used in our work. As can be seen, our study
respectively. Here eg = (0, −1) and the function ϕ (C ) = 2C − 1 varies (with ∼ 400 simulations performed) covers a wide range of these
linearly between − 1 and 1 for C ∈ [0, 1]. No slip boundary conditions parameters.
are applied at the walls. The fully developed velocity profile (plane
Poiseuille profile) and the outflow boundary conditions are prescribed 3. Computational code
at the inlet ( x = −L/4 ) and the outlet ( x = 3L/4 ) of the channel, re-
spectively. The system (1)–(5) was discretized using a mixed finite element-
The governing dimensionless numbers that appear in (1)–(3) are the finite volume method, using the classical augmented Lagrangian ap-
Reynolds number (Re), the densimetric Froude number (Fr), the At- proach of [19] to resolve unyielded zones. More details of the numer-
wood number (At) and the Péclet number (Pe). The Atwood number is a ical method are explained in [47,61]. The present numerical algorithm
dimensionless density difference, shown as At = (ρĤ − ρL̂ )/(ρĤ + ρL̂ ), in was implemented in C++ and solved using PELICANS (available at
which the density of the displacing fluid is represented by ρĤ and that https://gforge.irsn.fr/gf/project/pelicans/), which is an open source,
of the displaced fluid by ρL̂ . Since ρĤ > ρL̂ (i.e., the heavy fluid pushes
object oriented platform to solve PDEs. PELICANS is shared under the
the light fluid), At > 0 in this paper. In addition, displacement flows
CeCILL free software license agreement (http://www.cecill.info/
with small density differences are considered, for which the Boussinesq
licences/Licence_CeCILL_V2-en.html, 2010). More details about using
approximation is implemented. The other parameters are defined in
this platform for displacement flow simulations are given in [57,61].
Table 1.
In our simulations, 63,000 mesh cells (1500 × 42) were typically
The constitutive law for the displacing Newtonian fluid is presented
used while the meshes (regular rectangular) in the y-direction were
refined slightly towards the channel walls. The initial interface between
1
In this paper we adopt the convention of denoting dimensional quantities with the ^ the two fluids was placed at distance L/4 from the channel inlet in the
symbol and dimensionless quantities without. computational domain (see also Fig. 1a).

81
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

1 − C (t )
Fig. 2. a) Evolution of versus time for dif-
1 − C (0)
ferent mesh sizes for Bn = 100, Re = 50 and
Fr = 1000 . b) Evolution of the position of the dis-
placing front, xfront, versus time for different mesh
sizes for the same simulation. c&d) Concentration
colormaps at t = [0, 10, ...,40] for Bn = 5, Re = 200
and Fr = 0.5 for the mesh size c) 1500 × 42 and d)
2100 × 34. The domain size shown is 1 × 100.

The most obvious global feature of the displacement flow is the displacing front position. Mesh refinement from 1500 × 42 to
average value of the concentration, defined as 2100 × 34 leads to a small deviation of 0.75%.
Moreover, to illustrate that our results are not affected by the mesh
1 1 3L /4
C (t ) =
L
∫0 ∫−L/4 C (x , y, t ) dxdy.
(7)
density, simulations with two different mesh densities are performed.
Fig. 2c and d display the concentration colormaps, at different times for
For different mesh densities, the dependency of (1 − C (t ))/(1 − C (0)) Bn = 100,Re = 300 and Fr = 0.1, for 63,000 mesh cells (1500 × 42) and
and the position of the displacing front, xfront, for parameters Bn = 100, 71,400 mesh cells (2100 × 34), respectively. For these conditions, the
Re = 50 and Fr = 1000, are given in Fig. 2a and b, respectively. Ob- front detachment phenomenon (due to flow instabilities) is observed.
viously at t = 0, the quantity (1 − C (t ))/(1 − C (0)) approaches unity As seen, there exists good agreement between the results of the two
and it decreases with time as more displacing fluid (with C = 1) is in- simulations, in which the mesh refinement results in a small deviation
troduced into the computational domain. As can be seen, for small mesh of 1.54% in the mean wavelength of 10 interfacial waves (closest to the
sizes, a good convergence for the variation of (1 − C (t ))/(1 − C (0)) is channel end) at t = 40 .
achieved.
As Fig. 1a illustrates, initially (at t = 0 ) an imaginary gate valve 3.1. Code benchmarking
separates the two fluids at x = 0, meaning that interface between two
fluids is distinct and sharp at the beginning of the displacement process. We have validated our code by comparing our results against those
At t > 0, due to the mean imposed flow rate, the heavy layer starts to obtained in Taghavi et al. [52,54] and Alba et al. [3] for Newtonian
push the light fluid. Initially, the axial position of the displacing front displacement flows. For the Bingham fluids, we have benchmarked our
(xfront) is zero and as the time progresses the distance between xfront and code against the results of Wielage-Burchard & Frigaard [61], who
the initial interface position increases. Fig. 2b displays xfront versus time considered isodense displacement flows of a Bingham fluid (i.e., Fr →
for the same mesh densities as in Fig. 2a. It is evident that the position ∞). Fig. 3a shows an example where our results are compared against
of the displacing front varies only slightly with the mesh sizes chosen. those of Wielage-Burchard & Frigaard [61], with Re = 100,Fr = 1000 .
Comparing 13,600 mesh cells (400 × 34) to 71,400 mesh cells As can be seen, the results obtained from our code follow precisely the
(2100 × 34), the maximum relative error is 3.81% for the final value of results presented in [61]. Fig. 3b shows a qualitative comparison

Fig. 3. Code benchmarking: a) The average static re-


sidual wall layer thickness, have = (hu + hl )/2, from
our simulation ( ) at Re = 100 and Fr = 1000 against
the results of [61] ( ) for Re = 100 and Fr = ∞. b)
Concentration colormaps of the displacement flow at
times t = [0, 7, ...,28] for Bn = 10, Re = 100,m = 2 and
Fr = 7.071. The last image at the bottom of the sub-
figure is the colorbar of the concentration values (here
and elsewhere). The white broken lines display the
position of the initial interface x = 0 (here and later).
The domain size shown is 1 × 100 (unless otherwise
stated). The images in this subfigure can be qualita-
tively compared with Fig. 2 in [51]. To validate the
code against the results of Swain et al. [51]. we have
exceptionally considered a light-heavy displacement
flow. (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.)

82
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

against the results of Swain et al. [51], who investigated the displace- the static layers seem to decrease with Bn.
ment flow of a viscoplastic fluid by a lighter Newtonian fluid in a 2D
channel via a multiphase lattice Boltzmann method. This subfigure 4.2. Static residual wall layers
shows the concentration colormaps of a displacement flow in a hor-
izontal channel for Bn = 10,Re = 100,m = 2 and Fr = 7.071 (note that As discussed in the introduction section, static residual wall layers
for comparison against the results of Swain et al. [51], we have per- are of great significance in viscoplastic fluid displacement flows. In this
formed a light-heavy displacement simulation). The overall behaviors subsection, we focus on describing qualitative and quantitative features
are quite similar to those reported in Fig. 2 given in [51]: due to the of these layers in our computational results.
imposed flow, first the displacing finger penetrates into the bulk of the Fig. 5 shows examples of our results, for Bn = 50,Re = 500 and
displaced fluid in a symmetric way but soon the finger becomes slightly Fr = 1. Initially at x = 0, the displacing and displaced fluids are com-
asymmetric under the effect of gravity. In addition, we observe the pletely separated by an imaginary gate valve. As time grows, due to a
appearance of Kelvin–Helmholtz-like instabilities, more or less in the combined effect of the imposed flow and the density ratio, the heavy
same spatial locations as in [51]. fluid penetrates through the light fluid and attempts to push it out of
the channel. Fig. 5a displays the concentration colormaps of the dis-
4. Results and discussions placement flow for various times. As can be seen in this subfigure, at
longer spatial positions with respect to the initial gate valve, the dis-
In this section, we present and discuss our main findings for a wide placed layers adjacent to the walls become apparently static. Accord-
range of the flow parameters. In a typical simulation, as time pro- ingly, the displacing finger moves in a narrower channel (created by the
gresses, depending on the imposed mean velocity, the heavy fluid pe- upper and lower static layers) and advances significantly faster than the
netrates into the light fluid and displaces it. During the displacement mean imposed flow. Fig. 5b shows the shear stress colormaps at the
process, various interesting patterns are formed at the interface be- same times as the colormaps of concentration, revealing that the stress
tween the two fluids, which we will explain below in more detail. Our field asymmetry results in asymmetric static layers on the top/bottom
simulation results can be divided in three subcategories: i) quantifying walls.
static residual wall layers; ii) quantifying the main displacement flow For convenience, let us call the spatial locations where the static
regimes; and iii) analyzing displacement front velocities. residual layers become uniform as the upper static distance (xu) and the
lower static distance (xl), for upper and lower static residual layers,
4.1. Note of the effects of Bn on displacement flows respectively. Similarly, let us call the associated uniform static layer
thicknesses as the upper static layer thickness (hu) and the lower static
Before we proceed with presenting various flow features in the layer thickness (hl). In order to better understand the definitions of hu,
following subsection, it is worth discussing general effects of con- hl, xu and xl, at a given large time (here t = 40 ), they are plotted in
sidering a yield stress fluid as the displaced entity in place of a Fig. 5c, where a few velocity vectors are also plotted. As seen, the
Newtonian fluid. This can help clarify two aspects: (a) the difference of displaced fluid layers close to the top and lower walls are entirely static,
our work with the large body of recent computational studies, con- where the concentration of displacing fluid is zero. This means that the
sidering Newtonian displacements in channels; (b) general effects that displaced fluid layers have a zero velocity and therefore hu and hl can be
could be expected before analyzing all the simulations results. computed. Initially, the static layer thicknesses vary with x but reach
Fig. 4 presents panoramas of the concentration colormaps at a given plateau values after certain distance x, where uniform static residual
time, for fixed Re and Fr. By increasing Bn, the displacement flow layers in the upper and lower walls appear.
morphology changes significantly. Fig. 4a shows that, at small Fr, by Fig. 5d shows the speed contours, V = Vx 2 + Vy 2 , at t = 40, where
increasing Bn the interfacial instabilities decrease and the yield stress Vx and Vy denote the stream-wise and depthwise velocity components,
damps the interfacial waves. In addition, the trailing front, which respectively. The velocity profile is evidently similar to a Poiseuille
moves upwards at small Bn, completely stops moving at larger Bn. profile before the gate valve location such that the high speed regions
However, the variation in Bn does not seem to influence the slumping remain towards the channel center. After the gate valve, the speed
form, which is close to a 2-layer flow. Finally, as will be illustrated in contours are zero within the displaced fluid layers adjacent the upper
the following sections, at large Bn the displaced layer above the dis- and lower walls, which means that the remaining displaced layers are
placing finger becomes static (motionless). completely static. Fig. 5e and f show the velocity vector fields along the
Fig. 4b shows that, at large Fr, at Bn = 0 (i.e., the Newtonian limit) channel length, confirming that the existence of completely static re-
the displacement is nearly symmetric (e.g., the displacing finger ad- sidual layers wherein the velocity is zero. It should be noted that the
vances approximately along the center of the channel) and that the velocity profiles within the displaced fluid further downstream (where
displacing finger resembles a Poiseuille-like profile. At larger Bn, the the displacing front has not reached) are similar to a typical plug-type
flow remains symmetric, although the front shape changes and re- profile of a Bingham fluid in a channel. The velocity profiles of dis-
sembles a plug-like profile. Moreover, while at Bn = 0 there are no placing fluid in the finger are similar to a stable viscous profile over a
static wall layers, these layers are formed at larger Bn. The thickness of narrower channel; however, inertial effects exist at the displacing front.

Fig. 4. Panorama of concentration colormaps at t = 16 and


Re = 400 for a) Fr = 0.5 and b) Fr = 100 . The rows from top
to bottom show Bn = 0, Bn = 2, Bn = 5, Bn = 20, Bn = 50,
Bn = 100, Bn = 200 . The domain size shown is 1 × 60.

83
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 5. Computational results for Bn = 50, Re = 500 and Fr = 1; a) Concentration colormaps at times t = [0, 10, ...,40]; b) Shear stress colormaps at times t = [0, 10, ...,40]; c) Interface
heights (lower and upper layers) at t = 40 : hu and hl show the thickness of upper and lower static layers, respectively. The red arrows indicate the position at which uniform static layers
appear, with xu indicating the upper static distance and xl the lower static distance. d) Speed contours: V = Vx 2 + Vy 2 . e) Velocity vectors. f) The image is zoomed-in on the indicated box
of subfigure. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Our study is mainly governed by Re, Bn and Fr. Thus, it is logical to (when hu ≫ hl), a slump-type displacement regime is observed where the
attempt to quantify the effects of these parameters on the upper and displacing finger is found more or less near the lower wall of the
lower static residual layers in our displacement flows. Fig. 6 shows the channel; at high Fr (when hu ≈ hl), a center-type displacement regime is
effects of varying the Bingham number on the thicknesses of the upper found where the displacing finger appears approximately in the middle
and lower static residual layers, for various Reynolds and Froude of the channel. For Re = 100, as seen in Fig. 7e, Fr = 2, for which hu/
numbers. There are missing data for certain flows at very small Froude hl ≈ 1.69. Fig. 7f displays the variation of hu/hl versus Bn for Re = 1. At
numbers for which there are no uniform static residual layers formed. these conditions the flow is more or less symmetric about the channel
The subfigures demonstrate that for a wide ranges of Re and Fr, in- center and hu/hl tends to unity. Two other conclusions drawn from
creasing Bn generally results in decreasing both hu and hl. In fact, in- Fig. 7d– f can be summarized as: (1) the value of Fr (when the plateau
creasing Bn of the displaced fluid influences the plug ahead of the state is observed) is a function of Re and (2) hu/hl is more or less in-
displacing finger and results in a reduction of the static residual wall dependent of Bn.
layer thickness [61]. The subfigures in Fig. 6a show that the upper
static residual layer thickness decreases as Fr increases, whereas the 4.2.1. Comparison between our simulations and theoretical models from the
ones in Fig. 6b show that the lower layer thickness is enhanced by in- literature
creasing Fr. These effects may be expected since increasing Fr is In recent years, there has been some progress in analyzing static
equivalent to decreasing the density difference, an effect which pro- residual layers of viscoplastic fluids. In this subsection, we will explain
gressively pushes the displacing finger towards the channel center. a few of the previous studies and attempt to make comparison between
Comparison among different subfigures in Fig. 6a and b reveals their findings and our simulation results.
useful information. For example, for various Fr the effect of Re on hu is Allouche et al. [4] were first to rigorously investigate theoretically
not monotonic: hu increases by increasing Re at small Fr, while hu de- and computationally static residual wall layers of Bingham fluids. For
creases by increasing Re at large Fr. The effect of Re on hl is more or less miscible displacement flows in a 2D channel and in symmetric config-
monotonic such that by increasing Re, hl decreases for all values of Fr. urations, they showed that the static wall layer thickness can be pre-
In order to better observe some of the effects explained, Fig. 7a–c dicted via the recirculation layer thickness, hcirc, which is a thickness
show the variations in hu and hl for increasing Froude numbers, for corresponding to a steadily displacing finger advancing with the speed
different Bn, at fixed Re = 500,Re = 100 and Re = 1. These subfigures equal to that of the downstream flow center-line. Upon the condition
show that, for a given Bn, hu decreases initially with Fr but finally that h < hcirc, a recirculatory region ahead of the displacing finger
reaches a plateau value. On the other hand, hl increases by increasing Fr would occur, increasing viscous dissipation as a response. Therefore,
but eventually becomes constant. Although the observed behaviors for Allouche et al. [4] concluded that the displacement flow adjusts to
hu and hl at Re = 500,Re = 100 and Re = 1 are more or less the same, the avoid such situation. hcirc is defined by the downstream flow, as follows
plateau values are reached at different values of Fr. [4,61]:
The ratio of the upper to lower static layer thickness (hu/hl) may
help gain further insight about the formation of these layers as well as 2Y
hcirc = 1 − ,
the overall displacement flow. Fig. 7d–f show the variation of hu/hl Bn (1 − Y )2 (8)
versus Fr for three Re and different Bn. At Re = 500 (Fig. 7d), initially where Y can be found by the solution of
hu/hl sharply decreases with Fr but eventually reaches a near-plateau
value for all Bn. One may find the value of Fr when the plateau state is Y 3 − 3Y ⎡1 +
2
⎤ + 2 = 0,
reached roughly at Fr = 10, where hu/hl ≈ 1.46. As illustrated by the ⎣ Bn ⎦ (9)
superimposed snapshot images, this transition point also interestingly
for a given Bn. The aforementioned theory suggests that the residual
corresponds to a significant change in the flow morphology: at low Fr
wall layer thickness has an inverse relation to Bn, which is consistent

84
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 6. Effects of Bn, Fr and Re on static residual layer thicknesses: a) hu; b) hl. The data correspond to Re = 50 ( ), Re = 100 ( ), Re = 200 ( ), Re = 300 ( ), Re = 400 ( ) and Re = 500
( ). In this and the following figures, the error-bars (estimated through the standard deviation of the static layer thicknesses) are shown in one graph only, while the error-bars of the
other data are more or less similar (not shown). The insets are zoomed at the variation of the lower static thickness.

with our simulation results. Fig. 8a and b compare our numerical simulation results of the static
Taghavi et al. [55] also studied analytically the maximal static re- layer thickness with hmax,u, as explained above. It can be seen that, first
sidual wall layer thickness, hmax , for 2D viscoplastic displacement flows of all, the variation of our results versus Bn appears to follow the same
in a slumping 2-layer configuration. They used a lubrication/thin-film trend as hmax,u. However, the results indicate that at small Froude
approximation to quantify these layers in near horizontal channel in- number (Fr = 1) the model underestimates the static layer thickness
clinations, showing that, for Bingham fluids, the layer thickness de- whereas at large Froude number (Fr = 100 ), the model overestimates
pends on the ratio of the yield stress of the fluids and the ratio of axial the static layer thickness. In fact, using hmax,u is not a precise model
buoyancy stress to the yield stress of the displaced fluid. In our context since, among other factors, it does not take into account the re-
(i.e., a strictly horizontal channel), the axial buoyancy stress, which circulatory front region and variation of Fr. Fig. 8c and d demonstrate
depends on the interface slope, would be negligible at long times, im- that the trends of our results and hcirc are similar, both decreasing with
plying that hmax is controlled by the Bingham number only. Here we Bn. However, the calculation of hcirc through the symmetric 3-layer
extend their analysis of hmax for our asymmetric configuration. We model does not fully agree with our simulation results, at least due to
crudely can assume that the upper maximal static residual wall layer following reasons: First, our flow is asymmetric with the respect to the
thickness, hmax,u, is a function of the Bingham number and the center- channel center-line; Second, our simulations are run for a wide range of
line of the displacing fluid layer (behind the front) donated as y0 (ob- Fr, which is not included in hcirc; Third, our displacement flows are run
tained from our simulations). Therefore, hmax,u can be found as at considerable Re, which is absent in the analysis of hcirc. In addition,
hmax, ave (extended form the analysis above) is also unable to predict the
12 − 12y0 1/3 average static layer thickness.
hmax , u = ⎛ ⎞ .
⎝ Bn ⎠ (10)

85
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 7. The upper row shows the variation of hu (filled symbols) and hl (hollow symbols) versus Fr for different values of Bn at: a) Re = 500 ; b) Re = 100 ; c) Re = 1. The lower row shows
the variation of hu/hl versus Fr for: d) Re = 500 ; e) Re = 100 ; f) Re = 1. The data correspond to Bn = 2 ( ), Bn = 5 ( ), Bn = 20 ( ), Bn = 50 ( ), Bn = 100 ( ), Bn = 200 ( ).

Fig. 8. Comparison among static layer thicknesses from our


simulation results (filled symbols), hcirc (dashed line) and
hmax,u or hmax, ave (hollow symbols). The upper row shows hu
and the bottom row shows the mean value of the lower and
h +h
upper static layer thicknesses (have = u l ). The data cor-
2
respond to Re = 100 ( ) and Re = 500 ( ). The insets indicate
the ratio of the simulation upper layer thickness to the upper
maximal static layer thickness, hu/hmax , versus Bn for the
same datapoints as in the main graphs.

4.2.2. Static layer distance such that when the yield stress is large the distance between the ima-
In this subsection, we explain our main findings about the distance ginary gate valve and the uniform static layers is small and vice versa.
from the gate valve where uniform static layers appear in the channel. Moreover, Fig. 9 shows that the variations of xu and xl as a function of
For a typical simulation, we study the impact of Re, Bn and Fr on the Re and Bn are monotonic at large Fr, while non-monotonic behaviors
upper and lower static distances. are observed at small Fr. The latter slightly strange feature may be at-
Fig. 9 shows the effects of varying the Bingham number on xu and xl tributed to large buoyancy effects.
for various Reynolds and Froude numbers. For a wide range of Re and Fig. 10 shows the ratio of lower to upper static distances, xl/xu, for
Fr, increasing Bn generally results in decreasing both xu and xl. This different Fr, Bn and Re. The upper row shows that increasing the Froude
means that for larger Bingham numbers, the static residual layers be- number (from Fr = 1 to Fr = 100 ) leads to the fact that all the curves
come uniform closer to the initial interface. Furthermore, by increasing reach more or less xl/xu ≈ 1, implying that the flow becomes progres-
Re generally (not always) xu and xl increase. It seems that the locations sively more symmetric as Fr increases (or the density difference de-
wherein the uniform static layer appear are controlled by a balance creasing); thus, the uniform static layers are formed at the same upper
between the yield (stabilizing) and inertial stresses (destabilizing). The and lower locations. However, the lower row shows the same trend
interplay between these stresses results in creating uniform static layers appears if Bn decreases. Increasing Bn also helps the formation of static

86
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 9. Effects of Bn, Fr and Re on a) xu; b) xl. The data correspond to Re = 50 ( ), Re = 100 ( ), Re = 200 ( ), Re = 300 ( ), Re = 400 ( ), Re = 500 ( ). The insets display the same data
as in the main graphs but with a linear scale.

Fig. 10. The ratio xl/xu. In the upper row, the data correspond to Re = 50 ( ), Re = 100 ( ), Re = 200 ( ), Re = 300 ( ), Re = 400 ( ), Re = 500 ( ). In the lower row, the data
correspond to Fr = 0.5 ( ), Fr = 1 ( ), Fr = 2 ( ), Fr = 10 ( ), Fr = 100 ( ), Fr = 1000 ( ).

87
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

layers close to initial interface of two fluids (both xu and xl become In relation to our work but for Newtonian fluids, Taghavi et al. [57]
small). Regarding the effect of Re on xl/xu, we have not succeeded to have also studied in depth the movement of the trailing front in dis-
reach to a definitive conclusion. placement flows and they have quantified various possible movements
of the trailing front for near-horizontal pipe/channel inclinations.
4.3. Main flow regimes Fig. 12a and b show the concentration colormaps as time evolves,
for a no-back-flow displacement and a temporary-back-flow displace-
Our displacement flows present a variety of fascinating, complex ment, respectively. In Fig. 12a the trailing front seems to be pinned to
behaviors; therefore, it is useful to try to classify various flow regime the upper wall, so that its speed is zero throughout the simulation. In
observed. Fig. 12b, the trailing front initially moves upward but its speed gra-
dually decreases until it becomes zero. The explanation for this beha-
vior may be simple: the buoyancy stress associated to the interface
4.3.1. Center-type and slump-type regimes slope is initially large but it decreases as the interface elongates and
One example of the flow regimes that appear in our simulations is eventually reaches zero when the interface becomes nearly parallel to
the classification of slump-type and center-type flow regimes, which we the channel walls. Fig. 12c and d show the spatiotemporal diagram of
briefly explained earlier. In the center-type regime (hu/hl → 1), the the depth-averaged concentration values, where the advancement of
heavy displacing fluid flows as a finger more or less in the middle of the the trailing and leading fronts with time are clearly observed.
channel, while in the slump-type regime (hu/hl ≫ 1) the displacing layer Let us classify our results based on the appearance of the no-back-
moves closer to the lower region of the channel, below the bulk of the flow and temporary-back-flow regimes. In order to find the suitable
displaced fluid. We observed that slump-type displacements appear at dimensionless groups governing the transition between the no-back-
large density differences (small Fr), more or less independent of Bn. The flow and temporary-back-flow displacement flows, various combina-
latter is a feature that has been seen experimentally by Taghavi tions of the dimensionless numbers were examined. Fig. 13 classifies
et al. [53], but it has never been quantified computationally. Our si- the simulation data points in the plane of Fr and Re/Bn, where the two
mulations show that beyond a transition value of Fr, the ratio of the regimes are clearly segregated. Temporary-back-flows are observed for
upper to lower static layer thickness is not much affected by Re and higher Fr. Although at smaller Fr and fixed Re, more temporary-back-
remains at hu/hl ≈ 1.6, on average. We have used this criterion to flows are observed if Bn decrease, in general the transition between the
identify the slump- and center-type regimes. The flow deemed to be regimes has only a small dependency on Bn, implying the dominance of
slump-type if hu/hl > 1.6, and is center-type otherwise. Fig. 11 shows a buoyancy-inertia balance (quantified by Fr) for the back flow beha-
our simulation datapoints that belong to different flow regimes in the viors.
plane of Re/Fr and Bn. The slump-type and center-type regimes are
marked by different symbols. A horizontal dashed line at Re / Fr = 60 is 4.3.3. Stable and unstable flow regimes
superimposed on this figure, which approximately separates the two We frequently observe unstable flows in our displacement simula-
regimes. It is interesting to compare the specified value of Re / Fr = 60 tions, forming a variety of flows patterns. Thus, it is useful to study
for displacements in a 2D channel, with the experimental results of unstable flows in more detail and classify them versus the dimension-
Taghavi et al. [53], who found the threshold of Re / Fr = 600 for circular less groups.
pipe displacements, i.e. buoyancy is apparently more effective in the Fig. 14 shows two examples of unstable displacement flows, leading
channel than the pipe. to two different patterns. Concentration colormaps and velocity vectors
are shown. At least in their initial stages, the origin of the two forms of
4.3.2. No-back-flow and temporary-back-flow regimes instabilities observed can be attributed to Kelvin–Helmholtz-like me-
Another type of flow regime that is phenomenologically important chanism. Typically, when instabilities are present in slump-type flows,
is the class of no-back-flow and temporary-back-flow displacement re- sinusoidal shaped wave regions are observed along the interface be-
gimes. In a typical simulation, the leading front always moves forward tween the two fluids. This is demonstrated in Fig. 14a, and the wave
downstream the channel, advecting a large portion of the displaced appear similar in form to internal gravity waves. For center-type flows
fluid outwards. However, the trailing front can move upstream against that are unstable, as demonstrated in Fig. 14b, the instabilities arise at
the direction of the imposed flow, due to buoyancy forces. In our si- two sides of the displacing finger forming symmetric waves. These in-
mulations, we have observed two behaviors for the trailing front: i) the stabilities grow (often into roll waves) and the fluids eventually mix at
trailing front neither moves downstream nor upstream (no-back-flow the interface. For the case shown in Fig. 14b the growths of the inter-
regime); ii) the trailing front advances upstream against the mean flow facial waves become significant at t ≈ 31.
but stops and does not significantly move afterwards, for the rest of the When buoyant and inertial effects are strong, instabilities can grow
simulation time (temporary-back-flow regime). The relevance of back- significantly, resulting in a pattern that we call periodic-detachment. In
flow regimes is that they are frequently associated with interfacial in- this flow pattern, large pieces of the displacing fluid are cut from the
stability. bulk displacing fluid. Using 7 sequential time frames, Fig. 15 details the
detachment process as times grows, for a simulation with Bn = 100,
Re = 300 and Fr = 0.2 (the videos corresponding to these simulations
are included in the supplementary materials). As can be seen, the
number, the size and the mean concentration of the segments change
with time. For instance, by tracking the first segment (from the top to
the bottom image), the size of the separated segment is clearly reduced
over time. Generally, after being cut from the displacing bulk, the
Newtonian core is encapsulated by the surrounding Bingham fluid. The
encapsulated part gradually mixes with the displaced fluid over longer
times and becomes smaller. Since the mean concentration of each
formed segment is a function of time, the mixing degree between the
Fig. 11. Regime classification based on the class of the slump-type ( ) and center-type ( ) two fluids is time-dependent. Furthermore, the number of segments
regimes. The dashed line represents Re/ Fr = 60 . Snapshot images belong to Bn = 20, increases with time, not only due to the formation of new segments near
Re = 500, and Fr = 1000 at t = 8.9 (center-type) and Bn = 50, Re = 500, and Fr = 2 at the bulk of the displacing fluid, but also thanks to flow instabilities
t = 12.7 (slump-type).
which may divide each segment (e.g. see the highlighted boxed-region

88
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 12. Concentration colormaps at t = [0, 8, ...,32]


for a) Bn = 200, Re = 100 and Fr = 0.5 (a case
without a back flow) and b) Bn = 2, Re = 500 and
Fr = 0.5 (a case with a back flow). c) & d)
Spatiotemporal diagrams of the depth-averaged
concentration values for the same simulations as in
a) & b). The size of the domain shown is 1 × 76,
starting from before the gate valve at x = −12 .

shear stress values are symmetric with respect to the channel center
with different signs; the maximum positive/negative shear stresses are
in the vicinity of the lower/upper walls.
Alba et al. [2] experimentally studied the miscible displacement
flows of a viscoplastic fluid by a Newtonian fluid through a long in-
clined pipe. For slump-type flows, they observed break-up of the dis-
placed layer. Our periodic detachment pattern has certain similarities
with the “ripped-type” displacement observed in their experimental
study. For example, in both phenomena, large fluid pieces are cut from
the bulk fluid and the detached pieces appear to decrease in size fol-
lowing the initial break-up. Also the breakage becomes weak when
Fig. 13. Flow regime classification based on appearing back flows. The data corre- buoyancy effects are smaller. The periodic detachment processes (si-
sponding to the temporary-back-flow regime are marked by ( ) and the no-back-flow milar to the break up of viscoplastic layers) leads to form diverse
regime by ( ). Snapshot images belong to Bn = 2, Re = 400, and Fr = 0.5 at t = 7.3 morphologies. However, Alba et al. [2] observed the breakage of the
(temporary-back-flow) and Bn = 50, Re = 500, and Fr = 1 at t = 10.6 (no-back-flow). The
viscoplastic layer (Carbopol gel) in all slump-type displacement, which
broken red arrow indicates the location of trailing front. The domain size shown is
1 × 60. (For interpretation of the references to colour in this figure legend, the reader is is caused by a fast-moving front (and thin layer) of the displacing fluid
referred to the web version of this article.) advancing along the bottom of the pipe. Here the periodic detachment
flow occurs when the Froude number is very small (buoyant and in-
in Fig. 15). Regarding the velocity vectors, they obviously fluctuate ertial effects are strong) wherein the Newtonian fluid is ruptured.
with time and location, both inside and outside of the segments. It Furthermore, the leading front segments are almost separated from one
should also be noted that the variation in the mean spacing length other and the boundaries between the segments are more or less clear,
between the segments can be quantified as the function of Re, Fr and Bn. i.e. the thin layer of displacing fluid is cut here, whereas the thin layers
For example, the mean spacing length has an inverse relation with the in the experiments mentioned are not completely segregated. Generally,
Bingham and Froude numbers whereas it grows with the Reynolds after cutting the pieces of the Newtonian fluid from the bulk, the
number. Newtonian core is encapsulated by the surrounding yield stress fluid.
The concentration, stress and shear stress colormaps are plotted in Despite these differences, which may be largely due to geometry (pipe
Fig. 16a–c for a displacement flow with a periodic detachment pattern. vs channel), we believe the breakage/rupture mechanisms are similar
As can be seen, the concentration and stress vary from the center of to those observed experimentally.
each segment towards the outer surface. Moreover, while the segments The encapsulated parts move at different speeds and we can see in
are more or less separated from one other, the stress at the upper wall is Fig. 16d, both that there is considerable motion within the capsules and
maximum locally. The asymmetric stress field results in asymmetric that the Bingham fluid between the capsules does not become fully
segments. The velocity corresponding (Fig. 16d) to each segment is also unyielded. This, although conceptually similar to the visco-plastic lu-
different. Furthermore, the positive shear stress values are close to the brication flows of Hormozi et al. [22] or the encapsulation studies of
upper/lower walls and the negative values are observed within the Maleki et al. [29], our flows are far from fully developed.
segments. It is interesting to note that after cutting the Newtonian fluid To give a broader understanding of the detached segments, char-
pieces from the bulk, the shear stress locally exceeds the yield stress at acterizing them versus Bn, Re and Fr can be useful. Fig. 17a and b il-
the upper wall, while it does not necessarily do so at the lower wall. lustrate the variation in the mean characteristic diameter of the seg-
Finally, before rupturing and within the pure yield stress fluid, the ments (d ) versus Re and Bn, respectively. It is interesting to note that d

89
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 14. Concentration colormaps and velocity vec-


tors at t = [6.8, 13.85, ...,35] for a) Bn = 5, Re = 200
and Fr = 0.5 and b) Bn = 5, Re = 500 and Fr = 10,
showing two types of unstable displacements. The
domain size shown is 1 × 57, starting from the gate
valve position at x = 0 .

monotonically increases with Re. Initially, d is controlled by a balance simulations show that the number of segments is generally reduced by
between the yield and buoyancy stresses. At larger Re, inertia competes increasing Re whereas their areas are enhanced. Fig. 17b shows that by
with the yield stress to balance buoyancy; thus d increases. In fact, our increasing Bn, d decreases generally, although the effects of Bn are less

Fig. 15. Concentration colormaps and velocity vectors, at


t = [4, 6, 8, 10, 14, 19, 26] for Bn = 100, Re = 300 and Fr = 0.2, showing 7 se-
quential steps in a displacement flow with periodic detachment. The domain
size shown is 1 × 50, starting from the gate valve position at x = 0 .

90
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 16. Colormaps of a) concentration, b) stress (second invariant of


the deviatoric stress) and c) shear stress, as well as d) velocity vectors,
at t = [13] for Bn = 100, Re = 300 and Fr = 0.2, showing a displace-
ment flow with periodic detachment. The domain size shown is
1 × 18, starting from x = 25 .

significant compared to those of Re. tip, semi-detached, and fully-detached fronts. Fig. 19 classifies these
There is also another subtle effect caused by increasing Re, i.e., the front patterns in the plane of Fr and Re/Bn, and gives illustrative ex-
augmentation of the mixing degree between the segments and the amples. In the following subsections, we will review these front beha-
surrounding fluid. As Re increases, the separated segments mix more viors in detail.
with the displaced fluid so that their concentration decreases.
Now we can return to providing the stable/unstable regime classi- 4.4.1. Plug-like front
fication based on the dimensionless groups. Fig. 18 shows the data As mentioned earlier, channel flow displacement simulations in the
corresponding to stable (black hollow) and unstable (red filled) flows, current study cover a wide range of yield stresses, i.e., Bn = [2 − 200].
which are completely segregated in the plane of Fr versus Re/Bn, over a For considerable yield stress values, the displacing front pattern re-
wide range of these parameters. For the stable flows, there is almost a sembles that of a plug flow. Fig. 20a and b display the concentration
sharp transition between purely displacing and displaced fluid regions. colormaps and velocity vectors for two typical cases. As can be seen,
For these flows, the location and concentration of the interface and front plug-like fronts can appear within the flow regimes that are either
are clear. Approximately, there is no wave at the interface between the center-type or slump-type, depending on the buoyancy. Although the
Bingham and Newtonian fluids. However, for unstable cases, there exist displacing front is plug-like due to the large yield stress in both cases,
waves observed in the concentration field and there is no distinct there seems to be a smooth transition between two shear flow regimes,
boundary between the two pure fluids. seen from the velocity vectors. For the slump-type case the mean static
In general, stable displacements are located at higher Fr. The tran- h +h
layer thickness ( u 2 l ≈ 0.2 ) is larger than in the center-type case
sition between stable and unstable displacement flow regimes highly hu + hl
depends on Bn, which may be expected. At fixed Re, by increasing Bn, ( 2
≈ 0.1).
the transition Froude number decreases since instabilities are damped
by the large yield stress. The displacement flows with Fr ≥ 100 are 4.4.2. Inertial tip front
generally stable, regardless of Re and Bn, within the ranges explored. One of the interesting behaviors observed in our numerical simu-
lations is the appearance an inertial tip pattern at the displacing front.
For example, Fig. 21a and b display the concentration colormaps for
4.4. Leading front features two simulations, showing an inertial tip front for Bn = 20,Re = 100,
Fr = 0.5 and Bn = 2,Re = 400,Fr = 1, respectively. The inertial tip at
The behaviors of the leading front are of importance from both the the displacing front starts to form approximately at t = 4 and extends
physical and practical points of view. At least four different leading gradually. It can be seen from these subfigures that the displacing layers
order behaviors can be distinguished for the fronts: plug-like, inertial slump towards the bottom of the channel, while the front itself moves

Fig. 17. The mean characteristic diameter of detached seg-


ments (d ) at Fr = 0.1; a) versus Re with the data corre-
sponding to Bn = 2 ( ), Bn = 5 ( ), Bn = 20 ( ), Bn = 50
( ), Bn = 100 ( ), Bn = 200 ( ); b) versus Bn with the data
corresponding to Re = 50 ( ), Re = 100 ( ), Re = 200 ( ),
Re = 300 ( ), Re = 400 ( ), Re = 500 ( ). d is calculated using
d = 4A , where A is the segment area.
π

91
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 18. Flow regime classification based on stable/unstable


flows. The stable datapoints are marked by (◯) and the unstable
ones by ( ). Two snapshots corresponding to a stable flow
(Bn = 100, Re = 500, and Fr = 1 at t = 32 ) and an unstable flow
(Bn = 5, Re = 50, and Fr = 0.35 at t = 40 ) are included. Within the
unstable displacement flows, the ones with the periodic-detach-
ment patterns are marked by ( ), the semi-detached patterns by
( ) and the sinusoidal shaped wave patterns by ( ).

Fig. 19. Classification of front patterns: plug-like ( ), inertial tip


( ), semi-detached ( ) and fully-detached ( ). Four snapshots
corresponding to a plug front pattern (Bn = 100, Re = 200, and
Fr = 1 at t = 35 ), inertial tip pattern (Bn = 20, Re = 100, and
Fr = 0.35 at t = 28 ), semi-detached pattern (Bn = 20, Re = 300,
and Fr = 0.5 at t = 20 ) and fully-detached front pattern (Bn = 2,
Re = 300, and Fr = 0.35 at t = 16 ) are included.

slightly upward. The velocity vectors show that the inertial effects are formation of a clean inertial tip pattern. Moreover, it can be seen that
quite present at the displacing front, leading to the formation of the the size and the shape of the inertial tip region are enhanced by in-
front inertial tip pattern. A similar type of front pattern has been pre- creasing Re and decreasing Bn.
viously observed in Newtonian displacements [3], although the front is
much more dispersive in Newtonian cases. In our work, the yield stress 4.4.3. Semi-detached and fully-detached fronts
of the displaced fluids limits the dispersivity at the front, leading to the As discussed earlier in detail, an interesting phenomenon observed

Fig. 20. a&b) Concentration colormaps and velocity


vectors at t = [3.5, 7.2, ...,22] for a) Bn = 50, Re = 400
and Fr = 100 and b) Bn = 200, Re = 500 and
Fr = 0.5, showing two examples of the plug-like
front. The domain size shown is 1 × 40, starting
from the gate valve position at x = 0 .

92
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 21. Concentration colormaps and velocity vec-


tors at t = [3, 6.25, ...,16] for: a) Bn = 20, Re = 100
and Fr = 0.5 and b) Bn = 2, Re = 400 and Fr = 1,
showing two examples with an inertial tip pattern.
The domain size shown is 1 × 44, starting from the
gate valve position at x = 0 .

Fig. 22. a) Concentration colormaps and velocity vec-


tors, at t = [0, 2.5, ...,10] for Bn = 100, Re = 300 and
Fr = 0.1, showing a front detachment configuration. The
domain size shown is 1 × 60, starting from x = −25 . b)
Evolution of the leading ( ) and trailing ( ) front velo-
cities for the same simulation.

phenomenon due to flow instabilities is seen. As discussed earlier, the


detachment process periodically continues so that separated segments
are continuously formed. Each detached segment moves faster than the
bulk of the displacing fluid, although each may have a different size and
a different speed. The velocity field is quite unstable and the velocity
profiles around each segment fluctuate with time. Finally, although in
this work we focus mainly on the leading front, we can also notice the
formation of separated segments of the displaced fluid, moving back-
Fig. 23. Simulation results at t = 10, for Bn = 50, Re = 200 and Fr = 0.2, showing a de- ward. These segments are clearly smaller, have smaller speeds and
tachment configuration: a) Concentration colormaps; b) Speed contours
usually mix much more quickly with the surrounding fluid.
V = ( Vx 2 + Vy 2 ) ; c) Vorticity contours (ω = ∂Vy
∂x

∂Vx
∂y ). The domain size shown is Fig. 22b illustrates the leading and trailing front velocities, for the
1 × 66, starting from x = −25 . same simulation as in Fig. 22a. Since the fronts are periodically formed,
these velocities correspond to fastest advancing segments of the dis-
in some of our simulations is that pieces of the displacing front are cut placing and displaced fluids. Initially, the leading front velocity rapidly
from the rest of displacing finger, forming separated segments of the increases from 0, and after a peak, starts to slowly decrease (perhaps
heavy fluid advected within the displaced fluid. The detachment pro- due to partial mixing). This implies that the speed of the most advanced
cess nearly always starts at the front. For example, Fig. 22a shows the piece of the displacing fluid (within the displaced one) remains sig-
concentration colormaps and velocity vectors, at different times for nificant. On the other hand, the absolute value of the trailing front
Bn = 100,Re = 300 and Fr = 0.1, in which the front detachment velocity increases with time, but slows down afterwards. The absolute
trailing front velocity decreases (due to very strong mixing) until it

93
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Fig. 24. Effects of the dimensionless groups on the


leading front velocity for at a&c) Re = 100 and b&d)
Re = 300 . In the top row the data correspond to Fr = 0.5
( ), Fr = 1 ( ), Fr = 2 ( ), Fr = 10 ( ), Fr = 100 ( ),
Fr = 1000 ( ) and in the bottom row to Bn = 2 ( ),
Bn = 5 ( ), ( )Bn = 20, ( ) Bn = 50, Bn = 100 ( ),
Bn = 200 ( ).

Fig. 25. Variation of Vf versus Re for a) Fr = 0.5 ; b) Fr = 1; c) Fr = 10 ; and d) Fr = 100 . The data correspond to Bn = 20 ( ), Bn = 50 ( ), Bn = 100 ( ), Bn = 200 ( ).

both a Poiseuille flow and a plug flow. High speed regions are observed
within the separated segments. Fig. 23c shows the vorticity contours.
Positive vorticity values are within the separated segments and negative
vorticity values are outside. However, before and after each segment,
small islands of positive vorticity values can be also observed close to
the upper/lower walls. Therefore, temporarily flow reversals do occur
in the vicinity of the upper/lower walls, due to a mix of buoyancy and
interfacial instabilities.

4.4.4. Leading front velocity


It is interesting to investigate the leading front velocity as a function
of the dimensionless numbers of the flow. In our large aspect ratio
channel, the leading front velocity can provide an indication to how
efficiency the displacement process is. Generally, one may expect that
when the leading front velocity approaches unity, the displacement
efficiency tends to 100%. Details about methods of the calculation of Vf
Fig. 26. Effects of Re, Bn and Fr on Vf, the values of which are marked by the symbol size
and colors. can found in [3,54].
Let us start with analyzing the effects of the dimensionless numbers
on the leading front velocity. Fig. 24a and b show the variation of Vf
reaches zero at very long times (not shown).
versus Bn (for different Fr) at two fixed Re = 100 and Re = 300, re-
In order to provide further understanding about how the front is
spectively. Note that for some cases at very small Fr, the front velocity
detached, let us look into Fig. 23 showing the contours of concentra-
highly fluctuates with time and cannot reach a plateau value; these
tion, velocity and vorticity, at t = 10 for Bn = 50,Re = 200 and Fr = 0.2 .
datapoints are not included in the figure. As can be seen, Vf decreases
Fig. 23a shows that the separated segments of the displacing fluid have
by increasing Bn (in other words by increasing the yield stress). The
progressively smaller areas along the channel length. Fig. 23b illus-
explanation for this behavior is that by increasing Bn the thickness of
trates the absolute speed contours. The flow velocity profile is far from
the static residual layers of the displaced fluid generally decreases,

94
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

Table 2
Summary of the main observations.

Regime class Ref. figure Appropriate dimensionless group Comments

Center- or slump-type Fig. 11 Re/Fr • InImportant


center-type: h ≈ h & in slump-type: h /h ≫ 1
• Transition threshold:
u l u l
h /h ≈ 1.6
• Transition from
u l
slump- to center-type: Re/Fr ≈ 60
• For Binghamindependent of Bn & V
• Temporary-back-flows observed at likely
0
Temporary- or no-back-flow Fig. 18 Re/Bn & Fr fluids, no-back-flows occur
• At smaller Fr, by increasing Bn, no-back-flows
small Fr
• Transition with small dependency on Bn. increasingly observed
• Stable (unstable) flows with (no) clear boundary between phases
Stable or unstable Fig. 13 Re/Bn & Fr • Unstable (stable) flows with (no) wave at interface
• Transition with significant dependency on Bn
• Main front patterns: plug-like, inertial tip, semi-detached, fully-detached
Front patterns Fig. 19 Re/Bn & Fr • For Bingham fluid, plug-like flows likely occur
• At large Fr front pattern independent of Re/Bn

which therefore implies that Vf needs to decrease as well. The other locations where uniform static layers appear have been quantified.
effect that can be seen is that Vf generally decreases with Fr. The in- Next, the main displacement regimes have been introduced, in the three
terpretation of this behavior is that, for fixed Re (i.e., constant ratio of classes of slump-type/center-type, no-back-flow/temporary-back-flow
inertial to viscous forces), Vf decreases when buoyancy forces decrease and stable/unstable displacements. Moreover, the characteristics of the
(as Fr increase). As Fr decreases, these forces act as additional driving leading front have been looked into in detail. For example, various front
forces (in addition to the imposed flow), improving Vf and worsening patterns observed, e.g., plug-like, inertial tip, semi-detached and fully-
the displacement efficiency. Fig. 24c and d show the variation of Vf detached fronts, have been studied. Finally, the leading front velocity
versus Fr (for different Bn) at two fixed Re = 100 and Re = 300 . For has been quantified as a function of the dimensionless numbers. A
fixed Bn, it can be seen that at small Fr, Vf highly depends on Fr, since summary of the main observations is presented in Table 2.
buoyancy forces are significant. Initially, Vf sharply decreases with Fr
but becomes independent of it at very large Fr values, where the dis- Acknowledgements
placement flow approaches an isodensity, symmetric three-layer dis-
placement flow. This research has been carried out at Laval University, supported
Fig. 25 shows the effects of Re on Vf for Bn and Fr. It is interesting to partially by the Canada Foundation for Innovation (Grant no.s
note that Re has opposite effects on Vf, at small and large Fr: by in- GF112622, GQ113034 and GF517657) and the Discovery Grant of the
creasing Re the leading front velocity increases at small Fr, but the Natural Sciences and Engineering Research Council of Canada (Grant
opposite is true at large Fr. no. CG10915). The research has been enabled in part by support pro-
At fixed, large value of Fr, as the displacing finger is more towards vided by Calcul Qubec. We wish to also acknowledge the help of Ali
the channel center, the upper/lower static layers are affected by Re to Roustaei in using PELICANS.
the extent that increasing Re leads to decreasing the static layer
thickness (due to inertial dissipation). More rigorously, Supplementary material
Wielage–Burchard & Frigaard [61] have shown that, for Fr → ∞, in-
creasing Re improves the rate of energy production of the steady flow, This supplementary section includes videos showing the con-
resulting in decreasing the residual wall layer thicknesses. Since the centration colormap, the velocity vectors, the absolute speed contours
static layer effects are dominant at larger Fr, Vf decreases with Re. On and the vorticity contours in a displacement flow in a periodic-de-
the other hand, at small Fr, the displacing finger is more towards tachment regime. The numerical simulation is run for Bn = 100,
channel center (i.e., the flow approaches a two-layer displacement). In Re = 300 and Fr = 0.2 . The domain size shown is 1 × 50, starting from
this case, for fixed Bn and Fr, increasing Re may be interpreted as in- the gate valve position at x = 0 .
creasing buoyancy, which is a significant driving force at small Fr. Thus, Supplementary material associated with this article can be found, in
since buoyant effects dominate the flow, Vf increases with Re. Fig. 26 the online version, at 10.1016/j.jnnfm.2017.10.001.
presents the variation of Vf, as a function of Fr, Bn and Re, where Vf
values are marked by the symbol size and colors. It is seen that in- References
creasing Bn generally results in decreasing Vf, and that the variation of
Vf is not monotonic versus Re. Moreover, the variation of Vf with Fr is [1] K. Alba, I. Frigaard, Dynamics of the removal of viscoplastic fluids from inclined
negligible for Fr ≥ 5. pipes, J. Non Newt. Fluid Mech. 229 (2016) 43–58.
[2] K. Alba, S. Taghavi, J. de Bruyn, I. Frigaard, Incomplete fluid-fluid displacement of
yield-stress fluids. part 2: highly inclined pipes, J. Non Newt. Fluid Mech. 201
(2013) 80–93.
5. Summary [3] K. Alba, S. Taghavi, I. Frigaard, Miscible heavy-light displacement flows in an in-
clined two-dimensional channel: a numerical approach, Phys. Fluids 26 (12) (2014)
Using numerical simulations, we have considered displacement 122104.
[4] M. Allouche, I. Frigaard, G. Sona, Static wall layers in the displacement of two
flows of a Bingham fluid by a Newtonian fluid, along a 2D uniform visco-plastic fluids in a plane channel, J. Fluid Mech. 424 (1) (2000) 243–277.
plane channel wherein the heavier fluid pushes the lighter fluid. The [5] A. Amiri, F. Larachi, S. Taghavi, Buoyant miscible displacement flows in vertical
long-time behaviors of the displacement flow have been characterized pipe, Phys. Fluids 28 (10) (2016) 102105.
[6] S. Bittleston, J. Ferguson, I. Frigaard, Mud removal and cement placement during
in the terms of three important dimensionless numbers: the Reynolds primary cementing of an oil well–laminar non-newtonian displacements in an ec-
number (Re), the Bingham number (Bn) and the densimetric Froude centric annular hele-shaw cell, J. Eng. Math. 43 (2–4) (2002) 229–253.
number (Fr). Details associated with various displacement flow beha- [7] R. Brandao, J. Fontana, J. Miranda, Interfacial pattern formation in confined power-
law fluids, Phys. Rev. E 90 (1) (2014) 013013.
viors have been uncovered. First, static residual wall layers of the dis- [8] D. Burfoot, K. Middleton, J. Holah, Removal of biofilms and stubborn soil by
placed fluid have been discussed. In particular, the upper/lower pressure washing, Trends Food Sci. Technol. 20 (2009) S45–S47.
thicknesses of the residual wall layers as well as the upper/lower [9] C. Chen, E. Meiburg, Miscible displacements in capillary tubes. part 2. numerical

95
A. Eslami et al. Journal of Non-Newtonian Fluid Mechanics 249 (2017) 79–96

simulations, J. Fluid Mech. 326 (1996) 57–90. viscoplastic material inside a tube or between two parallel disks: conditions for wall
[10] Y. Cinar, A. Riaz, H. Tchelepi, et al., Experimental study of CO2 injection into saline detachment of the advancing front, J. Rheol. 53 (5) (2009) 1155–1191.
formations, SPE ATCE, (2007). [40] M. Parvez, N. Ong, Y. Lam, S. Tor, Gas-assisted injection molding: the effects of
[11] N. Cogan, J. Keener, Channel formation in gels, SIAM J. Appl. Math. 65 (6) (2005) process variables and gas channel geometry, J. Mater. Process. Technol. 121 (1)
1839–1854. (2002) 27–35.
[12] P. Coussot, Saffman–Taylor instability in yield-stress fluids, J. Fluid Mech. 380 [41] H. Pascal, Dynamics of moving interface in porous media for power law fluids with
(1999) 363–376. yield stress, Int. J. Eng. Sci. 22 (5) (1984) 577–590.
[13] Y. Dimakopoulos, J. Tsamopoulos, Transient displacement of a viscoplastic material [42] H. Pascal, Rheological behaviour effect of non-newtonian fluids on dynamic of
by air in straight and suddenly constricted tubes, J. Non Newt. Fluid Mech. 112 (1) moving interface in porous media, Int. J. Eng. Sci. 22 (3) (1984) 227–241.
(2003) 43–75. [43] H. Pascal, Rheological effects of non-newtonian behavior of displacing fluids on
[14] A. Eslami, S. Taghavi, Viscous fingering regimes in elasto-visco-plastic fluids, J. Non stability of a moving interface in radial oil displacement mechanism in porous
Newt. Fluid Mech. 243 (2017) 79–94. media, Int. J. Eng. Sci. 24 (9) (1986) 1465–1476.
[15] J. Fontana, E. Dias, J. Miranda, Controlling and minimizing fingering instabilities in [44] P. Petitjeans, T. Maxworthy, Miscible displacements in capillary tubes. part 1. ex-
non-newtonian fluids, Phys. Rev. E 89 (1) (2014) 013016. periments, J. Fluid Mech. 326 (1996) 37–56.
[16] J. Freitas, E. Soares, R. Thompson, Viscoplastic-viscoplastic displacement in a plane [45] A. Poslinski, P. Oehler, V. Stokes, Isothermal gas-assisted displacement of visco-
channel with interfacial tension effects, Chem. Eng. Sc. 91 (2013) 54–64. plastic liquids in tubes, Polym. Eng. Sci. 35 (11) (1995) 877–892.
[17] P. Fryer, G. Christian, W. Liu, How hygiene happens: physics and chemistry of [46] G. Rousseaux, A. De Wit, M. Martin, Viscous fingering in packed chromatographic
cleaning, Int. J. Dairy Technol. 59 (2) (2006) 76–84. columns: linear stability analysis, J. Chromatogr. A 1149 (2) (2007) 254–273.
[18] C. Gabard, J. Hulin, Miscible displacement of non-newtonian fluids in a vertical [47] A. Roustaei, Yield Stress Fluid Flows in Uneven Geometries: Applications to the Oil
tube, Eur. Phys. J. E. 11 (3) (2003) 231–241. and Gas Industry, Ph.D. thesis, University of British Columbia, 2016.
[19] R. Glowinski, P. Le Tallec, Augmented lagrangian and operator-splitting methods in [48] K. Sahu, H. Ding, P. Valluri, O. Matar, Pressure-driven miscible two-fluid channel
nonlinear mechanics, Phys. Fluids 19 (2007) 084106. flow with density gradients, Phys. Fluids 21 (4) (2009) 043603.
[20] S. Hill, et al., Channeling in packed columns, Chem. Eng. Sci. 1 (6) (1952) 247–253. [49] T. Séon, J. Hulin, D. Salin, B. Perrin, E. Hinch, Buoyancy driven miscible front
[21] K. Holloway, P. Habdas, N. Semsarillar, K. Burfitt, J. de Bruyn, Spreading and dynamics in tilted tubes, Phys. Fluids 17 (3) (2005) 031702.
fingering in spin coating, Phys. Rev. E 75 (4, 2) (2007) 046308. [50] D. de Sousa, E. Soares, R. de Queiroz, R. Thompson, Numerical investigation on gas-
[22] S. Hormozi, K. Wielage-Burchard, I. Frigaard, Entry, start up and stability effects in displacement of a shear-thinning liquid and a visco-plastic material in capillary
visco-plastically lubricated pipe flows, J. Fluid Mech. 673 (2011) 432–467. tubes, J. Non Newt. Fluid Mech. 144 (2) (2007) 149–159.
[23] P. Howell, S. Waters, J. Grotberg, The propagation of a liquid bolus along a liquid- [51] P.A. Swain, G. Karapetsas, O. Matar, K. Sahu, Numerical simulation of pressure-
lined flexible tube, J. Fluid Mech. 406 (2000) 309–335. driven displacement of a viscoplastic material by a newtonian fluid using the lattice
[24] D. Huh, H. Fujioka, N. Tung Y.C.and Futai, R. Paine, J. Grotberg, S. Takayama, boltzmann method, Eur. J. Mech. B Fluid 49 (2015) 197–207.
Acoustically detectable cellular-level lung injury induced by fluid mechanical [52] S. Taghavi, K. Alba, I. Frigaard, Buoyant miscible displacement flows at moderate
stresses in microfluidic airway systems, Proc. Natl. Acad. Sci. 104 (48) (2007) viscosity ratios and low atwood numbers in near-horizontal ducts, Chem. Eng. Sc.
18886–18891. 69 (2012) 404–418.
[25] P. Huzyak, K. Koelling, The penetration of a long bubble through a viscoelastic fluid [53] S. Taghavi, K. Alba, M. Moyers-Gonzalez, I. Frigaard, Incomplete fluid-fluid dis-
in a tube, J. non-Newtonian Fluid Mech. 71 (1) (1997) 73–88. placement of yield stress fluids in near-horizontal pipes: experiments and theory, J.
[26] L. Lake, Enhanced Oil Recovery, Prentice Hall, 1989. Non Newt. Fluid Mech. 167–168 (2012) 59–74.
[27] A. Lindner, P. Coussot, D. Bonn, Viscous fingering in a yield stress fluid, Phys. Rev. [54] S. Taghavi, K. Alba, T. Séon, K. Wielage-Burchard, D. Martinez, I. Frigaard, Miscible
Lett. 85 (2) (2000) 314. displacement flows in near-horizontal ducts at low atwood number, J. Fluid Mech.
[28] A. Lindner, P. Coussot, D. Bonn, Viscous fingering in a gel, Branching in Nature, 696 (2012) 175–214.
Springer, 2001, pp. 433–438. [55] S. Taghavi, T. Seon, D. Martinez, I. Frigaard, Buoyancy-dominated displacement
[29] A. Maleki, S. Hormozi, A. Roustaei, I. Frigaard, Macro-size drop encapsulation, flows in near-horizontal channels: the viscous limit, J. Fluid Mech. 639 (2009)
J. Fluid Mech. 769 (2015) 482–521. 1–35.
[30] N. Maleki-Jirsaraei, A. Lindner, S. Rouhani, D. Bonn, Saffman–Taylor instability in [56] S. Taghavi, T. Séon, D. Martinez, I. Frigaard, Influence of an imposed flow on the
yield stress fluids, J. Phys. Condens. Matter 17 (14) (2005) S1219. stability of a gravity current in a near horizontal duct, Phys. Fluids 22 (3) (2010)
[31] S. Malhotra, M.M. Sharma, Impact of fluid elasticity on miscible viscous fingering, 031702.
Chem. Eng. Sci. 117 (2014) 125–135. [57] S. Taghavi, T. Séon, K. Wielage-Burchard, D. Martinez, I. Frigaard, Stationary re-
[32] K. McCloud, J. Maher, Experimental perturbations to Saffman–Taylor flow, Phys. sidual layers in buoyant newtonian displacement flows, Phys. Fluids 23 (4) (2011)
Rep. 260 (3) (1995) 139–185. 044105.
[33] G. Moisés, M. Naccache, K. Alba, I. Frigaard, Isodense displacement flow of visco- [58] G. Taylor, Cavitation of a viscous fluid in narrow passages, J. Fluid Mech. 16 (04)
plastic fluids along a pipe, J. Non Newt. Fluid Mech. 236 (2016) 91–103. (1963) 595–619.
[34] R. Mollaabbasi, S. Taghavi, Buoyant displacement flows in slightly non-uniform [59] G. Vinay, Modélisation du Redémarrage des écoulements de Bruts Paraffiniques
channels, J. Fluid Mech. 795 (2016) 876–913. Dans les Conduites Pétrolieres, Ph.D. thesis, École Nationale Supérieure des Mines
[35] S. Mora, M. Manna, Saffman–taylor instability for generalized newtonian fluids, de Paris, 2005.
Phys. Rev. E 80 (1) (2009) 016308. [60] A. White, T. Ward, Constant pressure gas-driven displacement of a shear-thinning
[36] J. Nase, D. Derks, A. Lindner, Dynamic evolution of fingering patterns in a lifted liquid in a partially filled radial hele-shaw cell: thin films, bursting and instability,
hele-shaw cell, Phys. Fluids 23 (12) (2011) 123101. J. Non Newt. Fluid Mech. 206 (2014) 18–28.
[37] E. Nelson, D. Guillot, Well Cementing, second ed., Schlumberger Educational [61] K. Wielage-Burchard, I. Frigaard, Static wall layers in plane channel displacement
Services, 2006. flows, J. Non Newt. Fluid Mech. 166 (5) (2011) 245–261.
[38] F. Orr, J. Taber, Use of carbon dioxide in enhanced oil recovery, Science 224 (4649) [62] J. Wiklund, M. Stading, C. Trägårdh, Monitoring liquid displacement of model and
(1984) 563–569. industrial fluids in pipes by in-line ultrasonic rheometry, J. Food Eng. 99 (3) (2010)
[39] G. Papaioannou, J. Karapetsas, Y. Dimakopoulos, J. Tsamopoulos, Injection of a 330–337.

96

Вам также может понравиться