You are on page 1of 14

Powder Technology 395 (2022) 604–617

Contents lists available at ScienceDirect

Powder Technology

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

Numerical investigation of non-uniform sand retention behavior


in sand screens
Noor Ilyana Ismail a, Shibo Kuang a,⁎, Mengmeng Zhou a, Aibing Yu a,b
a
ARC Research Hub for Computational Particle Technology, Department of Chemical Engineering, Monash University, Clayton, Victoria 3800, Australia
b
Center for Energy and Environment, Southeast University-Monash University Joint Research Institute, Suzhou 215123, China

a r t i c l e i n f o a b s t r a c t

Article history: Non-uniform sand retention behavior often occurs to the serviced screen deteriorating erosion. However, this
Received 9 August 2021 phenomenon is poorly understood. This paper presents a numerical study of the sand retention on wire-
Received in revised form 11 October 2021 wrapped screens, with special reference to non-uniform behaviors. This is done by the combined approach of
Accepted 13 October 2021
computational fluid dynamics (CFD) and discrete element method (DEM). The validity of the model has been val-
Available online 16 October 2021
idated for dry and wet sand screen systems. It is used here to study sand retention behaviors at different solid
Keywords:
concentrations and particle size distributions (PSD). Via this model, five distinct sand retention modes are iden-
Sand screen tified: No sand retention (Mode I), partial sand retention (Mode II), sand retention with slow sequential bridging
Sand retention (Mode III), sand retention with fast sequential bridging (Model IV) and sand retention with instantaneous bridg-
Retention mode ing (Mode V). Modes II and III belong to non-uniform sand retention, which develops strong local flows that in-
Non-uniform duce local erosion or hot spot on the screen. A phase diagram is introduced to predict these five modes and their
CFD-DEM transition with respect to solid concentration and PSD. Additionally, the predicted flow and force structures are
Converging slot analyzed in detail. The results indicate that the bridging over a slot heavily relies on the particle accumulation on
the screen. A new screen with a converging slot configuration is proposed to improve this particle accumulation.
This improvement helps develop uniform sand retention on the screen.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction occur through different processes such as size exclusion and particle
bridging. Sand retention by size exclusion occurs when particles are
Sand production is a migration of failed sand grains from a reservoir larger than the slot opening of a screen. Thus, such particles are trapped
into a well during hydrocarbon extraction. It has been one of the major above the screen. In contrast, particle bridging occurs when particles are
problems that operators need to deal with in developing unconsoli- smaller than the slot opening. It is affected by many variables related to
dated reservoirs, which represent about 70% of the oil and gas reserves the screen opening to particle diameter ratio, particle size distribution,
around the world [1]. Sand, along with the hydrocarbon, causes erosion fluid velocity, sand concentration, screen type, fluid properties, etc.
in piping systems, accumulation of sand in wells and surface vessels, This complicates the screen design very much. To date, our knowledge
and disposal problems. Therefore, sand production is highly undesir- underpinning sand screen applications is not established [2].
able, promoting the development of various control methods. One of For example, non-uniform sand retention behavior is commonly en-
the most commonly used methods is installing sand screens in a well countered during the sand retention process. It occurs when a screen is
to function as a downhole filter to exclude sand while allowing hydro- partially covered by sand, leaving only some sections of the screen open
carbon to be passed. With sand screen, good well productivity should for flow. This condition produces highly localized fluid flows in the
be maintained with small operational complexity and low cost. One of screen as fluid is forced to produce through the remaining open sec-
the criteria that are used to evaluate screen performance is sand reten- tions. As this problem persists for a certain period, a screen could be par-
tion. In practice, a screen is expected to achieve maximum sand reten- tially plugged in the area covered by particles. Meanwhile, the open
tion while maintaining an excellent passing rate for hydrocarbon. section suffers severe erosion by the high-velocity particle-fluid flow.
However, sand retention of the screen is a complex process. It can Such phenomena have been experimentally observed by Hamid et al.
[3]. Generally, the resulting erosion is very rapid and could occur in
the order of few minutes. Similar problems have also been reported
by other investigators [4–6], who studied screen failure mechanisms
⁎ Corresponding author. and reported that random sand retention or plugging of the screen
E-mail address: shibo.kuang@monash.edu (S. Kuang).

https://doi.org/10.1016/j.powtec.2021.10.016
0032-5910/© 2021 Elsevier B.V. All rights reserved.
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

leading to very high erosion rates in other sections, inducing local ero- some of its features to some degree. However, the non-uniform sand re-
sion or hot spot on the screen. In practice, this non-uniform sand reten- tention phenomenon observed in serviced screens has not been re-
tion reduces hydrocarbon production rates and causes the industry ported in the previous CFD-DEM studies. To date, it is unclear whether
millions of dollars each year to protect and repair screens. Although crit- the CFD-DEM approach is capable of predicting this challenging phe-
ical to screen performance, non-uniform sand retention is rarely studied nomenon.
due to the lack of a viable method. In fact, at present, the design and con- In the current work, the sand retention of the wire-wrapped screen
trol of sand screens rely heavily on experimental methods. To cover var- is studied by the reported CFD-DEM model [19]. It focuses on non-
ious variables related to screen geometries, operating conditions, and uniform sand retention behavior. This phenomenon is observed when
fluid and sand properties, extensive experiments are needed to be con- sand retention occurs randomly and only on some sections of the screen
ducted, which is time-consuming and labor-intensive, causing extra leaving other sections open for flow. On the other hand, uniform sand
costs. Furthermore, measuring the sand retention and production retention is observed when the entire screen section is fully covered
under controlled and reproducible flow conditions is currently difficult by the retained sand particles. Even though non-uniform sand retention
to achieve [7]. Also, at this stage of development, the particle and flow is frequently observed, the conditions under which it occurs have re-
structures that govern sand screen performance cannot be assessed mained unknown. In addition, in the past, there have been no studies
using experimental methods. All these problems bring difficulties to un- reported to analyze the non-uniform sand retention in detail due to
derstanding and quantifying the sand retention behavior experimen- lack of effective tool and, thus, the knowledge regarding its flow behav-
tally. This is particularly true for non-uniform sand retention as it iors through a sand screen is limited. A better understanding of the flow
occurs more locally and thus is more complicated. behaviors of this process would help the design and optimization of
Computer models provide a promising alternative because of the sand screens. Therefore, the aim of this study is first to investigate the
merits of high efficiency, low cost, and zero risks while providing direct influence of operational conditions such as solid concentration and par-
mechanistic insights into particle-fluid flow systems. Among the most ticle size distribution on the sand retention over screen. Then, a phase
prominent models is the combined approach computational fluid dy- diagram is established to identify the conditions under which different
namics (CFD) for the fluid phase and discrete element method (DEM) sand retention modes, including non-uniform ones, occur. Furthermore,
for particles. In such an approach, the trajectories of and the forces act- the simulation results are analyzed in detail in terms of particle patterns
ing on individual particles are tracked directly. It offers a sound theoret- as well as flow and force structures to understand the nature of non-
ical base for describing the interactions among particles, screen wall, uniform sand retention behaviors. Finally, a new screen is proposed to
and surrounding fluid. These interactions determine screen perfor- mitigate the non-uniform sand retention based on the understanding
mance but cannot be generally modeled by continuum methods such developed.
as two-fluid models (TFM), especially for particle-particle/wall interac-
tions [8,9]. For this reason, the TFM approach has not been applied to 2. Mathematical model
simulate sand screens. As such, to our knowledge, the CFD-DEM ap-
proach is only the viably available numerical method to simulate sand The current CFD-DEM model used here is the same as the one re-
screen. cently reported elsewhere [19]. Therefore, only the key features of the
In recent years, the CFD-DEM approach has been widely accepted as model are outlined below for brevity. However, detailed modeling and
an effective tool to study various particle-fluid systems, as reviewed by numerical treatments can be found elsewhere [19,21–35].
different investigators [10–15]. Despite the broad applications, only a
few efforts have been reported to use the CFD-DEM method to study 2.1. Particle motion
sand retention. One of the earliest CFD-DEM studies on sand retention
over screen was by Feng et al. [16]. They conducted a parametric In the DEM model, the motions of individual particles are traced
study to elucidate the effects of solid concentration and liquid flow by an explicit numerical solution of Newton's equation of motion.
rate on monosized sand retention of a wire-wrapped screen. Thereafter, For particle i, its translational and rotational motions are determined
several CFD-DEM studies have focused on simulating sand retention by:
over screen using a polydisperse particles system. For example, Wu
ki  
et al. [17] used polydisperse particles to investigate the effects of solid dvi
mi ¼ f d,i þ f ∇p,i þ f ∇:τ,i þ ∑ f cn,ij þ f ct,ij þ f dn,ij þ f dt,ij þ mi g ð1Þ
concentration, fluid velocity and screen opening on sand retention of a dt j¼1
wire-wrapped screen. Shaffee et al. [18] used CFD-DEM simulations to
investigate the effects of adhesive polydisperse particles on the sand re- and
tention process of a wire-wrapped screen for enhanced sand control. Al-
ki  
though useful, these studies focused on examining screen efficiency, dωi
Ii ¼ ∑ Tij ð2Þ
and the effects of key variables were not studied in detail. Besides, dt j¼1

none of the previous CFD-DEM studies have attempted to analyze


particle-fluid flow characteristics and force structures to understand where mi is the particle mass, Ii is the moment of inertia, vi is the particle
the screen performance and identify the underlying mechanisms. Re- translational velocity, ωi is the particle angular velocity, and ki is the
cently, the authors [19] conducted CFD-DEM studies to investigate the number of particle/wall in contact with particle i. The forces involved
effects of slot width-particle size ratios and wetting fluids on sand re- are: the particle-fluid interaction forces, including the drag force, fd,i,
tention and production behaviors of widely used wire-wrapped screens. calculated by the model formulated using sub-particle simulation re-
The analysis of particle flow patterns and force structures, revealed that sults [26,31,36], the pressure gradient force, f∇p,i and the viscous force,
sand retention could occur through three distinct mechanisms: stable f∇.τ,i, the gravitational force, mi g and the inter-particle forces between
bridging, bridging with intermittent collapse and bridging with contin- particles which include elastic force, fc,ij and viscous damping force, fd,
uous collapse. Razavi et al. [20] also studied sand retention mechanisms ij. These inter-particle forces are solved into the normal and tangential

of screens via CFD-DEM simulations but with and without porous media components at a contact point. Tij is the torque acting on particle i by
on top of the slot opening. They reported that the presence of a porous particle j, which causes particle i to rotate. The particle motions are
region on top of the slot reduced the number of particles passing calculated based on the soft-particle approach originally developed by
through the slot in the specified time compared with the case without Cundall and Strack [37]. The inter-particle contact force is determined
the porous region. All these preceding CFD-DEM studies have shown according to the Hertz collision theory [10,38,39] and the Coulomb
the viability of this method in modeling sand screen and predicting sliding friction model is used to determine the frictional force. For

605
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Table 1 cell into 48 elements to calculate the initial values of porosity in the
Components of forces and torque acting on particle i. CFD cell, which is expressed as
Forces and torques Symbols Equations
n
pffiffiffiffiffi 3=2 ∑i¼1 V i
Normal elastic force fcn,ij − 43 E∗ R∗ δn,ij ni εf ¼ 1− ð5Þ
qffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffi V cell
Normal damping force fdn,ij 
−2 56 η Sn,ij m∗ vr,ij ∙nij ni
∗ pffiffiffiffiffiffiffiffiffiffiffiffi


where Vcell and Vi are the volume of a computational cell and the volume
Tangential elastic force fct,ij −8G R δn:ij ti
qffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffi  
Tangential damping fdt,ij −2 56 η St,ij m∗ ðvr ∙ti Þti þ ωi  ri −ωj  rj of particle i inside this cell, respectively, and n is the number of particles
force in the cell. In the second step, the diffusion-based averaging method
Sliding friction force fct,ij −μs|fcn,ij|ti
[45] is used to determine the porosity in the computational domain
Drag force fd,i 2−βðε,〈 ;Re ;〉Þ
8 C DO πρf α ðdi , ;xi , ;εÞju−vi jðu−vi Þεf
1
using the initial values given by Eq. (6). Correspondingly, a transient,
Pressure gradient force f∇p,i − ∇ p ∙ Vp,i homogeneous diffusion equation is solved over the pseudo-time
Viscous force f∇∙τ,i −(∇ ∙ τ)Vp,i
period,τp. This porosity calculation method is applicable to different
where: mesh sizes and easy for parallelism. In addition, the volumetric

      ð1:5− log 〈 Re 〉Þ
2

β εf , 〈 Re 〉 ¼ 2:65 εf þ 1 − 5:3−3:5εf ε2f exp − 2


particle-fluid interaction force in the considered cell is calculated by
  0:5ð1þεf Þ  
α di , xi , ε f ¼ xi þ 0:5yi þ 0:5 1−εf y2i 1 n  

i yi
2 Fpf ¼ ∑ f ð6Þ
" #−1 V cell i¼1 d,i
di εf ju−vi jhdi
hdi ¼ ∑xi =di , yi ¼ hdi
, h Re i ¼ v
i
  Note that corresponding to model A, the calculation of the volumet-
1−σ 2j
ð1−σ 2i Þ ln ðeÞ ric particle-fluid force needs to exclude the pressure gradient and vis-
1
E∗ ¼ Ei þ Ej , 1
R∗ ¼ R1i þ R1j , 1
m∗ ¼ m1i þ m1j , η ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2

ln ðeÞþπ 2

2ð2þσ j Þð1−σ j Þ pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi cous forces. This is not the case for Model B, as detailed elsewhere [24].
1
G∗ ¼ 2ð2þσ iEÞið1−σ i Þ þ Ej , Sn,ij ¼ 2E∗ R∗ δn,ij , St,ij ¼ 28 R∗ δt,ij

2.4. Coupling scheme and numerical solution

convenience, the particle-particle/wall and particle-fluid force and The modeling of particle flow by the DEM is at an individual particle
torque models are listed in Table 1. level, while the fluid flow by the CFD model is at a computational cell
level. According to the work of Xu and Yu [21], their coupling is numer-
2.2. Fluid flow ically achieved as follows. At each time step, the DEM will provide infor-
mation, such as the positions and velocities of individual particles for
In the CFD model, the liquid flow is modeled by the continuum fluid the evaluation of porosity and volumetric particle-fluid force in a com-
method, the continuity of mass and momentum based on local mean putational cell. The CFD model will then use these data to determine
variables over a computational cell. The model formulation has two dif- the fluid flow field, which yields the particle-fluid forces acting on indi-
ferent forms due to different treatments of the pressure [24,27]. If the vidual particles. Incorporating the resulting forces into the DEM will
pressure is attributed to the fluid phase only, it is referred to as Model produce information about the motion of individual particles for the
B. If both the fluid and solid phases share the pressure, it is called next time step. The particle-fluid forces acting on individual particles
Model A. This difference corresponds to whether the void fraction is in- will react on the fluid phase so that Newton's third law of motion is sat-
cluded in the pressure gradient term of Navier-Stokes equations. Model isfied.
A is used in the current study. The fluid equations in (3) and (4) are solved with the finite volume
  method. Convection and diffusion terms are discretized second-order
∂ ρf εf   central schemes, and time integrations are based on a second-order im-
þ ∇∙ ρf εf uf ¼ 0 ð3Þ plicit scheme. A PISO (Pressure Implicit Splitting Operation) algorithm
∂t
is used for velocity–pressure coupling [46]. The equations for a particle
and are explicitly integrated. The DEM time step selection refers to the Ray-
  leigh time ΔtR:
∂ ρf εf uf     sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
þ ∇∙ ρf εf uf uf ¼ −εf ∇p þ Fpf þ ∇∙ εf τ þ ρf εf g ð4Þ πr i 2ρp ð1 þ σ i Þ
∂t TR ¼ ð7Þ
ð0:1631σ i þ 0:8766Þ Yi
where uf,ρf and p are, respectively, the fluid velocity, density and
pressure; τ,εf and Fpf are the fluid viscous stress tensor, porosity and where ri,pp,Yi and σi are respectively, the radius, density, Young
the volumetric particle-fluid interaction forces, respectively. Note that modulus and Poisson's ratio of particle i. To prevent numerical
fluid flow velocities through screens are expected low [40]. To simulate instabilities and nonphysical results, the ΔtDEMis often set to be lower
field conditions, the fluid velocity is set to 0.005 m/s, as done in typical than TR. Based on the smallest particle diameter of particle size
experiments simulating sand retention [41–44]. Under this flow condi- distribution and particles properties used in the simulation, the TR is
tion, the Reynold number is smaller than 10; thus, the fluid flow is as- calculated as 5.0 × 10−6. Therefore, the ΔtDEMis set to 5.0 × 10−7
sumed laminar. while the CFD time step is set to 5.0 × 10−5 to minimize the computa-
tional loadings.
2.3. Calculation of porosity
3. Boundary and simulation conditions
In the CFD-DEM method, the coupling of fluid and particles is
achieved by solving the porosity (εf) and the fluid-particle interaction The geometry of the screen model used in the present study is
force term (Fpf), where CFD cells and particles are linked with each shown in Fig. 1. The wire-wrapped screen is represented by a rectangu-
other via their locations [23]. Various methods have been proposed to lar model consisting of three slot openings separated by wires with a
calculate porosity, particularly for CFD cells of irregular shapes and dif- width of 2.3 mm. The computational domain is meshed with 80,000
ferent sizes, to ensure numerical stability [15]. Here, the porosity is cal- hexahedra grid cells using a commercial software package ANSYS
culated through two steps. The first step divides each particle in a CFD ICEM™. As done in the model validation, the grids are finer near walls

606
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 2. Particle size distribution of the particles simulated.

1
pffiffiffiffiffiffi e−2ð σ Þ
1 x−μ 2
f n ðxÞ ¼ ð8Þ
σ 2π

where x is the particle diameter. The mean particle diameter studied is


192 μm, which represents a typical field sand size. The PSD width is de-
fined non-dimensionally as the standard deviation normalized by the
Fig. 1. Geometry and mesh representation of the computational model. mean diameter of particles (σ/μ). Fig. 2 shows the PSD widths investi-
gated here, with values ranging from 3 to 10%. Note that the size of
monodisperse particles is equivalent to the mean particle diameter of
and for slot openings compared with the remainder of the domain [19]. 192 μm. The screen geometry is fixed in the study of solid concentration
The widely used outflow boundary is applied at the outlet, where the and PSD effects, where the slot width (w) to mean particle diameter ra-
normal gradients of all variables are set to zero for the fluid phase, as- tio is 1.5. At this size ratio, sand retention by particle bridging is ex-
suming that the flow is fully developed there. Differently, when parti- pected [47]. Such a process is more complicated than sand retention
cles flow out of the outlet, they are removed directly from the by size exclusion. Furthermore, the screen geometry is modified and
simulation. For each simulation, a uniform fluid velocity profile is tested to mitigate non-uniform sand retention.
adopted at the inlet corresponding to the case of constant flow rate, fol- In each simulation, particles are loaded randomly into the flow
lowing the previous experiment [17] and particles are injected with the system near the inlet according to the specified solid concentration.
same velocity as the fluid. The number of injected particles varies ac- The particles descend under gravity, buoyancy and fluid forces. Once
cording to the fluid flow rate and solid concentration specified. No-slip particles flow out of the slots, those particles are regarded as pro-
boundary conditions are taken at the walls for the fluid flow. The front duced sand. The mass flow rate of produced sand particles is re-
and back sides take symmetric boundary conditions for the fluid corded with time for results analysis. Stable sand retention occurs
phase but periodic boundary conditions for particles. Facilitated with when no particles flow out of the slots while the height of the accu-
the above boundary conditions, the CFD-DEM model is solved using mulated particles over the screen continues to increase. The simula-
an in-house code. Fluid with properties of crude oil is adopted to simu- tions stop after 7 s or when the particle bed height reaches a pre-set
late the typical oilfield sand retention and production environments. level.
Table 2 lists the current simulation conditions.
Two variables are considered, including solid concentration (Cv) and 4. Results and discussion
particle size distribution (PSD) width, followed by a new screen design.
The values of solid concentration are in the range of 2 to 16%. The PSDs 4.1. Model validation
are characterized using the normal distribution function with a constant
mean diameter (μ) but different standard deviations (σ): To validate the current CFD-DEM model, the experimental data
from two different applications have been considered [19]. The re-
sults are summarized here. The first application focused on the dry
particles system, following the experimental work of Coberly [47].
Table 2 The predicted bridging criteria are the same as the criteria obtained
Simulation conditions in the current study.
experimentally. The particle retention process is simulated using
Solid phase Fluid phase mixtures of two different particle sizes following this experimental
Material Sand Fluid type Crude oil study. The predicted critical slot widths at which bridging occurs
Shape Spherical Fluid density (kg/m3) 825 are in good agreement with the experimental values. Further
Particle density (kg/m3) 2500 Fluid viscosity (kg/m.s) 0.003 model validation was conducted based on the experimental work
Particle diameter (μm) 166–220 Temperature (°C) 54 of Wu et al. [17], who studied sand retention over a wire-wrapped
Young's modulus (N/m2) 1 × 107 Mesh and geometry screen under slurry flow conditions. The predicted total mass of
Poison's ratio 0.3 Grid type Hexahedra sand produced reasonably accord with that measured in the experi-
Sliding friction coefficient 0.5 Grid number 80,000 ment, with the prediction errors range from −8% to 4%. All the re-
Rolling friction coefficient 0.1 CFD time step 5 × 10−5 sults obtained so far suggest that the current model is reliable for
Restitution coefficient 0.6 DEM time step 5 × 10−7 predicting the screen retention performance and particle-fluid flow
w/D50 1.5
behavior in sand screen, at least qualitatively.

607
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

particle flow, and thus a majority of particles can pass through the
screen, which is undesirable.
ii. Partial sand retention (Mode II): It is featured by single or two-time
drops in the flow rate, followed by a steady flow rate at the later
stage. The single drop in the flow rate corresponds to bridging for-
mation over one slot (Fig. 4 (b)) or simultaneous bridging over
two slots. Fig. 5 (b) demonstrates the partial sand retention where
particles bridge only one slot. As time evolves, particles accumulate
on the top of the blocked slot and build a small dune (t = 6.0 s).
Meanwhile, particle bridges fail to form on the other two slots, and
thus a steady particle flow is observed (t = 7.0 s). Note that stable
sand retention may be possible if the screen operation lasts for a
long time because the number of particles accumulated above the
screen increases with time, as observed experimentally by Wu
et al. [17]. This suggests that Mode II could take considerable time
to achieve stable sand retention, and thus the sand production is
high. Moreover, the wires of the remaining open slots are more vul-
nerable to erosion due to the passing particle-fluid flow with rela-
tively high velocities resulting from the slot plugging.
iii. Sand retention with slow sequential bridging (Mode III): It is fea-
tured by at least two or three-time drops in the flow rate to zero.
These drops correspond to the sequential formation of bridging
over two or three slots (Fig. 4 (c)). This sequential bridging is
shown in Fig. 5 (c). At the beginning, particles are produced through
the three slots. After a short period, one of the slots is bridged by par-
ticles, resulting in the first drop in the flow rate (e.g., t = 1.5 s). As
time evolves, another slot is bridged by particles, and thus the flow
rate drops for the second time. Then, particles are produced merely
from the remaining open slot while sand dune starts to develop over
the two blocked slots (e.g., t = 3.0 s). The slight increase in the flow
rate at t > 4.0 s is attributed to the increase in the particle velocities
from the remaining open slot. At the later stage, bridging forms on
the last slot and a complete stop of particle flow develop eventually
(t = 6.24 s). In Mode III, stable sand retention is achieved but not
Fig. 3. Comparison of the mass flow rate of produced particles at: (a) different grid uniform. The bridging forms in sequential order very slowly, partic-
resolutions; (b) different drag models. ularly over the last slot that may be eroded significantly, hereby the
sand production is high.
iv. Sand retention with fast sequential bridging (Mode IV): This mode
4.2. Sensitivity tests of grid resolution and fluid drag correlation resembles two-time drops in the flow rate to zero, corresponding
to the sequential formation of bridging over two slots, and the se-
The independence of numerical results on grid resolution and fluid quential bridging takes place very quickly (Fig. 4 (d)). As shown
drag model is examined under the Mode III (σ/μ = 3% and Cv = 10%), from Fig. 5 (d), bridging forms almost at the same time over two
where non-uniform sand retention is observed. For this purpose, three slots, resulting in a rapid drop in the flow rate at the first stage
grid solutions are considered, corresponding to the cell number of (e.g., t = 1.46 s). Shortly after that, bridging forms on the last open
180,000 (fine grid), 80,000 (medium grid) and 30,000 (coarse grid), re- slot, and then the particle flow stops (e.g., t = 2.2 s). Because of
spectively. Also, three widely used fluid drag correlations are tested, the fast formation of bridging over these slots, the sand retention
which were reported by Di Felice and Rosa [48], Gidaspow et al. [49] is nearly uniform in Mode IV. Also, the particle flow stops quickly,
and Rong et al. [26]. Fig. 3 (a) and (b) compare the mass flow rate of pro- and thus a majority of particles are retained. In this circumstance,
duced particles at different grid resolutions and drag correlations. It is of the screen sections are quickly covered by particles and protected
interest to note that the numerical results do not show a significant dif- from erosion.
ference with respect to grid solution and drag correlation. The following v. Sand retention with instantaneous bridging (Mode V): It is featured
results are generated based on the medium grid scheme and Rong by a single drop in the flow rate to zero, corresponding to the instan-
et al.'s drag correlation. taneous formation of bridging over three slots (Fig. 4 (e)). As shown
in Fig. 5 (e), after the initial sanding, bridging forms instantaneously
4.3. Sand retention modes and their transition over the three slots (t = 1.12 s) in Mode V, establishing rapid and
uniform sand retention (e.g. t = 1.5 s). This model produces mini-
Fig. 4 presents representative variations of the mass flow rate of pro- mum sand production and prevents local erosion.
duced particles with time at different solid concentrations and PSD
widths. The corresponding particle patterns are given in Fig. 5. Five dis- A phase diagram as given by Fig. 6 is proposed to predict the condi-
tinct sand retention modes can be identified according to the mass flow tions under which different sand retention models occur. It covers the
rate characteristics: solid concentrations and PSD widths considered in the current study.
In this diagram, non-uniform and uniform sand retention can also be
i. No sand retention (Mode I): It is featured by a stable flow rate in the
identified. The non-uniform mode includes Modes II-IV, while the uni-
entire simulation time (Fig. 4 (a)), which indicates a continuous
form one includes Modes I and V. As shown in Fig. 6, Mode I is dominant
sanding event. The particle flow patterns in Fig. 5 (a) show that in
at low solid concentrations such as 2 to 4% for both monodisperse and
Mode I, no particle bridging forms over the three slots to stop the
polydisperse particles. However, at a slightly higher solid concentration

608
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 4. Mass flow rates of produced particles during the retention process: (a) Mode I, (b) Mode II, (c) Mode III, (d) Mode IV and (e) Mode V.

such as 6%, this sand retention mode is only observed for monodisperse examined. Figs. 7 and 8 show the representative results, which corre-
and relatively small PSD widths (e.g., σ/μ ≤ 10 here). Mode II is observed spond to the cases for the monodisperse particles at Cv = 10% and
for both monodisperse and polydisperse particles at intermediate solid Cv = 16%, respectively. In these figures, particles and their vectors
concentrations such as 8%. However, it is also observed at a low solid are colored by velocity magnitudes, while fluid velocities are shaded
concentration, such as 6% for the widest PSD (σ/μ = 10%) and at a higher by their axial component. This treatment applies other figures in the
solid concentration, such as 10% for monodisperse particles. Differently, following discussion. Fig. 7 shows the particle and fluid flow patterns,
at a solid concentration of 10%, Mode III is observed for polydisperse whose sand retention is in Mode II. At the beginning of the sand re-
particles. High solid concentrations such as 13% and 16% lead to, respec- tention process (t = 0.7 s), as particles fall on the screen randomly,
tively, Mode IV and Mode V for both monodisperse and polydisperse some of these particles settle on the wires, forming small sand beds
particles. Overall, with increasing solid concentration, the sand reten- and ridges, while particles above the slots would be produced along
tion transits from Mode I to Mode V, which means that the sand reten- with the fluid. At this stage, the fluid flow is well distributed along
tion is improved from no occurrence to non-uniform behavior, and the length of the screen, featured by similar velocity profiles inside
finally to uniform behavior. It is consistent with the experimental obser- the three slots. As time evolves, the sand beds and ridges grow, and
vation for the solid filtration system [47]. This has not been obtained in at this stage (t = 1.02 s), bridging starts to form on the middle slot
the existing CFD-DEM studies of screen retention, as discussed previ- (S2). It occurs when particles in the vicinity of the slot flow to the
ously. Fig. 6 also shows that the sand retention mode transitions to slot after colliding with the descending particles. The particle flow
Mode V faster with increasing PSD widths. Furthermore, at high solid near the slot is evidenced by the fact that the particle velocity
concentrations such as 13% and 16% where particle number is sufficient vectors are pointing to the slot as shown in the Fig. 7. (ii). This phe-
for bridging, the sand retention mode does not change with increasing nomenon is also observed by Wu et al. [17] in their experiment during
PSD widths. The results highlight that sand retention is very sensitive the initial stage of the sand retention process. At lower solid concen-
to solid concentration under the conditions considered, and the impact trations such as 2 to 6%, there is no particle accumulation on the
of PSD widths is more pronounced at low to intermediate solid concen- screen wall; thus, bridging does not occur. Another interesting obser-
trations such as 6 to 10%. vation is the slots are closed in succession, and the closing of each slot
influences the bridging on the neighboring slots. For example, when
4.4. Analysis of solid concentration effect no subsequent particles can pass through the S2, the sand dune starts
to develop on this slot, covering more area of the screen gradually.
To understand the effect of solid concentration, the particle flow After that (t = 1.58 s), bridging is formed on the S1. Meanwhile, the
patterns, together with the particle and liquid velocity vectors, are fluid velocity inside the S2 slightly decreases because of the particle

609
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 5. Snapshots showing particle flow patterns at different sand retention modes: (a) Mode I, (b) Mode II, (c) Mode III, (d) Mode IV and (e) Mode V.

accumulation over this slot, as shown in Fig. 7 (iii). This leads to a S1 is completed, more particles are retained above the screen; thus,
small increase in the downward fluid velocity magnitude in the left the sand dune grows and span over the S1 and S2 (t = 3.0 s). This
slot (S1) and the right slot (S3). After the bridge formation on the causes an increase in the downward fluid velocity magnitude inside
the S3. At the later stage (t = 5.24 s), the fluid velocity in the S3 is sig-
nificantly increased as the sand dune continues to grow. Conse-
quently, the descending particles are entrained strongly by the fluid
and pass through the slot. This strong particle flow in this slot and
its neighboring region, featured by large particle velocities, hinders
the formation of particle bridges over this slot.
Fig. 8 shows the instantaneous formation of bridging over three slots
in Mode V, which occurs at the beginning of the retention process (t =
0.88 s). At a high solid concentration such as 16%, considerable particles
are transported to the screen at a time, increases the number of particles
arriving at the slot entrance and settling on the screen wall, as shown in
Fig. 8 (i). As a result, particles attempting to enter these slots are more
than those able to pass through, which facilitates rapid bridging (t =
0.88 s). After all, slots are closed, a uniform layer of sand forms on the
Fig. 6. Phase diagram showing sand retention modes and their transition with respect to
screen (t = 1.04 s and t = 1.5 s). Correspondingly, the fluid flow
PSD width and solid concentration. along the length of the screen is evenly distributed and undisturbed,

610
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 7. Sand retention behavior for monodisperse particles at Cv = 10%: (a) particle flow pattern, (b) particle flow field and (c) fluid flow field.

Fig. 8. Sand retention behavior for monodisperse particles at Cv = 16%: (a) particle flow pattern, (b) particle flow field and (c) fluid flow field.

611
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 9. Evolution of inter-particle force network at different solid concentrations for monodisperse particles: (a) Cv = 10% and (b) Cv = 16%.

Fig. 10. Sand retention behavior at σ/μ = 10% and Cv = 6% corresponding to Mode II: (a) particle flow pattern, (b) particle flow field and (c) fluid flow field.

612
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

featured by similar and constant velocity profiles inside the three slots the smallest and largest particles. Therefore, the case of σ/μ = 10% con-
throughout the retention process (Fig. 8 (iii)). tains the largest particles under the current work, which underlies the
As shown in Fig. 9, the evolution of the inter-particle force network is occurrence of sand retention on the screen. Such a result can be clearly
examined to understand the difference in bridging between Mode II and seen in Fig. 10 (i). It shows that different sized particles are widely
Mode V, corresponding to the particle flow patterns shown in Fig. 7 spread when flowing downward. Part of the particles flows out of the
(i) and Fig. 8 (i). Here, each stick represents one connection of two par- slots at the early stage of the retention process (t = 1.0 s). Bridging is
ticle centers, and its thickness denotes the magnitude of the contact hard to form because the number of particles is not sufficient. Only a
force. The rectangular boxes S1, S2 and S3, are the left, middle and few particles arrive at the slot entrance at a time and settle on the
right slots, respectively. As shown in Fig. 9 (a), for mode II, the first screen. Consequently, there is less particle-particle interaction near
bridging occurs over the S2 when small force chains appearing on this the slot. However, at the later stage (t = 4.38 s), bridging occurs when
slot; and these chains propagate horizontally and connect the force the largest particles bridge over the S1. This can happen because at a
chains at the left and the right sides of the slot, leading to the formation given slot opening, increasing the particle diameter would reduce the
of bridges. At the later stage, while the force chains on the S2 continue to size ratio between particle and slot, which in turn increases the proba-
grow, small force chains start to appear on the S1 (t = 1.58 s), which bility of bridging configuration. In this circumstance, the largest parti-
later also propagate over the entire slot. The result again shows cles in the distribution are the controlling factor that causes sand
that bridging over these slots are formed in succession. However, retention. These particles are the first to bridge and begin the formation
above the S3 region, contacts between particles cannot be apparently of sand bed on the screen, which is consistent with experimental obser-
observed, indicating that no bridge is formed here, so particles in this re- vations by Ballard et al. [42,50]. Fig. 10 (iii) shows that after sand dune
gion can flow almost freely. Conversely, small force chains simulta- forms over the S1 (t = 7.0 s), the magnitudes of the downward fluid ve-
neously appear on the S1, S2 and S3 at the very beginning (t = 0.88 s) locities inside this slot decrease, which causes an increase in the fluid
in Mode V, as shown in Fig. 9 (b). The particle flow stops after these velocity magnitudes inside the S2 and S3. Such a situation could hinder
force chains propagate over these slots to form strong bridging. Note the formation of particle bridges on these two slots.
that despite the retention modes, particles over the slots with bridging Another example that demonstrates a different bridging phenome-
experience large forces. This is due to the compression exerted by the non is shown in Fig. 11, which is in model III at σ/μ = 10% and Cv =
fluid flow through the slots. 10%. In such cases, stable sand retention can occur to polydisperse
particles but not to monodisperse particles. As shown in Fig. 11 (i), at
4.5. Analysis of particle size distribution width effect t = 1.02 s, bridging begins from the middle slot (S2) when the largest
particles bridge over this slot. As the sand dune develops over the S2,
Fig. 10 demonstrates how the PSD width (σ/μ) affect the sand reten- another bridging occurs over the S3. It also happens when the largest
tion process at σ/μ = 10% and Cv = 6%. Under these conditions, sand particles bridge over this slot (t = 1.9 s). After that (t = 3.0 s), the
retention can occur in the form of Mode II. However, this does not sand dune grows over the S2 and S3. Meanwhile, the fluid velocity
occur to the monodisperse particles and smaller σ/μ such as 3% and magnitudes inside these two slots decrease, which increases the
5%. Note that a larger PSD width implies a bigger difference between downward fluid velocity magnitudes inside the S1. Later, the fluid

Fig. 11. Sand retention behavior at σ/μ = 10% and Cv = 10% corresponding to Mode III: (a) particle flow pattern, (b) particle flow field and (c) fluid flow field.

613
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 12. Mesh distribution of (a) the conventional wire-wrapped screen model, (b) θ = 2° and (c) θ = 10°.

velocity magnitudes inside the S1 increase significantly when the sand has slots with a “V” cross-sectional area, aiming to bring particles into
dune grows larger. Despite the high flow velocities that develop inside contact to bridge. The new screen with different angles of convergence
and surrounding the S1, bridging occurs over this slot (t = 5.24 s). It (θ) (see Fig. 12) is tested at different solid concentrations and PSD
happens when the largest particles initiate the bridge formation, as widths. Based on the current concept, the conventional screen corre-
demonstrated in Fig. 11 (iii). After that (t = 6.26 s), stable sand sponds to θ = 0°.
retention is quickly achieved with no further particles can flow out of Fig. 13 shows some representative results of the new screen for the
the slot while the height of the accumulated particles continues monodisperse particle at Cv = 6%, including the mass flow rates of
to increase. These results also suggest that the likelihood of sand produced particles and the corresponding particle flow patterns. At
retention by bridging increases with the presence of large particles in θ = 0° (Fig. 13 (a)), particles flow out of the three slots continuously;
a distribution. therefore, the flow rate is constant. This flow pattern corresponds to
Mode I, in which sand retention does not occur. However, at θ = 2°, par-
4.6. New screen design ticles flow out of the three slots only at the beginning (t > 2.0 s), and
then a complete stop of particle flow is established. The flow rate
As discussed previously, controlling bridging plays a critical role in quickly decreases to zero. This sand retention mode corresponds to
managing sand retention and mitigating sand production. In this direc- Mode IV, in which sand retention is nearly uniform. However, with in-
tion, a novel screen system with a converging slot configuration is pro- creasing angle of convergence (e.g., θ = 10°), sand retention occurs
posed to facilitate bridging in this study. Unlike the conventional wire- but is not uniform. The flow rate drops when two slots (S1 and S2 in
wrapped screens with straight or keystone shaped slots, the new screen Fig. 13 (c)) bridge by particles simultaneously (e.g., t = 4.0 s). After

Fig. 13. Mass flow rates of produced particles (left) and particle flow patterns (right) at different angle of convergences for monodisperse particles at Cv = 6%: (a) θ = 0°, (b) θ = 2° and
(c) θ = 10°.

614
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 14. Mass flow rates of produced particles (left) and particle flow patterns (right) at different angle of convergences for monodisperse particles at Cv = 8%: (a) θ = 0°, (b) θ = 2° and
(c) θ = 10°.

that, particles accumulate above these two slots, forming a thick sand Fig. 16 (a) for θ = 2°, the two particles initially flow parallelly inside
dune. However, bridges fail to form inside the S3 to stop the particle the slot (t = 0.6 s) and then bridge at the upper section where the
flow (t = 7.0 s). This sand retention mode corresponds to Mode II. slot is relatively confined (t = 0.62). The bridging is evidenced by a
Fig. 14 is another example to demonstrate the benefit of the new
screen. For monodisperse particles at Cv = 8%, sand retention occurs
but is not uniform at θ = 0°. The flow rate firstly drops (e.g., t = 5.0 s)
due to the bridges formed over the S1 and then drops again (e.g., t =
7.0 s) after the second bridge forms over the S3. However, bridges fail
to develop over the S2, and thus particles can flow through this slot.
This sand retention mode corresponds to Mode II. At θ = 2°, sand pro-
duction occurs only within a short period (t < 1.5 s) before the particle
flow stops completely. The flow rate drops to zero almost instantly
after the retention begins. This phenomenon is caused by the fast
bridging inside the three slots (e.g., t = 1.5 s). This sand retention
mode corresponds to Mode V, in which sand retention is uniform, as
demonstrated in Fig. 14 (b). However, when the θ continues to increase
(e.g., θ = 10°), sand retention becomes non-uniform, corresponding to
Mode III.
The phase diagram predicting sand retention modes is also estab-
lished for the new screen, as given in Fig. 15. The comparison of Figs. 6
and 15 suggests the new screens with θ = 2° and θ = 10° are better
in developing uniform sand retention than the conventional screen,
and the one with θ = 2° is more effective. Fig. 16 compares the particle
flows and inter-particle force network in the new screen system with
θ = 2° and θ = 10° to achieve a better understanding. The analysis fo-
cuses on one slot for clarity, and the results of the monodisperse parti-
cles at Cv = 8% are considered as an example. For the conventional
screen (θ = 0°), bridging forms over the slot, and it occurs only after
enough particles accumulate on the screen. Differently, as particles are
flowing in the new screen system, bridging forms inside the slot. It
occurs when particles enter the slot and flow parallel to each other
inside the slot. When the slot angle is narrow such as θ = 2°, bridging
forms very quickly. This happens because these particles cannot
reposition themselves inside the slot, thus they bridge when they Fig. 15. Phase diagram showing sand retention modes and their transition for the new
arrive at the confine section of the slot. For instance, as shown by screen system: (a) θ = 2° and (b) θ = 10°.

615
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Fig. 16. Development of particle behavior until bridging in the new screens with θ = 2° (left) and (b) θ = 10° (right): (a) particle flow pattern, (b) particle flow field and (c) force network.

large contact force appearing at the upper section of the slot, resulting (2) Increasing solid concentration enhances particle flow near slots,
from the high-speed fluid flow there. After the bridging event, particles which induces the instantaneous formation of particle bridges
pile up inside the slot and start forming a bed on the screen. However, on the three slots and the uniform sand retention. Also, bridging
bridging inside a slot becomes harder when the angle of convergence depends on the fluid flow velocity in the slots, which changes
increases to θ = 10° (Fig. 16 (b)). Consequently, bridging occurs at the with the slot closure. Therefore, with the increasing proportion
lower slot section, where particle flow is relatively strong due to the re- of closed slots, the bridging in the remaining open slot needs a
duced passage. Bridging could be hindered by high particle velocities, longer time. By contrast, the introduction of polydispersity
which leads to the reduced sand retention performance at θ = 10° helps increase the bridging probability at low to intermediate
than θ = 2°. solid concentrations. This is attributed to the presence of large
particles in a distribution, which helps initiate the bridging and
5. Conclusion thus cause sand retention on the screen.
(3) A novel screen system with a converging slot configuration has
The non-uniform sand retention behaviors over wire-wrapped been successfully designed. New screens significantly improve
screens are studied using a CFD-DEM model, considering the effects of the sand retention with the angle of convergence of 2° and 10°,
solid concentration, particle size distribution width and screen design. in which uniform sand retention occurs at a much lower solid
The major findings of the current study can be summarized as follows: concentration compared to the conventional screen. It is mainly
because bridging forms inside the slot in the new screen. How-
(1) Five distinct sand retention modes can be identified from mass
ever, when the angle of convergence changes from 2° to 10°,
flow rate characteristics of produced particles: Mode I, in which
the performance slightly drops. Correspondingly, the position
sand retention does not occur; Mode II, in which sand retention
of bridging moves down from the upper section of the slot to
occurs only on one or two slots; Mode III and IV, in which sand
the lower section. The latter position has a relatively strong par-
retention occurs via sequential bridging over three slots, and
ticle flow, hindering bridging to some extent.
this occurs much faster for the latter; and, Mode V, in which
sand retention occurs instantaneously over three slots, resulting
Declaration of Competing Interest
in uniform sand retention. A phase diagram has been established
to predict these sand retention modes with respect to solid con-
The authors declare that they have no known competing financial
centration and PSD width. It is shown that the sand retention
interests or personal relationships that could have appeared to influ-
transits from Mode I to Mode V with increasing solid concentra-
ence the work reported in this paper.
tions, and a larger PSD width enhances this transition.

616
N.I. Ismail, S. Kuang, M. Zhou et al. Powder Technology 395 (2022) 604–617

Acknowledgements [24] Z. Zhou, S. Kuang, K. Chu, A. Yu, Discrete particle simulation of particle–fluid flow:
model formulations and their applicability, J. Fluid Mech. 661 (2010) 482–510.
[25] S. Kuang, K. Li, R. Zou, R. Pan, A. Yu, Application of periodic boundary conditions to
The authors are grateful to the Australian Research Council (ARC) CFD-DEM simulation of gas–solid flow in pneumatic conveying, Chem. Eng. Sci. 93
(IH140100035 and LP160100819) for the financial support of this work (2013) 214–228.
and the National Computational Infrastructure (NCI) for the use of its [26] L.W. Rong, K.J. Dong, A. Yu, Lattice-Boltzmann simulation of fluid flow through
packed beds of uniform spheres: effect of porosity, Chem. Eng. Sci. 99 (2013) 44–58.
computational facilities. [27] K. Chu, S. Kuang, Z. Zhou, A. Yu, Model a vs. model B in the modelling of particle-
fluid flow, Powder Technol. 329 (2018) 47–54.
References [28] L. Ji, K.W. Chu, S.B. Kuang, J. Chen, A.B. Yu, Modelling the multiphase flow in
hydrocyclones using coarse-grained VOF-DEM and mixture-DEM approaches, Ind.
[1] E. Khamehchi, E. Reisi, Sand production prediction using ratio of shear modulus to Eng. Chem. Res. 57 (2018) 9641–9655.
bulk compressibility (case study), Egypt. J. Pet. 24 (2015) 113–118. [29] M. Zhou, S. Wang, S. Kuang, K. Luo, J. Fan, A. Yu, CFD-DEM modelling of hydraulic
[2] N.N.A. Razak, S.J. Abdulkadir, M.A. Maoinser, S.N.A. Shaffee, M.G. Ragab, One- conveying of solid particles in a vertical pipe, Powder Technol. 354 (2019) 893–905.
dimensional convolutional neural network with adaptive moment estimation for [30] M. Zhou, S. Kuang, K. Luo, R. Zou, S. Wang, A. Yu, Modeling and analysis of flow re-
modelling of the sand retention test, Appl. Sci. 11 (2021) 3802. gimes in hydraulic conveying of coarse particles, Powder Technol. 373 (2020)
[3] S. Hamid, S. Ali, Causes of sand control screen failures and their remedies, SPE 543–554.
European Formation Damage Conference, Society of Petroleum Engineers, 1997. [31] Z. Qi, S. Kuang, T. Qiu, A. Yu, Lattice Boltzmann investigation on fluid flows through
[4] A. Procyk, M. Whitlock, S. Ali, Plugging-induced screen erosion difficult to prevent, packed beds: interaction between fluid rheology and bed properties, Powder
Oil & Gas J. 96 (1998) 80–90. Technol. 369 (2020) 248–260.
[5] T. Hallman, V. Yung, A. Albertson, Analysis of Wellbore Failures and re-Design of [32] S.B. Kuang, K. Li, A.B. Yu, CFD-DEM simulation of large-scale dilute-phase pneumatic
Slotted Liners for Horizontal Wells Applied in a Heavy Oilfield, SPE Western Re- conveying system, Ind. Eng. Chem. Res. 59 (2020) 4150–4160.
gional Meeting, Society of Petroleum Engineers, 2015. [33] E.Z. Zheng, M. Rudman, S.B. Kuang, A. Chryssc, Turbulent coarse-particle suspension
[6] D.L. Tiffin, M.H. Stein, X. Wang, Drawdown Guidelines for Sand Control Completions, flow: measurement and modelling, Powder Technol. 373 (2020) 647–659.
SPE Annual Technical Conference and Exhibition, Society of Petroleum Engineers,
[34] E.Z. Zheng, M. Rudman, S.B. Kuang, A. Chryss, Turbulent coarse-particle non-
2003.
Newtonian suspension flow in a pipe, Int. J. Multiphase Flow 142 (2021) 103698.
[7] S. Mondal, C.-H. Wu, M.M. Sharma, R.A. Chanpura, M. Parlar, J.A. Ayoub, Character-
[35] M.M. Zhou, S.B. Kuang, F. Xiao, K. Luo, A.B. Yu, CFD-DEM analysis of hydraulic con-
izing, designing, and selecting metal mesh screens for standalone-screen applica-
veying bends: interaction between pipe orientation and flow regime, Powder
tions, SPE Drill. Complet. 31 (2016) 085–094.
Technol. 392 (2021) 619–631.
[8] P. Jop, Y. Forterre, O. Pouliquen, A constitutive law for dense granular flows, Nature
441 (2006) 727–730. [36] Z. Qi, S.B. Kuang, L.W. Rong, A. Yu, Lattice Boltzmann investigation of the wake effect
[9] M. Wang, X.Y. Lan, C.X. Wang, J.S. Gao, A.B. Yu, S.B. Kuang, Numerical simulation of on the interaction between particle and power-law fluid flow, Powder Technol. 326
the pilot-scale high-density circulating fluidized bed riser, Ind. Eng. Chem. Res. 60 (2018) 208–221.
(2021) 3184–3197. [37] P.A. Cundall, O.D. Strack, A discrete numerical model for granular assemblies,
[10] H. Zhu, Z. Zhou, R. Yang, A. Yu, Discrete particle simulation of particulate systems: Geotechnique 29 (1979) 47–65.
theoretical developments, Chem. Eng. Sci. 62 (2007) 3378–3396. [38] A. Di Renzo, F.P. Di Maio, Comparison of contact-force models for the simulation of
[11] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate sys- collisions in DEM-based granular flow codes, Chem. Eng. Sci. 59 (2004) 525–541.
tems: a review of mayor applications and findings, Chem. Eng. Sci. 63 (2008) [39] H. Kruggel-Emden, E. Simsek, S. Rickelt, S. Wirtz, V. Scherer, Review and extension
5728–5770. of normal force models for the discrete element method, Powder Technol. 171
[12] S.B. Kuang, Z.Y. Li, A.B. Yu, Review on modeling and simulation of blast furnace, Steel (2007) 157–173.
Res. Int. 89 (2018) 1700071–1700071/1700025. [40] R.A. Chanpura, R.M. Hodge, J.S. Andrews, E.P. Toffanin, T. Moen, M. Parlar, A review
[13] M. Sakai, Y. Mori, X. Sun, K. Takabatake, Recent progress on mesh-free particle of screen selection for standalone applications and a new methodology, SPE Drill.
methods for simulations of multi-phase flows: a review, KONA Powder Particle J. Complet. 26 (2011) 84–95.
37 (2020) 132–144. [41] T. Ballard, S.P. Beare, Sand Retention Testing: The More you Do, the Worse it Gets,
[14] Z. Peng, E. Doroodchi, B. Moghtaderi, Heat transfer modelling in discrete element SPE International Symposium and Exhibition on Formation Damage Control, Society
method (DEM)-based simulations of thermal processes: theory and model develop- of Petroleum Engineers, 2006.
ment, Prog. Energy Combust. Sci. 79 (2020) 100847. [42] T.J. Ballard, S.P. Beare, An Investigation of Sand Retention Testing with a View to De-
[15] S. Kuang, M. Zhou, A. Yu, CFD-DEM modelling and simulation of pneumatic convey- veloping Better Guidelines for Screen Selection, SPE International Symposium and
ing: a review, Powder Technol. 365 (2020) 186–207. Exhibition on Formation Damage Control, Society of Petroleum Engineers, 2012.
[16] Y. Feng, X. Choi, B. Wu, C.Y. Wong, Evaluation of Sand Screen Performance Using a [43] A.M. Mathisen, G.L. Aastveit, E. Alteraas, Successful Installation of Stand Alone Sand
Discrete Element Model, SPE Asia Pacific Oil and Gas Conference and Exhibition, So- Screen in More than 200 Wells-the Importance of Screen Selection Process and Fluid
ciety of Petroleum Engineers, 2012. Qualification, European Formation Damage Conference, Society of Petroleum Engi-
[17] B. Wu, S. Choi, Y. Feng, R. Denke, T. Barton, C. Wong, J. Boulanger, W. Yang, S. Lim, M. neers, 2007.
Zamberi, Evaluating Sand Screen Performance Using Improved Sand Retention Test [44] D.R. Underdown, S. Hopkins, Design and implementation of retention/filtration
and Numerical Modelling, Offshore Technology Conference Asia, Offshore Technol- media for sand control, SPE Drill. Complet. 23 (2008) 235–241.
ogy Conference, 2016.
[45] R. Sun, H. Xiao, Diffusion-based coarse graining in hybrid continuum–discrete
[18] S.N.A. Shaffee, P.F. Luckham, O.K. Matar, A. Karnik, M.S.A. Zamberi, Numerical inves-
solvers: applications in CFD–DEM, Int. J. Multiphase Flow 72 (2015) 233–247.
tigation of sand-screen performance in the presence of adhesive effects for en-
[46] R.I. Issa, Solution of the implicitly discretised fluid flow equations by operator-
hanced sand control, SPE J. 24 (2019) 2195–2208.
splitting, J. Comput. Phys. 62 (1986) 40–65.
[19] N.I. Ismail, S. Kuang, A. Yu, CFD-DEM study of particle-fluid flow and retention per-
formance of sand screen, Powder Technol. 378 (2021) 410–420. [47] C. Coberly, Selection of Screen Openings for Unconsolidated Sands, Drilling and pro-
[20] F. Razavi, A. Komrakova, C.F. Lange, CFD–DEM simulation of sand-retention mecha- duction practice, American Petroleum Institute, 1937.
nisms in slurry flow, Energies 14 (2021) 3797. [48] R. Di Felice, The voidage function for fluid-particle interaction systems, Int. J. Multi-
[21] B. Xu, A. Yu, Numerical simulation of the gas-solid flow in a fluidized bed by com- phase Flow 20 (1994) 153–159.
bining discrete particle method with computational fluid dynamics, Chem. Eng. [49] D. Gidaspow, R. Bezburuah, J. Ding, Hydrodynamics of Circulating Fluidized Beds: Ki-
Sci. 52 (1997) 2785–2809. netic Theory Approach, Illinois Inst. of Tech., Chicago, IL (United States). Dept. of
[22] S. Kuang, K. Chu, A. Yu, Z. Zou, Y. Feng, Computational investigation of horizontal Chemical …, 1991.
slug flow in pneumatic conveying, Ind. Eng. Chem. Res. 47 (2008) 470–480. [50] T. Ballard, S. Beare, Media Sizing for Premium Sand Screens: Dutch Twill Weaves,
[23] S. Kuang, A. Yu, Z. Zou, A new point-locating algorithm under three-dimensional hy- SPE European Formation Damage Conference, Society of Petroleum Engineers, 2003.
brid meshes, Int. J. Multiphase Flow 34 (2008) 1023–1030.

617

You might also like