0% found this document useful (0 votes)
10 views

Evaluation of RANS-DEM and LES-DEM Methods

This article evaluates the efficiency of RANS-DEM and LES-DEM methods in OpenFOAM for simulating particle-laden turbulent flows, highlighting challenges in CFD-DEM modeling due to computational demands and method limitations. The study finds that one-way coupling is sufficient for low particle concentrations and suggests approaches for estimating particle boundary conditions when data is lacking. Additionally, it emphasizes the importance of considering sub-grid scale effects in LES-DEM and the need for dynamic turbulence models to improve accuracy in simulations.

Uploaded by

azeez oluwaseyi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views

Evaluation of RANS-DEM and LES-DEM Methods

This article evaluates the efficiency of RANS-DEM and LES-DEM methods in OpenFOAM for simulating particle-laden turbulent flows, highlighting challenges in CFD-DEM modeling due to computational demands and method limitations. The study finds that one-way coupling is sufficient for low particle concentrations and suggests approaches for estimating particle boundary conditions when data is lacking. Additionally, it emphasizes the importance of considering sub-grid scale effects in LES-DEM and the need for dynamic turbulence models to improve accuracy in simulations.

Uploaded by

azeez oluwaseyi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 26

fluids

Article
Evaluation of RANS-DEM and LES-DEM Methods in
OpenFOAM for Simulation of Particle-Laden Turbulent Flows
Atul Jaiswal * , Minh Duc Bui and Peter Rutschmann

Chair of Hydraulic and Water Resources Engineering, Technical University of Munich (TUM), Arcisstr. 21,
80333 Munich, Germany
* Correspondence: [email protected]; Tel.: +49-(89)-289-23161

Abstract: CFD-DEM modelling of particle-laden turbulent flow is challenging in terms of the required
and obtained CFD resolution, heavy DEM computations, and the limitations of the method. Here,
we assess the efficiency of a particle-tracking solver in OpenFOAM with RANS-DEM and LES-DEM
approaches under the unresolved CFD-DEM framework. Furthermore, we investigate aspects of the
unresolved CFD-DEM method with regard to the coupling regime, particle boundary condition and
turbulence modelling. Applying one-way and two-way coupling to our RANS-DEM simulations
demonstrates that it is sufficient to include one-way coupling when the particle concentration is small
(O ~ 10−5 ). Moreover, our study suggests an approach to estimate the particle boundary condition
for cases when data is unavailable. In contrast to what has been previously reported for the adopted
case, our RANS-DEM results demonstrate that simple dispersion models considerably underpredict
particle dispersion and previously observed reasonable particle dispersion were due to an error in
the numerical setup rather than the used dispersion model claiming to include turbulence effects
on particle trajectories. LES-DEM may restrict extreme mesh refinement, and, under such scenarios,
dynamic LES turbulence models seem to overcome the poor performance of static LES turbulence
models. Sub-grade scale effects cannot be neglected when using coarse mesh resolution in LES-DEM
Citation: Jaiswal, A.; Bui, M.D.;
Rutschmann, P. Evaluation of
and must be recovered with efficient modelling approaches to predict accurate particle dispersion.
RANS-DEM and LES-DEM Methods
in OpenFOAM for Simulation of Keywords: particle-laden BFS; turbulent flows; dispersion model; unresolved CFD-DEM;
Particle-Laden Turbulent Flows. RANS-DEM; LES-DEM; OpenFOAM
Fluids 2022, 7, 337. https://doi.org/
10.3390/fluids7100337

Academic Editors: Mehrdad


1. Introduction
Massoudi and Sourabh V. Apte
Two-phase systems consisting of a continuum phase (fluid) and a discrete phase (parti-
Received: 5 September 2022
cles) are prevalent in industrial processes, biological phenomena, and nature. The behavior
Accepted: 12 October 2022
of solid particles in continuous fluid flow is determined by complex physics and depends
Published: 21 October 2022
on the particle and fluid characteristics and flow regime. According to Crowe et al. [1],
Publisher’s Note: MDPI stays neutral five key factors contribute to the turbulence modulation induced by particles: (1) Surface
with regard to jurisdictional claims in effects: particle size normalized by a length scale d p /l; (2) inertial effects: flow Reynolds
published maps and institutional affil- number Re and particle Reynolds number Rep ; (3) response effects: particle response time
iations. or Stokes number St; (4) loading effects: particle volume fraction αp ; and (5) interaction
effects: particle-particle as well as particle-wall. Numerical simulations of such systems
can be helpful in providing a detailed insight into the complex physics involved in particle
motion. However, modelling of particle motions in turbulent flow is difficult because it
Copyright: © 2022 by the authors.
involves the modelling of the surrounding flow field and resulting pressure gradients as
Licensee MDPI, Basel, Switzerland.
well as the particle-flow interaction, which involves the local flow around the particle and
This article is an open access article
the forces resulting from the stress applied on the particle by the flow [2].
distributed under the terms and
In computational fluid dynamics (CFD), both phases can be treated as continuum
conditions of the Creative Commons
Attribution (CC BY) license (https://
medium, also called the Eulerian–Eulerian method, in which Navier–Stokes (NS) equations
creativecommons.org/licenses/by/
are solved for each phase, including the momentum exchange between the phases. The
4.0/). Eulerian–Eulerian method is computationally less expensive but at the cost of losing the

Fluids 2022, 7, 337. https://doi.org/10.3390/fluids7100337 https://www.mdpi.com/journal/fluids


Fluids 2022, 7, x FOR PEER REVIEW 2 of 26

In computational fluid dynamics (CFD), both phases can be treated as continuum


Fluids 2022, 7, 337 medium, also called the Eulerian–Eulerian method, in which Navier–Stokes (NS) 2equa- of 26
tions are solved for each phase, including the momentum exchange between the phases.
The Eulerian–Eulerian method is computationally less expensive but at the cost of losing
the discrete
discrete nature
nature of particles.
of particles. OnOn thethe other
other hand,a aLagrangian
hand, Lagrangianmethodmethodisisadopted
adopted for for
granular
granular systems, also called the Discrete Element Method (DEM), where
called the Discrete Element Method (DEM), where each particle each particle is
tracked
is tracked using
usingNewton’s
Newton’s second
secondlaw law
of motion. With an
of motion. With increase in computational
an increase power,
in computational
advancements
power, in CFDin
advancements andCFDDEM,
andandDEM,improved solutionsolution
and improved algorithms over theover
algorithms last the
decade,
last
the newthe
decade, andnew
more anddetailed approach
more detailed combining
approach CFD andCFD
combining DEM forDEM
and multiphase systems,
for multiphase
called thecalled
systems, Eulerian–Lagrangian method (CFD-DEM),
the Eulerian–Lagrangian method (CFD-DEM),has become hasabecome
populara tool to inves-
popular tool
to investigate particle-laden flows. In particular, numerical approaches combining and
tigate particle-laden flows. In particular, numerical approaches combining the CFD the
DEMand
CFD haveDEMprovenhaveto be advantageous
proven over many other
to be advantageous over options in terms
many other of computational
options in terms of
efficiency and numerical
computational efficiency andconvenience
numerical [3]. In CFD-DEM,
convenience [3]. the continuum the
In CFD-DEM, phase (fluid) is
continuum
phase
resolved(fluid)
usingis resolved using thewhereas
the NS equations, NS equations, whereas
the discrete phasethe(particles)
discrete phase (particles)
is tracked is
by solv-
tracked by solving
ing Newton’s Newton’s
second law ofsecond
motionlaw forof motion
each for each
particle in theparticle in the fluid
fluid system. Thesystem. The
continuum
continuum
and discrete and discrete
phase are phase are alsowith
also coupled coupled
eachwith
othereach other
using using momentum
momentum transfer transfer
mecha-
mechanisms. This coupling level depends on the volumetric fraction
nisms. This coupling level depends on the volumetric fraction of solid material αp = Vp/V of solid material
p=
αin Vpsystem,
the /V in the system,
where Vp iswhere Vpvolume,
particle is particle
andvolume,
V is total and V is total
volume volumeand
of particles of particles
fluid. A
and fluid. A classification
classification map is depicted map isindepicted
Figure 1,inwhich
Figurecan
1, which
be usedcantobeincorporate
used to incorporate
the levelthe
of
level of coupling
coupling in numerical
in numerical simulations
simulations [4]. Furthermore,
[4]. Furthermore, the approximated
the approximated CFDCFD solution
solution can
can be obtained
be obtained using
using Reynolds-averaged
Reynolds-averaged Navier–Stokesequations
Navier–Stokes equations(RANS)
(RANS) or or Large-Eddy
Large-Eddy
simulation
simulation (LES), instead of solving the flow using Direct Numerical Simulation (DNS),
(LES), instead of solving the flow using Direct Numerical Simulation (DNS),
which
which could save significant
could save significantcomputational
computationalefforts,
efforts,especially
especially when
when tracking
tracking particles
particles us-
using
ing DEM.DEM.

Figure 1.1. Classification


Figure Classification map
map [4]
[4] showing
showing the
the level
level of
ofcoupling
coupling required
requiredfor
fornumerical
numericalsimulations
simulations
and interaction between particles and turbulence for  one-way coupling,  two-way
and interaction between particles and turbulence for 1 one-way coupling, 2 two-way coupling coupling
where particles enhance the turbulent production,  two-way coupling where particles enhance
where particles enhance the turbulent production, 3 two-way coupling where particles enhance the
the turbulence dissipation,  four-way coupling.
turbulence dissipation, 4 four-way coupling.

The CFD-DEM
The CFD-DEM methodmethod can
canbebecategorized
categorizedinto
intotwo
twoapproaches:
approaches: unresolved
unresolved andand re-
solved CFD-DEM. Unresolved CFD-DEM solves the flow at larger scales
resolved CFD-DEM. Unresolved CFD-DEM solves the flow at larger scales using filtering using filtering
(LES)/averaging (RANS)
(LES)/averaging (RANS) methods
methods and
and can
can only
only be applied to
be applied to particles
particles smaller
smaller than
than thethe
CFD cell
CFD cellsize.
size.InInunresolved
unresolvedCFD-DEM,
CFD-DEM, some
some empirical
empirical equations
equations are are
usedused to calculate
to calculate the
the fluid forces acting on the particles and the calculated fluid forces are included
fluid forces acting on the particles and the calculated fluid forces are included as additional as ad-
ditional terms in governing DEM equations. In contrast, resolved CFD-DEM
terms in governing DEM equations. In contrast, resolved CFD-DEM (particle resolved (particle re-
solved DNS) can be applied to larger particles than the CFD cell size, where
DNS) can be applied to larger particles than the CFD cell size, where extreme fine meshingextreme fine
meshing
is used to is useddetailed
obtain to obtain detailed information
information on flow
on turbulence turbulence flowacting
and forces and forces actingare
on particles on
directly obtained by integrating fluid stress over the particle surface. In resolved CFD-DEM,
various techniques such as Adaptive mesh refinement (AMR) and Immersed boundary
method (IBM) are becoming popular, but are limited to a minimal number of particles [5].
Additionally, the particles can be assumed as point-particles (PP), representing point objects
Fluids 2022, 7, 337 3 of 26

with a certain mass and whereupon DNS can be performed. Unlike resolved CFD-DEM,
here the forces acting on the particles are calculated using empirical equations, but the
application of PP DNS-DEM to particles larger than Kolmogorov scale is questionable and
highly discouraged.
DNS-DEM studies [6,7] are limited to simple flows, a small number of particles, and
mainly to the PP approach due to heavy computational requirements. DNS performed
on PP in rough-wall pipe with a small Reynold’s number indicates that particle volume
fraction (αp ) and Stokes number (St) play an important role in turbulent modification [8].
Recently, a two-way coupled DNS study on particle-laden flow highlighted the effects of
the particle Stokes number (St) on near-wall turbulence [9]. Resolved CFD-DEM (particle
resolved DNS) is only possible if the spacing of the computational grid is small compared
to the size of a particle, therefore restricting its application to large particles compared to
the smallest scales of the turbulent flows and/or relatively small number of particles [10].
Even for single-phase fluid systems, DNS is not possible for the cases with high Reynolds
numbers and complex geometries due to computational limitations.
RANS-DEM is another economical approach where the mean flow fields are obtained,
and additional dispersion models incorporate the turbulence effect on the particle’s tra-
jectories. An accurate evaluation of instantaneous velocity fluctuations is required for
the realistic evaluation of turbulent diffusion effects for accurate predictions of particle
dispersion and deposition on surfaces [11,12]. A number of RANS-DEM studies on simple
cases [13,14] indicate that these simple dispersion models cannot accurately obtain the lost
fluctuating component due to averaging of NS equations. In contrast, Greifzu et al. [15]
showed that the simple dispersion models are indeed able to predict correct particle dis-
persion even for more complex flow (particle-laden BFS flow). However, we found out
that their conclusion was due to an error in the numerical setup (refer to the results and
discussion section for details). Therefore, further investigation is necessary to reach a
unanimous conclusion about the ability of the simple dispersion models in incorporating
the effect of turbulent fluctuations on particle motion.
A sensible and efficient approach would be LES-DEM, which might be a good compro-
mise between accuracy and computational feasibility. However, one also has to be cautious
about the sub-grid scales (SGS) fluid fluctuating motion seen by the particles, because in
several investigations, the effects of SGS on particle motion were shown to be significant
and hence should not be neglected [16–19], especially when the particle response time is
the same order of magnitude as that of the smallest time scale resolved in LES. To recover
the dynamic consequences of the SGS in the LES-DEM framework, stochastic models such
as an explicit stochastic forcing in the equations of particle motion were suggested [20,21].
Furthermore, the size of LES meshes in unresolved CFD-DEM is restricted by the require-
ment of particles being significantly smaller than the CFD cell size, unless particles are
considered as PP. This restriction prevents finer meshes near the boundaries (y+ ~1) and
can lead to poor performance of the static LES turbulence model, which require very fine
boundary meshes. Dynamic LES turbulence models could be adopted to avoid the poor
performance of static LES models in cases of low mesh resolutions.
On one hand, commercial software such as Fluent EDEM, STAR-CCM+ and AVL-
fire [22–26], in-house CFD programs such as LESOCC [27], and research codes [28–31] can
be applied to CFD-DEM simulations, but the accessibility of these solvers is limited. On
the other hand, some open-source CFD codes, such as OpenFOAM [32] have accelerated
research in the field. Some coupled CFD-DEM codes [33,34] are also available, which
couple OpenFOAM and DEM software such as LAMPS/LIGGGHTS and are not limited
to PP.
Particle-laden backward facing step (BFS) flow is popular among researchers in the
field due to its simple geometry and ability to produce interesting turbulent features
concerning flow separation and re-attachment. A few LES-DEM simulations on particle-
laden BFS have been performed [35–40], and some have attained a reasonable agreement
with the experiment. These LES-DEM studies used extreme fine meshing (y+ ~1) and were
Fluids 2022, 7, 337 4 of 26

mainly based on the PP approach. The particle-laden BFS flow adopted in our study
had previously been numerically simulated by Greifzu et al. [15] in the context of RANS-
DEM. The authors concluded that RANS-DEM (and the simple dispersion model therein)
predicts the correct fluid and particle velocity profiles, as well as the particle dispersion.
Interestingly, we found that their OpenFOAM numerical setup was erroneous as a fluid
density of 1000 kg/m3 (water) instead of 1 kg/m3 (air) was used, although in the original
experiment the fluid was air, not water. Despite using an incorrect fluid density value,
the authors obtained an excellent agreement with the experimental data. However, these
results are questionable, as the density of the fluid should have been 1 kg/m3 as in the
original experiment—therefore requiring a reinvestigation.
Despite several experimental and numerical studies, fluid-particle systems remain
poorly understood due to the complex physics involved, such as turbulent modulation and
complex fluid-particle and particle-particle interaction. The CFD-DEM method could pro-
vide detailed insights into these multivariable and interdependent phenomena and can be
employed for large scale engineering applications. However, due to the huge computational
requirements and associated limitations of the CFD-DEM method, finding the trade-off
and compromise between the levels of flow resolution obtained (DNS/LES/RANS) and
the required computational efforts to predict the correct particle dispersion and trajectories
is difficult.
Here, we focus on different aspects of modelling such flows in terms of the compu-
tational requirements, the available models, as well as the challenges and limitations. In
particular, we perform RANS-DEM and LES-DEM simulations in 3D for particle-laden
BFS flow. Special attention is given to the ability of the respective approaches to predict
particle dispersion, coupling regime, particle boundary conditions, and turbulence mod-
elling. First, we discuss the theoretical background of the Eulerian–Lagrangian method
(CFD-DEM) in detail, focusing on the RANS and LES methods for resolving the fluid flow
fields. The following method and numerical setup section highlights the fundamental
structural differences in the numerical setup for different approaches adopted in our study.
Furthermore, the RANS-DEM and LES-DEM simulation results for the fluid and particle
phases are compared, especially in relation to particle dispersion. Additionally, different
aspects of the CFD-DEM method with regard to the coupling regime and particle bound-
ary conditions were investigated and their effects on fluid and particle phase results are
discussed. Before including particles into the system, single-phase RANS and LES simula-
tions are also performed and its accuracy in predicting mean and turbulent flow statistics
with different turbulence models are assessed. Taken together, RANS-DEM requires more
sophisticated dispersion models to predict correct particle dispersion, and LES-DEM has
limitations in terms of the flow resolution obtained, the computational resources required,
and the prerequisites of unresolved CFD-DEM, preventing extreme fine meshing unless
the particles are considered as PP.

2. CFD-DEM Approach and the Governing Equations


The unresolved CFD-DEM approach was adopted to investigate the two-phase system
containing fluid as a continuum and the particles as discrete mediums. The full NS
equations describe the continuum fluid phase for unsteady incompressible flow, which
is a slightly modified version of the standard NS equations to incorporate the particle
fraction in each computational cell. Newton’s second law of motion describes the discrete
particle phase.
∂α ∂αui
+ = 0, (1)
∂t ∂xi
p
∂αui ∂αui 1 ∂αp ∂2 αui f
+ uj =− +υ + αgi − i , (2)
∂t ∂x j ρ ∂xi ∂x j ∂x j ρ
p(k) nc(k)
∂ui

c(kl ) f (k) g(k)
m(k) = Fi + Fi + Fi , (3)
∂t l =1
Fluids 2022, 7, 337 5 of 26

p(k) nc(k)
∂wi

c(kl )
I (k) = Mi , (4)
∂t l =1
p(k)
dxi p(k)
= ui . (5)
dt
where:
α = volume fraction of fluid in each cell; unitless
ui = fluid velocity field in direction i; m/s
p = fluid pressure; N/m2
υ = kinematic viscosity of fluid; m2 /s
gi = gravitational acceleration in direction i; m/s2
ρ = density of fluid; kg/m3
m(k) = mass of particle k; kg
nc(k) = number of particles colliding with particle k; unitless
I (k) = moment of inertia of particle k; kgm2
p(k)
ui = velocity of particle k in direction i; m/s
p(k)
wi = angular velocity of particle k in direction i; 1/s
f (k)
Fi = surface forces acting on particle k (including drag,
lift, added-mass, basset
history forces etc.: coupled forces); N
p
f i = volumetric fluid-particle interaction momentum source in direction i; N/m3
g(k)
Fi = body forces acting on particle k; (gravity + buoyancy: uncoupled forces)
 
= m ( k ) gi 1 − ρ p ; N
ρ

ρ p = density of particle; kg/m3


c(kl )
Fi = particle-particle interaction/contact force between particle k and l; N
c(kl )
Mi = particle-particle interaction moment between particle k and l; Nm
x and t are space and time with units m and s, respectively.
c(kl )
Mi = particle-particle interaction moment between particle k and l; Nm
x and t are space and time with units m and s, respectively.
OpenFOAM considers the particles as point-particles (PP), meaning that they are
represented as point objects having a certain mass. This assumption automatically neglects
the torque acting on the particles, meaning that OpenFOAM does not consider Equation (4)
in calculating the trajectories of the particles.
In the above equations, the momentum transfer term consists of several forces coupled
between the continuum phase and discrete phase, such as drag force, lift force, pressure
gradient force, basset history force, added-mass force, etc. In the references, it has been
established that the major contribution in the momentum transfer term originates from the
drag force [41], and lift force is more relevant for light particles than heavy particles [10].
The particles considered in the present study are dense copper particles. Therefore, the lift
force seems to be insignificant for our case. However, we have also included the pressure
gradient force in addition to the drag force in our numerical setup. Ultimately, the coupled
forces term reduces to:
f (k) D (k) PG (k )
Fi = Fi + Fi , (6)

D (k) 3 ρ m(k) 
s(k) p(k)

s(k) p(k)
Fi = C D u i − u i ui − ui , (7)
4 ρ p d(k)
  2

24 1 p ( k )

p(k) 1 + 6 ( Re ) 3
when Re p(k) ≤ 1000
CD = Re , (8)
0.424 when Re p(k) ≥ 1000

s(k) p(k)
d (k ) ui − ui
p(k)
Re = , (9)
υ
Fluids 2022, 7, 337 6 of 26

s(k)
PG (k ) π (k) 3 Dui
Fi = (d ) ρ . (10)
6 Dt
From the equations above, one can see that to calculate forces acting on particles
(thus to calculate their trajectories), information is needed on the fluid velocity at the
location of particle (ui s(k) ). We obtain this information from the fluid phase solution (CFD).
The solution of fluid phase can be categorized into three types, namely Direct Numerical
Simulation (DNS), Large-Eddy Simulation (LES), and Reynolds-averaged Navier–Stokes
equations (RANS), depending on the level of flow resolution needed and the computational
resources available.

2.1. Direct Numerical Simulation (DNS)


DNS solves the full NS equations numerically, thus resolving everything from the
largest scale to the smallest dissipative eddies present in the system. In DNS under
consideration of the point-particles (PP) approach, the velocity of the fluid at the particle
location can be obtained directly from the DNS solution.

ui s(k) = ui DNS ( x p(k) (t), t) = ui ( x p(k) (t), t). (11)

Since turbulent flows possess a varying range of time and length scales, the exact
solution (DNS), even for the simplest turbulent flows, requires enormous computational
resources and extreme fine meshing. Initial estimation of the computational resources
required for DNS can be made based on Kolmogorov scales (smallest time, length and
velocity scales) in the system. In our case, the Kolmogorov length scale is about 170 µm,
meaning the mesh resolution should be smaller than 170 µm for DNS. It has been demon-
strated that the restrictions that DNS needs for simple channel flow in terms of grid point
and time steps [42], thus require huge computational resources even for simple channel
flow. Computational resources requirement by DNS in the sense of both processor speed
and memory size for storing intermediate results is vast and increases exponentially with
the Reynolds number of the flow. In order to obtain the maximum possible information
about the flow fields with an affordable computational cost, the full NS equations are
approximated with some averaging/filtering approaches. The resulting averaged/filtered
NS equations have almost the same form as the original NS equations with additional
terms, which can be calculated based on eddy viscosity.

2.2. Large-Eddy Simulation (LES)


LES aims to resolve large-scale turbulence while small-scale turbulence is modelled
using the filtering operation. Compared to DNS, where nearly all the computational effort
is used to resolve the smallest dissipative eddies, LES resolves only up to the inertial
subrange, not all the way to the dissipative scales. This can save significant computational
effort, yet preserves enough information of the fluid flow. LES converges to DNS when
finer meshing and smaller time steps are used.
The common idea behind LES is to decompose the instantaneous flow field u( x, t) into
resolved (or filtered) component u( x, t) and residual (or sub-grid scale; SGS) component
u0 ( x, t) by a filtering operation, as follows:

ui ( x, t) = ui ( x, t) + ui 0 ( x, t), (12)
Z
ui x 0 , t G x, x 0 ; ∆ dx 0 ,
 
ui ( x, t) = (13)

where, G ( x, x 0 ; ∆) is the filter function that depends on mesh discretization. The filtering
operation results in extra terms, called residual stresses (τijR ) in the original NS equations.
Calculation of residual stresses is based on the Boussinesq hypothesis of eddy viscosity
(turbulence viscosity υT ), which is being modelled. Various models are available for
this purpose, such as Smagorinsky, one-equation model (kEqn), dynamic Smagorinsky,
Fluids 2022, 7, 337 7 of 26

dynamicKEqn, Spallart–Allmaras, and many others. We found that the dynamicKEqn


turbulence model was able to predict correct flow fields in terms of mean and fluctuating
components with reasonable accuracy, whereas static turbulence models were performing
poorly with provided mesh resolution.
In cases where the particle response time (Stokes number) is large compared to the
smallest time scale resolved in LES, the fluid velocity of the sub-grid scales does not
significantly influence the particle’s motion [10,43,44]. Considering this, one does not
need an extra dispersion model to incorporate the effect of turbulence in the particle’s
motion; thus, the LES solution can be directly equated to the fluid velocity at the location
of the particle.
ui s(k) = ui LES ( x p(k) (t) , t) = ui ( x p(k) (t) , t). (14)

2.3. Reynolds-Averaged Navier–Stokes Equations (RANS)


RANS resolve only the mean flow statistics; thus, RANS solution for fluid flow fields
cannot be directly equated with the fluid velocity at the location of the particle. In RANS,
the instantaneous flow field u( x, t) decomposes into a time average component hu( x, t)i
and a fluctuating component u0 ( x, t):

ui ( x, t) =< ui ( x, t) > +ui 0 ( x, t), (15)


Z T
1
< ui ( x, t) >= lim ui ( x, t) dt. (16)
T →∞ T 0

The averaging operation results in some new terms, < ui0 u0j >, called Reynolds stresses,
to appear in the original NS equations, which are also modelled based on eddy viscosity.
The terms < ui0 u0j >, although named as Reynolds stresses, have a unit of stress only when
multiplied by the fluid density ρ. Similar to LES, eddy viscosity υT can be calculated based
on several models such as k-ε, k-ω, k-ω SST and many others. We used the k-ω SST model
to close the Reynolds-averaged NS equations of our RANS and RANS-DEM simulations,
as it was the best performing.
In RANS, the effect of fluctuating components (turbulence) on particles is incorpo-
rated using some dispersion models [45]. The resulting fluid velocity at the location of
a particle (ui s(k) ) can be equated to the sum of the RANS (mean) velocity field and the
modelled fluctuations.

ui s(k) = ui RANS ( x p(k) (t) , t)+ui 0 =< ui ( x p(k) (t) , t) > + ui 0 . (17)

All three approaches (DNS, LES, and RANS) for calculating the fluid velocity at the
location of particles have their limitations in terms of accuracy and computational cost
and can be adopted as per the details required and computational resources available.
Figure 2 shows the extent of modelling certain types of turbulent models [46], where
DNS resolves everything from the largest to the smallest dissipative eddies present in
the system. LES resolves up to energy-containing eddies while dissipative eddies are
modelled. RANS resolves only the mean flow statistics, and the fluctuating components are
modelled. More information on the implementation of the turbulence models used in our
RANS-DEM and LES-DEM and the dispersion models can be found on the OpenFOAM
webpage [32]. We have used RANS-DEM and LES-DEM approaches to solve particle-laden
BFS flow and investigated the case by comparing the fluid and particle-phase results with
the experimental data and literature.
RANS-DEM and LES-DEM and the dispersion models can be found on the OpenFOAM
webpage [32]. We have used RANS-DEM and LES-DEM approaches to solve particle-
laden BFS flow and investigated the case by comparing the fluid and particle-phase results
Fluids 2022, 7, 337 with the experimental data and literature. 8 of 26

Figure 2. Extent of modelling for certain types of turbulent flows [46].


Figure 2. Extent of modelling for certain types of turbulent flows [46].
2.4. Particle Dispersion Modelling
2.4. Particle Dispersion Modelling
The approximated NS equations provide mean (RANS)/filtered (LES) flow statistics
and information
The approximated about turbulent
NS equations fluctuations
provide mean are lost due to simplifications
(RANS)/filtered (LES) flow of statistics
original
f (k)
and information about turbulent fluctuations are lost due to simplifications
NS equations. The fluid-particle interaction forces (Fi ) such as pressure gradient force, of original NS
𝑓(𝑘)
equations.
added massThe fluid-particle
force, drag force interaction
and historyforces force, (𝐹 contain
𝑖 ) such as pressure
unfiltered componentsgradient andforce,
are
required
added mass to beforce,
estimated
dragtoforce
closeand certain terms
history in particle
force, contain equations
unfiltered of motion.
components In LES-DEM,
and are
the sub-grid
required to bescales only have
estimated a small
to close effectterms
certain on the inparticle’s trajectories,
particle equations of especially
motion. Inwhen LES-
the particle response time is large compared to the typical time
DEM, the sub-grid scales only have a small effect on the particle’s trajectories, especially scales of the turbulent
flow
whenand the the smallest
particle time time
response scaleisresolved in LES. to
large compared However,
the typicalthetime
sub-grid
scalesscales
of thecan be
turbu-
significant
lent flow and in many physicaltime
the smallest processes such as in
scale resolved inturbophoresis
LES. However, [17].
theOn the other
sub-grid hand,
scales canthebe
effect of turbulent
significant in many fluctuations must be included
physical processes such as in in turbophoresis
RANS-DEM to[17]. predict
On realistic
the other particle
hand,
trajectory.
the effect of There are mainly
turbulent two types
fluctuations must of be
modelling
includedapproaches
in RANS-DEM to account for missing
to predict realistic
turbulent
particle trajectory. There are mainly two types of modelling approaches to account[47]
fluctuations: either by adding stochastic noise forcing to the NS equations for
or by adding
missing an additional
turbulent velocity
fluctuations: to the
either particlestochastic
by adding equation of motion
noise [48].toThe
forcing thestochastic
NS equa-
models,
tions [47]which
or by aim
adding to recover the lostvelocity
an additional turbulence to theeffects
particleon equation
a particle’s trajectories,
of motion [48]. can
The
be formulated based on (a) transport equations of the turbulent
stochastic models, which aim to recover the lost turbulence effects on a particle’s trajecto- kinetic energy (Discrete
Random
ries, can Walk; DRW, or Eddy
be formulated basedInteraction Model; equations
on (a) transport EIM) or (b)of generalized
the turbulent Langevin
kinetic equation
energy
(Continuous Random Walk; CRW). We have considered the
(Discrete Random Walk; DRW, or Eddy Interaction Model; EIM) or (b) generalized Lange- DRW/EIM stochastic model in
our numerical simulation. Here, particles are assumed
vin equation (Continuous Random Walk; CRW). We have considered the DRW/EIM sto- to be trapped by an eddy during
its lifetime,
chastic model resulting
in our in the meansimulation.
numerical flow fieldsHere,seen by the particles
particles being those
are assumed to be of the fluid
trapped by
and the fluctuating components, which are randomly distributed
an eddy during its lifetime, resulting in the mean flow fields seen by the particles being following Gaussian
distribution, whose
those of the fluid root
and themean square values
fluctuating components,are equal whichandarededuced
randomly fromdistributed
the turbulent fol-
kinetic energy.
lowing Gaussian distribution, whose root mean square values are equal and deduced
In OpenFOAM, two dispersion models are available to model the turbulent fluctua-
from the turbulent kinetic energy.
tions ui 0 , namely StochasticDispersionRAS and GradientDispersionRAS models, which are
basically DRW/EIM type of stochastic dispersion models. These dispersion models use the
turbulent kinetic energy (TKE; k) provided by the RANS solution to model the turbulence
fluctuations such that, r
0 2
ui = kx e (18)
3 rnd i
where xrnd is a random factor that reproduces a probability density function
q with Gaussian
2
distribution with an expected value µ = 0 and standard deviation σ = 3 k. ei is a unit
Fluids 2022, 7, 337 9 of 26

vector that points either in a random direction (in StochasticDispersionRAS) or in −∇k (in
GradientDispersionRAS).

2.5. Coupling between Continuum and Discrete Phases


As shown in Figure 1, the continuum and discrete phases interact with each other and
are required to be coupled. Several studies have emphasized on one-way and two-way
coupling [49–52] for dilute particle-laden flows and fluid-structure interaction. Depending
on the coupling regime, specific terms vanish in the respective governing equations. The
coupling regime required for numerical simulations can be made using the volumetric
particle concentration (αp ).
• One-way coupling (fluid → particle): when the volumetric concentration of particles
is small (αp < 0.0001%), the fluid flow fields affect the particle motion, but particles
have a negligible effect on fluid flow fields. This results in only specific terms being
considered in the governing CFD-DEM equations as follows:

f (k) p c(kl )
Fi 6= 0; f i = 0; Fi = 0. (19)

• Two-way coupling (fluid ←→ particle): when the volumetric concentration of par-


ticles increases (0.1% < αp < 0.0001%), both fluid and particles affect each other’s
motion. Two-way coupling can be further categorized into two categories, one in
which particles enhance the turbulence dissipation and a second in which particles
enhance turbulence production, which depends on the ratio of particle reaction time
(τp = ρ p d2p /18ρυ) to the Kolmogorov time scale (τk = ( υε )1/2 ) and to the turnover time
of large eddies (τe = l/u), respectively, where ρ p is the density of particle, d p is the
diameter of particle, ρ is the density of fluid, υ is the kinematic viscosity of fluid, ε
is turbulence dissipation rate, l is turbulence length scale, and u is the fluid velocity.
Two-way coupling results in the following CFD-DEM equations:

f (k) p c(kl )
Fi 6= 0; f i 6= 0; Fi = 0. (20)

• Four-way coupling (fluid ←→ particle, particle ←→ particle): When the volumetric


concentration of the particles further increases (αp > 0.1%), the interaction among
particles becomes significant. In this regime, fluid and particle affect each other’s
motion; additionally, the particle collision term needs to be included in the govern-
ing equations:
f (k) p c(kl )
Fi 6= 0; f i 6= 0; Fi 6= 0. (21)

3. Method and Numerical Setup


The original experiment [53] includes a blower and particle feeders, where the blower
injects fluid (air), and the particle feeders feed the particles into the system at a specific
mass flow rate. This arrangement provides uniform fluid velocity and particle loading
at the inlet of the development section, which has a length of 5.2 m, ensuring that the
turbulent flow is fully developed at the inlet of the test section (backward facing step; BFS)
and gives the particles enough time/length to become uniformly mixed with the fluid flow
and attain equilibrium with the fluid phase before it reaches the inlet of BFS. The fully
developed turbulent flow has a centerline velocity of 10.5 m/s, and the Reynolds numbers
based on it are 13,800 and 18,400 at the channel section (based on channel half-height h)
and at the step (based on step height H), respectively. The geometry of the test section, fluid
(air) and particle-phase description in the experiment and their corresponding adoption for
numerical simulations are shown in Table 1.
Fluids 2022, 7, 337 10 of 26

Table 1. Geometry, fluid, and particle phase characteristics in the original experiment [53] and their
corresponding setting for the numerical setup.

Experimental Setup Our Numerical Setup


Channel half-height, h = 20 mm Channel half-height, h = 20 mm
Channel span, Lu = 457 mm Channel span, Lu = 457 mm
Geometry Step height, H = 26.7 mm Expansion channel length, Ld = 935 mm
Expansion ratio (H + 2h/2h) = 5:3 Step height, H = 26.7 mm
Aspect ratio (B/H) = 17:1 Width (B) = 114.25 mm
Centerline velocity, U0 = 10.5 m/s
Friction velocity, uτ = 0.5 m/s Density (ρ) = 1.225 kg/m3
Viscous length scale = 31 µm Kinematic viscosity (υ) = 1.5 × 10−5 m2 /s
Dissipation,  (centerline estimated) = 4.3 m2 /s3 Centerline velocity (U0) at
Continuum phase (air)
Kolmogorov length scale, η Inlet = (10.5 0 0) m/s
(Centerline estimated) = 170 µm Average velocity (Uavg) at
Large eddy time scale, inlet = (9.39 0 0) m/s
τf = 5H/U0 = 12.7 ms
Nominal diameter, d p = 70 µm
Material = copper
Number mean diameter = 68.2 µm
Standard deviation of diameter = 10.9 µm
Particle size distribution (PSD): Normal
Density, ρ p = 8800 kg/m3
distribution with expected value of 68.2 µm and
Mass flow rate of particle
Discrete phase (copper . . standard deviation of 10.9 µm
(Mass loading), m p /m f = 10%
particle) Density, ρ p = 8800 kg/m3
Stokes mean particle time constant,
Mass flow rate of particle (mass loading),
τp, stokes = (2ρ p + ρ)d2p /36µ= 130 ms . .
m p /m f = 10%
Modified mean  particle time constant,

τp = τp, stokes / 1 + 0.15Re0.687
p = 88 ms
d p Urel
Particle Reynolds number, Re p = υ = 5.5

The geometry considered (test section; BFS) for numerical simulation can be seen in
Figure 3. Due to computational limitations, the development channel before the inlet of BFS
is not considered in the geometry. To achieve a fully developed turbulent flow at the inlet of
BFS without providing a development channel, mapped boundary conditions are applied,
in which flow fields are mapped from 400 mm downstream of the inlet of BFS, resulting in
a fully developed fluid velocity profile at the inlet of BFS with a centerline velocity (U0)
of 10.5 m/s and average velocity (Uavg) of 9.39 m/s. Regarding the particle boundary
condition in OpenFOAM, in addition to the mass flow rate (mass flux) of particles, one
also needs to provide a particle injection velocity. In the original experiment, particle
velocity was not measured at the inlet of the BFS; thus, no data is available for specifying
boundary conditions at the model inlet concerning particle injection velocity. We have
tested several boundary conditions for the particle injection velocity. Assuming that all the
particles have attained a constant velocity (injection velocity) as they reach the inlet of the
BFS, two extreme bounds of particle injection velocity are tested, where the particles are
injected with (10.5, 0, 0) m/s (upper bound) and (0, 0, 0) m/s (lower bound) of injection
velocity. Furthermore, we assume that the particles have attained a velocity that follows
the mean fluid velocity profile at the inlet, which is also tested (i.e., particles injected at the
center will have an injection velocity similar to the centerline velocity of the fluid, whereas
particles injected near to the walls will have almost zero injection velocity due to the no-slip
boundary condition for the fluid phase). The options mentioned above for obtaining the
best approximation of particle boundary conditions are tested against both the approaches
adopted in our study to simulate particle-laden BFS flow.
injection velocity. Furthermore, we assume that the particles have attained a velocity that
follows the mean fluid velocity profile at the inlet, which is also tested (i.e., particles in-
jected at the center will have an injection velocity similar to the centerline velocity of the
fluid, whereas particles injected near to the walls will have almost zero injection velocity
due to the no-slip boundary condition for the fluid phase). The options mentioned above
Fluids 2022, 7, 337 for obtaining the best approximation of particle boundary conditions are tested against 11 of 26
both the approaches adopted in our study to simulate particle-laden BFS flow.

Figure 3. The geometry considered for numerical simulation.


Figure 3. The geometry considered for numerical simulation.

RANS-DEM and LES-DEM simulations are performed in 3D. In the first step, RANS-
RANS-DEM and LES-DEM simulations are performed in 3D. In the first step, RANS-
DEMDEM simulations
simulations were
wereperformed
performed for foraa10%
10% mass
mass loading
loading of ~70ofµ~70 µm copper
m copper particlesparticles
for for
one-way and two-way coupling. Although the volumetric particle −5 (O ~ 10−5 )
fraction
one-way and two-way coupling. Although the volumetric particle fraction (O ~ 10 ) lies
lies in therange
in the range of two-way
of two-way coupling,
coupling, we havewealsohave also one-way
considered considered one-way
coupling coupling in
in addition
to two-way coupling under RANS-DEM to facilitate the comparison
addition to two-way coupling under RANS-DEM to facilitate the comparison between between them. Inter-
them.estingly, our RANS-DEM
Interestingly, results almost
our RANS-DEM overlap
results for one-way
almost overlap coupling and two-way
for one-way couplingcou- and two-
pling, indicating that it should be enough to include one-way coupling, even when the
way coupling, indicating that it should be enough to include one-way coupling, even when
particle concentration is slightly greater than the threshold suggested by Elghobashi [4].
the particle
However, concentration
we decided toisonlyslightly greater than
use two-way the for
coupling threshold suggested
our LES-DEM by Elghobashi
simulations for [4].
However, we decided to only use two-way coupling for our LES-DEM
10% mass loading of ~70 µ m copper particles due to the fact that the particle volume frac-simulations for 10%
masstion loading
lies inof ~70ofµm
range coppercoupling
two-way particlesanddue
the to
CPUthetime
factfor
that
thethe
caseparticle
was roughlyvolume the fraction
lies insame
range of two-way
for one-way coupling
and two-way and the
coupled CPU time
RANS-DEM for the case
simulations (~10was roughly
min more). the same for
RANS-
one-wayDEMand and LES-DEM
two-wayhave basic RANS-DEM
coupled structural differences in their (~10
simulations numerical
min setup
more). as they both
RANS-DEM and
require different input parameters, depending on the models used to close the aver-
LES-DEM have basic structural differences in their numerical setup as they both require
aged/filtered NS equations. We have used the k-ω SST (kOmegaSST) and dynamicKEqn
different input parameters, depending on the models used to close the averaged/filtered
turbulence models in the case of RANS-DEM and LES-DEM, respectively. We have en-
NS equations.
sured that the Wesingle-phase
have usedresults
the k-ωmatchSST
the(kOmegaSST)
experimental data and dynamicKEqn
before the particles are turbulence
models in the case of RANS-DEM and LES-DEM, respectively. We
included in it. In our single-phase LES simulations, we have tested the predictive capabil- have ensured that
the single-phase resultsmodel
ity of static turbulence match(kEqn)
the experimental
as well. Once thedata beforeofthe
correctness the particles
single-phase are re-included
in it.sults has been
In our verified, theLES
single-phase casessimulations,
were modifiedwe to incorporate
have tested particles and solved using
the predictive capability of
the DPMFoam solver. Although DPMFoam is based on the discrete parcel
static turbulence model (kEqn) as well. Once the correctness of the single-phase results method (DPM),
we have considered only one particle in each parcel, therefore it is equivalent to the
has been verified, the cases were modified to incorporate particles and solved using the
DPMFoam solver. Although DPMFoam is based on the discrete parcel method (DPM), we
have considered only one particle in each parcel, therefore it is equivalent to the discrete
element method (DEM). More details about the RANS-DEM and LES-DEM case setup can
be found in Table 2.
As mentioned in the introduction, unresolved CFD-DEM requires the CFD mesh
to be much larger than the particle size. This is assured as the smallest CFD mesh cell
size is 0.2 mm and 0.15 mm in our RANS and LES setup, respectively, which is much
greater than the particle size (~70 µm), thus considering them as point-particles (PP) is
justified. To resolve the interesting flow features developing near walls, mesh grading is
performed, providing us with the flexibility to provide larger cells away from the wall,
which saves some additional computational efforts. Mesh is also refined in streamwise
direction near the step, whereas uniform mesh is used in spanwise direction. Before we
finalize our final mesh, the mesh is refined in a stepwise manner until we obtain almost the
same results for fluid and particle velocity profiles between consecutive refinements (grid
independence). The selection y+ as 3 instead of 1 in our LES-DEM simulations allowed us
to keep the particle size significantly smaller than the CFD cell (PP approach), and extreme
fine meshing is avoided due to computational limitations and the resources available. In the
review published on LES simulations [54], it has been reported that dynamic LES models
are expected to perform better than static models. Our investigation on LES turbulence
models also showed that static LES turbulence models such as Smagorinsky or kEqn
models cannot predict correct fluid velocity flow fields with the provided mesh resolution
(y+ ~3). Therefore, we used the dynamicKEqn model in our LES-DEM simulations, which
Fluids 2022, 7, 337 12 of 26

seems to overcome the poor performance of static turbulence models when using relatively
coarse mesh resolution. However, standard DPMFoam solver does not come with dynamic
turbulence models in OpenFOAM-v1912 (the solver used in our RANS-DEM and LES-
DEM study), and we needed to compile the dynamicKEqn model as a separate library for
two-phase systems and included it in our LES-DEM simulations.

Table 2. RANS-DEM and LES-DEM settings.

Description RANS-DEM LES-DEM


Solution domain 3-Dimension 3-Dimension
dynamicKEqn with
Turbulence model kOmegaSST
cubeRootVolume delta function
Dispersion model StochasticDispersionRAS -
Mesh resolution 1,105,280 hexahedra cells 3,533,376 hexahedra cells
y+ - ~3
Resolved TKE - 80–90%
boundary condition at mappedPatch (Mapped from mappedPatch (Mapped from
inlet (air) 400 mm downstream of inlet) 400 mm downstream of inlet)
Front and back
Cyclic Cyclic
boundary treatment
Wall treatment Wallfunctions Resolved
(10.5, 0, 0) m/s (10.5, 0, 0) m/s
Particle injection velocity
(0, 0, 0) m/s (0, 0, 0) m/s
at inlet (particle
Varying as per fluid velocity Varying as per fluid velocity
boundary condition)
distribution distribution
Mass loading of particles 10% 10%
Coupling regime One-way and two-way Two-way
Simulation duration 1s 3s

OpenFOAM solvers, namely pimpleFoam and DPMFoam, are used for single and two-
phase simulations, respectively, in OpenFOAM-v1912. Both pimpleFoam and DPMFoam
use the pimple algorithm to couple velocity and pressure fields. Backward and least Squares
schemes are used for time and gradient discretization, respectively. All the divergence
terms are discretized using Gauss linear method. The resulting discretized equations were
solved using algebraic multigrid (AMG) and algorithms based on a point-implicit linear
equation solver (Gauss–Seidel). DEM data (particle position, velocity, etc.) are mapped
onto CFD mesh, and particle volume fraction in each computational cell is calculated. The
interaction forces are locally averaged in each cell and incorporated in NS equations and
the calculated flow data are communicated back to the DEM side. All the simulations were
performed in parallel on 56 processors in the Linux cluster of Leibniz Supercomputing
Centre (LRZ). The total CPU computational time corresponding to different simulations
can be found in Table 3.

Table 3. Total CPU computational time.

Total CPU Computational Time (in Seconds)


Simulation
RANS-DEM (Run Time = 1 s) LES-DEM (Run Time = 3 s)
Single-phase 3925 29,682
Two-phase (one-way coupled) 12,693 -
Two-phase (two-way coupled) 13,385 109,921
Fluids 2022, 7, 337 13 of 26

Fluid and particle velocity profiles at the measurement locations are compared with
the experimental data. Fluid flow fields can be directly extracted and plotted at the
measurement locations from the OpenFOAM results. Time averaging on particles cannot
be performed as done for the continuum phase (air) due to the discrete nature of the
particles. The particles are evaluated in a slice around the measurement locations with a
thickness of 0.15H and the center lies precisely at the measurement locations for every 0.1 s
interval from the start to the end of the simulation. The slice thickness of 0.15H is adopted
as the same slice thickness was considered for sampling the particles in the previous
study [15]. The particle data collected for all the selected time intervals are combined,
and averaging is performed in each slice on the particles with the same location (only the
y-component). The resulting average particle velocity profiles are then compared with the
experimental data.

4. Results and Discussion


The results shown below are arranged in such a way that they indicate the workflow
adopted to investigate and solve the particle-laden BFS flow in OpenFOAM. The normal-
ized mean and fluctuating fluid velocity profiles are compared with the experimental data
at their respective measurement locations (x/H = 2, 5, 7, 9, 14). Normalized average particle
velocity profiles are compared with observed data in the experiment at their respective
measurement locations (x/H = 2, 5, 7, 9, 12) to assess the particle-phase results. We have
somewhat amplified the normalized velocity fields in order to highlight the deviations.

4.1. Single-Phase RANS and LES


To obtain correct results for discrete phase (particles), one must have acceptable results
for the continuum phase (air) with reasonable accuracy. Therefore, we first performed the
single-phase RANS and LES simulations using the pimpleFoam solver, which is able to
predict fluid velocity profiles with acceptable precision. For RANS and LES results, the
methods and formulae used to calculate the fluctuating components (Urms) are different.
Regarding the RANS models, for everything above mean flow fields, the fluctuating
component is calculated (with the assumption of isotropic turbulence holds true) directly
from the modelled turbulent kinetic energy (TKE; k) using Equation (22), which is the
modelled part of the fluctuating component.
r
2
Urms = k (22)
3
In LES simulations, we aim to resolve 80–90% TKE, and the calculated Urms represents
the resolved fluctuating component. Urms is calculated by subtracting the mean flow field
from the instantaneous flow fields, as shown below:
v
u1 N
u
Urms = t ∑ (Ui − U Mean)2 (23)
N i =0

N
1
U Mean =
N ∑ Ui (24)
i =0

where, Ui is the instantaneous velocity, U Mean is the time-averaged velocity, and N is the
total number of time steps.
Figure 4 compares simulated streamwise air mean and fluctuating velocity profiles
under RANS (with kOmegaSST turbulence model) and LES (with kEqn and dynamicKEqn
turbulence models) frameworks with the experiment data. It is evident from the streamwise
air mean velocity profile plots (Figure 4a) that both RANS and LES (with dynamicKEqn
turbulence model) predict the correct mean flow and are able to predict the re-attachment
point (x/H ~ 7) quite accurately as observed in the original experiment (x/H ~ 7.4). On the
Fluids 2022, 7, 337 14 of 26

other hand, Figure 4b shows that the streamwise air fluctuating components (Urms) from
RANS are somewhat underpredicted, which is not surprising as they represent only the
modelled part of the turbulent fluctuations. LES (with dynamicKEqn turbulence model)
is able to resolve the turbulent fluctuations more accurately and realistically than RANS.
LES simulation with static turbulence model (kEqn) performs very poorly with provided
LES mesh resolution (y+ ~ 3). The static LES turbulence model is not only unable to predict
flow separation but also overpredicts the mean flow and turbulent fluctuating velocities
by a huge margin. In LES, the calculated mean and fluctuating velocities represent the
statistical fields, which is a function of time over which the averaging is performed (in
our case: 3 s). A more accurate and realistic approximation of flow statistics would be
obtained if performed over a longer duration. In Figure 5, the simulated velocity fields are
shown under (a) RANS and (b) LES (using dynamicKEqn turbulence model) frameworks,
representing the level of resolution obtained under these approaches. RANS flow fields
are very smooth as they represent mean values, while eddy generation and decay can be
seen in LES flow fields. It must be emphasized that with the mesh resolution used in our
LES simulations (y+ ~ 3), static LES turbulence models were unable to predict correct fluid
Fluids 2022, 7, x FOR PEER REVIEW velocity profiles, whereas the dynamic LES turbulence model resulted in a relatively real
15 of 26
estimation of mean and fluctuating fluid velocity with the used relatively coarser mesh, as
shown in our fluid velocity plots.

(a)

(b)

Figure 4. Comparisons of experimental and OpenFOAM simulated streamwise air (a) mean and (b)
Figure 4. Comparisons of experimental and OpenFOAM simulated streamwise air (a) mean and
fluctuating velocity profiles at the measurement locations under RANS and LES frameworks.
(b) fluctuating velocity profiles at the measurement locations under RANS and LES frameworks.

(a)
Fluids 2022, 7, 337 15 of 26
Figure 4. Comparisons of experimental and OpenFOAM simulated streamwise air (a) mean and (b)
fluctuating velocity profiles at the measurement locations under RANS and LES frameworks.

(a)

(b)

Figure 5. OpenFOAM simulated velocity fields (magnitude) under (a) RANS (kOmegaSST) and (b)
Figure 5. OpenFOAM simulated velocity fields (magnitude) under (a) RANS (kOmegaSST) and
LES (dynamicKEqn) frameworks.
(b) LES (dynamicKEqn) frameworks.

In our next
In our nextRANS-DEM
RANS-DEMand andLES-DEM
LES-DEM sections,
sections, wewe mainly
mainly focused
focused on theon simulation’s
the simula-
tion’s
accuracy in predicting particle dispersion on the influence of the carrier fluid fluid
accuracy in predicting particle dispersion on the influence of the carrier (air) (air)
flow
flow turbulence. However, even the presence of the particles can modify
turbulence. However, even the presence of the particles can modify the turbulence, as the turbulence,
as discussed
discussed in in
thethe introduction.The
introduction. Thepotential
potentialmodification
modificationin in air
air turbulence
turbulence duedue to
to the
the
presence
presence of
of particles
particles isis not
not explicitly
explicitly discussed
discussed in
in this
this study.
study. Additionally,
Additionally, thethe turbulence
turbulence
models
models employed
employeddo donotnotconsider
considerthe presence
the presenceof particles as they
of particles are developed
as they are developedfor sin-
for
gle-phase fluid flow and some works have noticed the potential failure
single-phase fluid flow and some works have noticed the potential failure of employedof employed tur-
bulence models
turbulence models in in
specific flow
specific scenarios
flow scenarios[55].
[55].However,
However,new newmathematical
mathematical turbulence
turbulence
models considering the particle’s presence are yet to be developed and do not come under
the scope of this study.

4.2. RANS-DEM
We have investigated the effect of one-way and two-way coupling on the continuum
and discrete phase results and found almost no difference between the results of either
coupling regime. The mass flow rate of 10% and the corresponding volumetric fraction
of particles is in order of ~10−5 , which is indeed in the range of the two-way coupling
threshold suggested by Elghobashi [4]. Interestingly our one-way and two-way coupling
results for fluid phase (Figure 6) and particle phase (Figures 7 and 8; red circles) almost
overlap each other, suggesting that one-way coupling might also be adopted when particle
volumetric fraction is slightly greater (O ~ 10−5 ) than the threshold for one-way coupling.
As demonstrated by our fluid and particle phase results, one-way coupling seems to be
sufficient even for particle concentration in O ~ 10−5 , which is slightly greater than the
standard threshold for one-way coupling.
Figure 9 also shows that when using RANS-DEM, the fluid mean velocity profiles
agree very well with the experimental data, as in the single-phase results. For brevity, we
show here streamwise air mean velocity profiles (Figure 9) only for two-way coupling
corresponding to different particle boundary conditions (injection velocity), as they are
very similar to RANS-DEM results for one-way coupling and single-phase RANS results.
Figure 9 shows that the air-phase results are independent of particle injection velocities due
to small particle concentration. Under the RANS-DEM framework, turbulent fluctuations
are significantly underpredicted, which can be seen in single-phase plots, and this under-
prediction of turbulent fluctuation also reflects in particle velocity plots (Figures 7 and 8;
red circles), where particles move roughly like a patch and do not disperse below the
step (y/H < 1) even after flow re-attachment (x/H ~ 7). This underprediction of particle
dispersion behind the step can also be seen in Figure 10a, which shows the particle spread
behind the step at t = 1 s.
ations are significantly underpredicted, which can be seen in single-phase plots, and this
underprediction of turbulent fluctuation also reflects in particle velocity plots (Figures 7
and 8; red circles), where particles move roughly like a patch and do not disperse below
the step (y/H < 1) even after flow re-attachment (x/H ~ 7). This underprediction of particle
dispersion behind the step can also be seen in Figure 10a, which shows the particle spread
Fluids 2022, 7, 337 16 of 26
behind the step at t = 1 s.

Fluids 2022, 7, x FOR PEER REVIEW Figure


Figure6.6.Comparisons
Comparisonsof
of experimental and OpenFOAM
experimental and OpenFOAMsimulated
simulated(RANS-DEM)
(RANS-DEM) streamwise
ofair
17air
streamwise 26
mean
meanvelocity
velocityprofiles
profilesat
atthe
themeasurement locationsfor
measurement locations forone-way
one-wayand
andtwo-way
two-way coupling
coupling forfor
thethe case
case
corresponding to a particle injection velocity of 10.5 m/s.
corresponding to a particle injection velocity of 10.5 m/s.

One-way coupling

(a)

(b)

Figure 7. Cont.

(c)
Fluids 2022, 7, 337 17 of 26

(c)

Figure 7. Comparisons of experimental and OpenFOAM simulated average streamwise particle ve-
Figure 7. Comparisons of experimental and OpenFOAM simulated average streamwise particle
locity profiles at the measurement locations for one-way coupling and 10% mass loading of copper
velocity profiles at the measurement locations for one-way coupling and 10% mass loading of copper
particles under RANS-DEM framework. (a) Particle injection velocity as (10.5, 0, 0) m/s, (b) particle
Fluids 2022, 7, x FOR PEER REVIEW particles under RANS-DEM framework. (a) Particle injection velocity as (10.5, 0, 0) m/s, (b) 18 particle
of 26
injection velocity same as that of the fluid velocity profile at inlet, (c) particle injection velocity as (0,
injection velocity same as that of the fluid velocity profile at inlet, (c) particle injection velocity as (0,
0, 0) m/s.
0, 0) m/s.

Two-way coupling

(a)

(b)

Figure 8. Cont.

(c)
Fluids 2022, 7, 337 18 of 26

(c)

Figure 8. Comparisons of experimental and OpenFOAM simulated average streamwise particle ve-
Figure 8. Comparisons of experimental and OpenFOAM simulated average streamwise particle
locity profiles at the measurement locations for two-way coupling and 10% mass loading of copper
velocity
particlesprofiles
under at the measurement
RANS-DEM locationsframeworks.
and LES-DEM for two-way(a)
coupling
Particleand 10% mass
injection loading
velocity of copper
as (10.5, 0, 0)
particles under RANS-DEM and LES-DEM frameworks. (a) Particle injection velocity
Fluids 2022, 7, x FOR PEER REVIEW m/s, (b) a particle injection velocity same as that of the fluid velocity profile at the inlet, (c)19as (10.5, 0,
of 26
particle
0) m/s, (b)velocity
injection a particle injection
as (0, velocity same as that of the fluid velocity profile at the inlet, (c) particle
0, 0) m/s.
injection velocity as (0, 0, 0) m/s.

Two-way coupling

(a)

(b)

Figure 9. Cont.

(c)
Fluids 2022, 7, 337 19 of 26

(c)

Figure 9. Comparisons of experimental and OpenFOAM simulated streamwise air mean velocity
Figure 9. Comparisons of experimental and OpenFOAM simulated streamwise air mean velocity
profiles at the measurement locations for two-way coupling and 10% mass loading of copper parti-
profiles at the measurement locations for two-way coupling and 10% mass loading of copper particles
cles under RANS-DEM framework. (a) Particle injection velocity as (10.5, 0, 0) m/s, (b) particle in-
under RANS-DEM framework. (a) Particle injection velocity as (10.5, 0, 0) m/s, (b) particle injection
Fluids 2022, 7, x FOR PEER REVIEW jection velocity same as that of the fluid velocity profile at the inlet, (c) particle injection velocity
20 of as
26
velocity same as that of the fluid velocity profile at the inlet, (c) particle injection velocity as (0, 0,
(0, 0, 0) m/s.
0) m/s.

(a)
a

(b)
a

Figure 10. Screenshot showing particle dispersion behind the step at 1 s under (a) RANS-DEM and
Figure 10. Screenshot showing particle dispersion behind the step at 1 s under (a) RANS-DEM and
(b) LES-DEM frameworks.
(b) LES-DEM frameworks.
One
One ofof the main questions
the main questions that
that we
we tried
tried to answer is:
to answer is: how
how effectively
effectively can can RANS-
RANS-
DEM, with simple dispersion models, predict particle dispersion
DEM, with simple dispersion models, predict particle dispersion in turbulent flows? in turbulent flows? In In
other
other words,
words, itsits effectiveness
effectiveness in
in obtaining
obtaining instantaneous
instantaneous flow flow fields
fields so
so that
that the
the turbulence
turbulence
effect
effect on
on the
the particle’s
particle’smotion
motioncan canbe bemodelled
modelledaccurately.
accurately.For Forthis
thispurpose,
purpose, two twosimple
sim-
dispersion
ple models
dispersion are available
models in OpenFOAM,
are available in OpenFOAM, namely,
namely,GradientDispersionRAS
GradientDispersionRAS and and
Sto-
chasticDispersionRAS, based
StochasticDispersionRAS, on Equation
based on Equation(18).(18).
An initial investigation
An initial investigationreveals that that
reveals the Sto-
the
chasticDispersionRAS dispersion model gives slightly better results in relation
StochasticDispersionRAS dispersion model gives slightly better results in relation to particle to particle
dispersion
dispersion andand was
was also
also adopted
adopted by by Greifzu
Greifzu etet al.
al. [15]
[15] in
in their
their RANS-DEM
RANS-DEMnumerical
numericalsimu- sim-
ulations. Therefore, we decided to use the StochasticDispersionRAS
lations. Therefore, we decided to use the StochasticDispersionRAS model in our RANS-DEM model in our RANS-
DEM simulations.
simulations. FigureFigure
7 shows 7 shows the normalized
the normalized streamwise streamwise
averageaverage
particle particle
velocity velocity
profiles
profiles corresponding
corresponding to one-way to coupling
one-way for coupling
different forparticle
different particleconditions
boundary boundary(injection
conditions ve-
(injection
locity). Thevelocity). The results demonstrate
results demonstrate that particle-phase
that particle-phase results are alsoresults are alsofor
very similar very simi-
one-way
lar for one-way
(Figure (Figure 7)
7) and two-way and two-way
coupling (Figurecoupling (Figure particle
8), but different 8), but different
injectionparticle
velocities injection
have a
velocities have a notable effect on particle-phase results, unlike air-phase
notable effect on particle-phase results, unlike air-phase results. When particles are injected results. When
particles
with are0,injected
(10.5, 0) m/swith (10.5, 0, velocity
of injection 0) m/s offrom
injection
the velocity
inlet of BFSfrom(Figure
the inlet of BFS
7a), (Figure
the particle
7a), the particle velocities at all the measurement locations are overpredicted, and their
dispersion (spread in the y-direction) is underpredicted. When particles are injected with
the velocity that follows the fluid velocity profile at the inlet (Figure 7b), the results im-
proved slightly but were still far from the experimental data, in which the particles moved
Fluids 2022, 7, 337 20 of 26

velocities at all the measurement locations are overpredicted, and their dispersion (spread
in the y-direction) is underpredicted. When particles are injected with the velocity that
follows the fluid velocity profile at the inlet (Figure 7b), the results improved slightly but
were still far from the experimental data, in which the particles moved roughly like a patch,
and a minimal number of particles were dispersed below the step (y/H < 1). When particles
are injected with (0, 0, 0) m/s of injection velocity (Figure 7c), the particle velocity profiles
provide a better match with the experimental data, but the dispersion of the particles is still
considerably underpredicted, similar to other RANS-DEM cases. It can also be seen from
Figure 8 that a very similar particle dispersion behavior is observed, even for two-way
coupling under the RANS-DEM framework. In the RANS-DEM framework, all the cases
with different boundary conditions for the particle-phase and under the considered cou-
pling regime underpredict the particle dispersion by a considerable margin and a minimal
number of particles are found below the step (y/H < 1) even after the flow re-attachment
point (x/H ~ 7). On the other hand, experimental data shows that particles disperse across
the extended channel section until they reach the re-attachment point (x/H ~ 7), and, after
this, the particle concentration below and above the step becomes almost uniform. Our
analysis shows that the simple dispersion models (DRW) are ineffective in incorporating
turbulence effects on the trajectory of particles.
In comparison to the RANS-DEM results of Greifzu et al. [15], our fluid flow results
are in good agreement with them and with the experimental data as well, but the particle-
phase results are entirely different. Their results showed that the RANS-DEM (and simple
dispersion model therein) can predict the correct particle dispersion and their velocity
profiles. Interestingly, we found that they were using a fluid density of 1000 kg/m3 (water)
instead of 1 kg/m3 (air) in their OpenFOAM numerical setup, even though the fluid used
in the experiment was air, not water. Using a density of 1000 kg/m3 results in a higher
body (buoyant force) and coupled forces (drag and pressure gradient force) acting on the
particles, which would disperse the particles in the domain even before the re-attachment
point, as observed in the particle velocity plots of Greifzu et al. [15]. It is demonstrated
that particles will be dispersed into the recirculation region only when their large-eddy
Stokes numbers are less than one [56]. The large-eddy Stokes numbers for fluid as air
(density = 1 kg/m3 ) and as water (density = 1000 kg/m3 ) are found to be 6.9 and 0.053,
respectively. Obviously, when the fluid density is that of water, the Stokes number is
significantly smaller than one, so the particles will also be dispersed into the recirculation
region (as they behave similar to tracers). Fluid flow results remain almost the same even
if the density of water is used instead of air because the momentum transferred from the
particle phase to the fluid phase remains of the same order of magnitude. This can be
explained as small particle concentrations result only in a few numbers of particles in each
computational cell and are simply not numerous enough to modify the fluid flow fields.
So, even by considering two-way coupling and a density of fluid as that of water does not
modify the fluid flow fields. This explanation also indicates that the previous observation
regarding the ability of the simple dispersion model (DRW) in accurately incorporating
turbulence effects on particle’s trajectory is not true and supports our observation about
simple dispersion models.
The dispersion models available in OpenFOAM are essentially discrete random walk
(DRW) type models and are calculated using the modelled turbulent kinetic energy (TKE).
These models are simple models based on rough assumptions, e.g., that the turbulence is
isotropic in the whole domain, which leads to the inappropriate modelling of turbulence
seen by particles. However, the turbulence is very anisotropic in the boundary layers, and
this anisotropic behavior is even more significant for wall-bounded flows with complex
geometries such as BFS flow. The shortcomings of the discrete random walk (DRW) type
of dispersion models can be avoided by better treatment of boundary layer effects. For
this purpose, an option could be the Continuous Random Walk (CRW) method to be
included in OpenFOAM, which offers a more physically sound way of modelling particle
dispersion [57]. The anisotropic behavior of turbulent flow is better modelled using the
Fluids 2022, 7, 337 21 of 26

CRW method, recently presented by Mofakham [13]. An alternative stochastic approach to


describe the particle dispersion due to turbulence could be a straightforward generalization
of the stochastic approach introduced by Pope [58], which was originally developed to
describe single-phase flow. This approach is extended to describe the two-phase system by
Peirano et al. [59]. Xiang [14] used this stochastic model in their numerical simulation in
OpenFOAM and reported that it performs better than the already implemented dispersion
models (DRW models) in OpenFOAM. Although even with their implemented stochastic
model, the simulated particle dispersion was far from the reference data. There are also
the Reynolds stress transport models (RSTM) that directly evaluate the components of
Reynolds stresses and account for the anisotropy of turbulent fluctuations [60–62]. However,
knowledge of instantaneous fluctuations is required for specific problems, such as the one
involving particle dispersion and deposition. It was recently shown that a RANS approach
in conjunction with RSTM and DRW does not improve the results in terms of particle
dispersion, and that more sophisticated dispersion models such as CRW must be used [13].
In the original experiment, the particles traversed a sufficient distance (a development
channel length of 5.2 m) before they reached the inlet of BFS. This assured that the particles
had attained equilibrium with the fluid phase before the inlet of BFS, as they had enough
time (at least three particle response time, in the worst case) to come to equilibrium with the
fluid phase [53]. As at the inlet of BFS, the particle velocity profile was not measured and/or
available in the literature, and we do not know the exact particle velocity when they reach
the inlet of BFS. It is quite difficult to approximate the real particle velocities at the inlet of
BFS without this development channel. Our results demonstrate that the straightforward
assumption that all the particles have attained mean flow velocity seems unreasonable. The
best approximation of particle injection velocity was found to correspond to an injection
velocity of (0, 0, 0) m/s. This might be due to the fact that the particles may obtain real
physical velocity depending upon the particle reaction time and fluid flow around it. The
additional injection velocity, which one needs to provide along with the mass flux of the
particles, does not seem to be necessary as we specify mass flux of particles (e.g., kg/s)
that is injected from the inlet of the BFS. Once the particles are in the domain, they attain
velocities depending upon the flow around them and the particle response time (Stokes
number). However, this approach may vary in individual cases, and the results might look
different when the inlet channel (before BFS) length is increased. It also depends on the
different algorithms that different software use for particle generation and insertion. The
best practice guidelines for CFD should still be the extension of the inlet channel length and
allowing particles to develop real physical velocity, but this might be extra computational
overhead for CFD-DEM simulations.

4.3. LES-DEM
We decided to perform LES-DEM simulations to investigate the case in more detail;
fluid flow fields are calculated using the LES approach, then particle trajectories are cal-
culated based on resolved LES fluid flow fields without considering additional models to
include the effects of SGS on the motion of the particle. LES-DEM simulated fluid-flow
fields agree well with the experimental data, such as single-phase LES simulations. LES-
DEM simulated fluid velocity profiles are not shown here for the sake of brevity. The
LES-predicted turbulent fluctuation was a significant improvement over the RANS ap-
proach (see Figure 4b), and this is also reflected in particle dispersion, as seen in Figure 8
(green circles) and Figure 10b.
Figure 8 shows normalized streamwise average particle velocity profiles for 10% mass
loading of ~70 µm copper particles for two-way coupling corresponding to the different
options concerning the particle injection velocities under RANS-DEM (red circles) and
LES-DEM (green circles) frameworks. Compared to the RANS-DEM results, particle dis-
persion and velocity profiles have been improved considerably due to the ability of LES
to resolve flow fields in greater detail and improved predictions in view of the improved
representation of the flow field seen by the particles. Moreover, when the particles are
Fluids 2022, 7, 337 22 of 26

injected with an injection velocity of (10.5, 0, 0) m/s from the inlet (Figure 8a), the par-
ticle velocities at all measurement locations are slightly overpredicted. When particles
are injected with an injection velocity that follows the fluid velocity profile at the inlet
(Figure 8b), the particle velocity profiles seem to be slightly overpredicted compared to
those observed in the experiment. When particles are injected with an injection velocity of
(0, 0, 0) m/s from the inlet of BFS (Figure 8c), the particle velocity profiles give a reasonably
good match with the experimental data at all the measurement locations. In all the three
LES-DEM cases, a very small number of particles is found in the recirculation region,
but after the re-attachment point (x/H ~ 7), enough particle dispersion is predicted as
observed in the original experiment. Compared to the original experiment, our LES-DEM
still underpredicts the particles’ dispersion, especially in the recirculation region. Taking
both the correctness of predicting particle velocity and their dispersion into account, we
found that the best results were obtained when particles are injected with (0, 0, 0) m/s of
injection velocity, and the probable reason of it being the best option to approximate the
particle boundary condition without extending the inlet channel is already explained in
our RANS-DEM section.
It can be clearly seen that the overall particle-phase results have improved considerably
with the LES-DEM approach. The particle-phase results, especially its dispersion, under
the LES-DEM framework can be further improved with increased mesh resolution (y+ ~1)
and/or with the inclusion of missing SGS in the equation of particle motion. However,
one must always bear in mind that extreme fine meshing can cause stability problems,
as CFD cells cannot be smaller than particle diameter under the unresolved CFD-DEM
framework, and can thus only be applied under consideration of point-particles (PP).
Additionally, further mesh refinement would require much more computational resources.
A recent study indicates that the use of stochastic dispersion models is necessary even in
the LES-DEM framework, especially for the fine particles, where the corresponding particle
relaxation time is of the same order as the smallest fluid flow time scale [63]. Our LES-DEM
results also indicate that the effect of SGS cannot be neglected and has a significant effect on
particle trajectories, especially for the particles with a small Stokes number, which was also
suggested by Ref. [10]. The effect of SGS on particle’s trajectory seems to play an even more
important role in LES-DEM simulations, where very fine meshes might not be possible
due to the prerequisite of unresolved CFD-DEM, where a particle must be significantly
smaller than the CFD cell size and must be recovered with efficient modelling approaches,
as shown in our results.

5. Summary and Conclusions


In this work, we have assessed the capability of OpenFOAM to solve the particle-laden
BFS flow in the different frameworks of RANS-DEM and LES-DEM. The RANS method
only provides the mean flow fields. Therefore, an additional dispersion model is used to
include the effect of turbulent fluctuation on the trajectory of particles. In contrast, LES
can resolve up to energy-containing eddies, but several constraints such as particles being
smaller than the CFD mesh in the unresolved CFD-DEM and computational limitations
restrict the extreme refinement of CFD mesh. Thus, it is necessary to make a compromise
between the accuracy obtained and the computational resources required, which is quite
challenging. Collectively, the following conclusions can be drawn from our study:
• We found that the threshold of coupling regime suggested by Elghobashi [4] is rigidly
formulated and it might be sufficient to include one-way coupling even when the
particle concentration is in O ~ 10−5 , since we found almost no difference between the
fluid and particle phase results for one-way and two-way coupling;
• Under the RANS-DEM framework, simple dispersion models based on DRW sig-
nificantly underpredicted the particle dispersion. Consequently, more sophisticated
dispersion models such as CRW must be used in conjunction with RANS-DEM. Previ-
ously claimed results about the ability of the simple dispersion models in accurately
incorporating turbulence effects on particles were due to error in the numerical setup;
Fluids 2022, 7, 337 23 of 26

• When using relatively coarse mesh resolution (y+ > 1), Dynamic LES turbulence
models seem to overcome the poor performance of static LES turbulence models in
predicting the mean and fluctuating components of turbulent flow. We recommend
using dynamic LES models when extreme mesh refinement is not possible due to the
limitation of particles being smaller than CFD cell size in unresolved CFD-DEM;
• Resolved CFD-DEM (particle resolved DNS) requires huge computational resources
and is restricted to a small number of particles. Point particle DNS-DEM is also com-
putationally expensive and should not be applied for larger particles than Kolmogorov
length scale. The LES-DEM seems to be a good compromise between accuracy and
computational feasibility. However, its application is mostly restricted to simple cases
(point-particles or small particles) due to the constraint of the particles being smaller
than the CFD cells in unresolved CFD-DEM. In addition, the unresolved component
of the turbulent velocity (SGS) seems to have a significant effect on particle dispersion
and cannot be neglected, especially when using larger y+ in LES-DEM;
• Our analysis of different options for approximating the initial particle velocity (particle
injection velocity) indicates that a suitable numerical approach might be to inject
particles with (0, 0, 0) m/s of particle injection velocity. The difference between the
results is small, but still might be appropriate so as to let the particles attain the real
physical velocity according to physics, for the cases where the initial particle velocity
is unavailable. However, this approach is case-dependent and software-specific. The
best practice guidelines for CFD should still be the extension of the inlet channel length
and allowing particles to develop real physical velocity, but this might be an extra
computational overhead for CFD-DEM simulations.
From our point of view, one of the best options for gaining success in predicting
dilute particle dispersion in turbulence flow can be an accurate calculation of the mean
flow statistics and a good stochastic model, although here, further benchmarking is still
necessary. More fundamental research and validations are required in both RANS-DEM
and LES-DEM before the complex physics related to fluid-particle systems can be studied
in detail, considering all factors such as surface, inertial, response, loading, and interaction
effects into account. Application of RANS-DEM and LES-DEM for real problems would
require larger CFD meshes, resulting in loss of information about turbulent fluctuations.
If we could recover this lost information with simple yet efficient methods, then it would
be of great engineering application. With the currently available computational resources,
both resolved CFD-DEM (particle-resolved DNS-DEM) and point-particle DNS-DEM are
still limited to simple cases with a small number of particles. More efficient algorithms and
computer architecture are required to achieve this, and more research should be encouraged
in this field.

Author Contributions: A.J. (conceptualization, investigation, data curation, writing—original draft),


M.D.B. (supervision, writing—review and editing), P.R. (supervision). All authors have read and
agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Acknowledgments: We would like to acknowledge the CFD-online community (https://www.
cfd-online.com/ (accessed on 4 September 2022)), which provided us with a platform to discuss
the problem and solutions with experts from all over the world. We would like to thank Rüdiger
Schwarze and his team (TU Freiberg) for confirming the error in their numerical setup, when reported.
We would also like to thank Silvio Schmalfuß (Leibniz Institute for Tropospheric Research, Leipzig)
and Robert Kasper (University of Rostock) for their valuable suggestions. We would like to thank the
Leibniz Supercomputing Centre (LRZ) for providing computational resources. We appreciate the
Fluids 2022, 7, 337 24 of 26

help and continuous encouragement from Nils Rüther (TU Munich). A.J. gratefully acknowledges
the Konrad-Adenauer-Stiftung (KAS) for a PhD scholarship. We thank the editor and anonymous
reviewers for their constructive comments, which helped us to improve the manuscript.
Conflicts of Interest: On behalf of all authors, the corresponding author states that there is no conflict
of interest.

References
1. Crowe, C.T.; Schwarzkopf, J.D.; Sommerfeld, M.; Tsuji, Y. Multiphase Flows with Droplets and Particles, 2nd ed.; CRC Press: Boca
Raton, FL, USA; London, UK; New York, NY, USA, 2011.
2. Brennen, C.E. Fundamentals of Multiphase Flows; Cambridge University Press: Pasadena, CA, USA, 2005.
3. Zhu, H.P.; Zhou, Z.Y.; Yang, R.Y.; Yu, A.B. Discrete Particle Simulation of Particulate Systems: Theoretical Developments. Chem.
Eng. Sci. 2007, 62, 3378–3396. [CrossRef]
4. Elghobashi, S. On predicting particle-laden turbulent flows. Appl. Sci. Res. 1994, 52, 309–329. [CrossRef]
5. Salih, S.Q.; Aldlemy, M.S.; Rasani, M.R.; Ariffin, A.K.; Ya, T.M.Y.S.T.; Al-Ansari, N.; Yaseen, Z.M.; Chau, K.W. Thin and sharp
edges bodies-fluid interaction simulation using cut-cell immersed boundary method. Eng. Appl. Comput. Fluid Mech. 2019, 13,
860–877. [CrossRef]
6. Squires, K.D.; Eaton, J.K. Particle response and turbulence modification in isotropic turbulence. Phys. Fluids A Fluid Dyn. 1990, 2,
1191–1203. [CrossRef]
7. Elghobashi, S.; Truesdell, G.C. On the two-way interaction between homogeneous turbulence and dispersed solid particles. I:
Turbulence modification. Phys. Fluids A Fluid Dyn. 1993, 5, 1790–1801. [CrossRef]
8. Chan, L.; Zahtila, T.; Ooi, A.; Philip, J. Turbulence Modification Due to Inertial Particles in a Rough-Wall Pipe. In Proceedings
of the 22nd Australasian Fluid Mechanics Conference AFMC2020, Brisbane, Australia, 7–10 December 2020; The University of
Queensland: Brisbane, Australia, 2020.
9. Lee, J.; Lee, C. The Effect of Wall-Normal Gravity on Particle-Laden Near-Wall Turbulence. J. Fluid Mech. 2019, 873, 475–507.
[CrossRef]
10. Kuerten, J.G.M. Point-Particle DNS and LES of Particle-Laden Turbulent flow—A state-of-the-art review. Flow Turbul. Combust.
2016, 97, 689–713. [CrossRef]
11. Bocksell, T.L.; Loth, E. Stochastic modeling of particle diffusion in a turbulent boundary layer. Int. J. Multiph. Flow 2006, 32,
1234–1253. [CrossRef]
12. Loth, E. Numerical approaches for motion of dispersed particles, droplets and bubbles. Prog. Energy Combust. Sci. 2000, 26,
161–223. [CrossRef]
13. Mofakham, A.A.; Ahmadi, G. On random walk models for simulation of particle-laden turbulent flows. Int. J. Multiph. Flow 2020,
122, 103157. [CrossRef]
14. Xiang, L. RANS Simulations and Dispersion Models for Particles in Turbulent Flows. Master’s Thesis, KTH, School of Engineering
Sciences (SCI), Mechanics, Stockholm, Sweden, 2019.
15. Greifzu, F.; Kratzsch, C.; Forgber, T.; Lindner, F.; Schwarze, R. Assessment of particle-tracking models for dispersed particle-laden
flows implemented in OpenFOAM and ANSYS FLUENT. Eng. Appl. Comput. Fluid Mech. 2016, 10, 30–43. [CrossRef]
16. Innocenti, A.; Marchioli, C.; Chibbaro, S. Lagrangian filtered density function for LES-based stochastic modelling of turbulent
particle-laden flows. Phys. Fluids 2016, 28, 115106. [CrossRef]
17. Kuerten, J.G.M.; Vreman, A.W. Can turbophoresis be predicted by large-eddy simulation? Phys. Fluids 2005, 17, 011701. [CrossRef]
18. Marchioli, C.; Salvetti, M.V.; Soldati, A. Some issues concerning large-eddy simulation of inertial particle dispersion in turbulent
bounded flows. Phys. Fluids 2008, 20, 40603. [CrossRef]
19. Salmanzadeh, M.; Rahnama, M.; Ahmadi, G. Effect of Sub-Grid Scales on Large Eddy Simulation of Particle Deposition in a
Turbulent Channel Flow. Aerosol Sci. Technol. 2010, 44, 796–806. [CrossRef]
20. Bianco, F.; Chibbaro, S.; Marchioli, C.; Salvetti, M.V.; Soldati, A. Intrinsic filtering errors of Lagrangian particle tracking in LES
flow fields. Phys. Fluids 2012, 24, 45103. [CrossRef]
21. Geurts, B.J.; Kuerten, J.G.M. Ideal stochastic forcing for the motion of particles in large-eddy simulation extracted from direct
numerical simulation of turbulent channel flow. Phys. Fluids 2012, 24, 81702. [CrossRef]
22. Eppinger, T.; Seidler, K.; Kraume, M. DEM-CFD Simulations of Fixed Bed Reactors with Small Tube to Particle Diameter Ratios.
Chem. Eng. J. 2011, 166, 324–331. [CrossRef]
23. Fries, L.; Antonyuk, S.; Heinrich, S.; Palzer, S. DEM–CFD modeling of a fluidized bed spray granulator. Chem. Eng. Sci. 2011, 66,
2340–2355. [CrossRef]
24. Jajcevic, D.; Siegmann, E.; Radeke, C.; Khinast, J.G. Large-Scale CFD–DEM Simulations of Fluidized Granular Systems. Chem.
Eng. Sci. 2013, 98, 298–310. [CrossRef]
25. Mohammadreza, E. Experimental and Numerical Investigations of Horizontal Pneumatic Conveying. Ph.D. Thesis, The University
of Edinburgh, Edinburgh, UK, 2014.
26. Spogis, N. Multiphase Modeling Using EDEM-CFD Coupling for FLUENT. In Proceedings of the 3rd Latin Americal CFD
Workshop Applied to the Oil Industry, Rio de Janeiro, Brazil, 18–19 August 2008.
Fluids 2022, 7, 337 25 of 26

27. Breuer, M.; Alletto, M. Efficient simulation of particle-laden turbulent flows with high mass loadings using LES. Int. J. Heat Fluid
Flow 2012, 35, 2–12. [CrossRef]
28. Capecelatro, J.; Desjardins, O. An Euler–Lagrange strategy for simulating particle-laden flows. J. Comput. Phys. 2013, 238, 1–31.
[CrossRef]
29. Deb, S.; Tafti, D.K. A Novel Two-Grid Formulation for Fluid–Particle Systems Using the Discrete Element Method. Powder Technol.
2013, 246, 601–616. [CrossRef]
30. Drake, T.G.; Calantoni, J. Discrete particle model for sheet flow sediment transport in the nearshore. J. Geophys. Res. 2001, 106,
19859–19868. [CrossRef]
31. Wu, C.L.; Ayeni, O.; Berrouk, A.S.; Nandakumar, K. Parallel Algorithms for CFD-DEM Modeling of Dense Particulate Flows.
Chem. Eng. Sci. 2014, 118, 221–244. [CrossRef]
32. OpenCFD Limited. OpenFOAM, v1912; OpenCFD Limited: Bracknell, UK, 2019. Available online: https://www.openfoam.com/
news/main-news/openfoam-v1906 (accessed on 4 September 2022).
33. Kloss, C.; Goniva, C.; Hager, A.; Amberger, S.; Pirker, S. Models, algorithms and validation for opensource DEM and CFD-DEM.
Prog. Comput. Fluid Dyn. 2012, 12, 140. [CrossRef]
34. Sun, R.; Xiao, H. SediFoam: A General-Purpose, Open-Source CFD-DEM Solver for Particle-Laden Flow with Emphasis on
Sediment Transport. Comput. Geosci. 2016, 89, 207–219. [CrossRef]
35. Bing, W.; Hui Qiang, Z.; Xi Lin, W. Large-Eddy Simulation of Particle-Laden Turbulent Flows over a Backward-Facing Step
Considering Two-Phase Two-Way Coupling. Adv. Mech. Eng. 2013, 5, 325101. [CrossRef]
36. Kasper, R.; Turnow, J.; Kornev, N. Numerical modeling and simulation of particulate fouling of structured heat transfer surfaces
using a multiphase Euler-Lagrange approach. Int. J. Heat Mass Transf. 2017, 115, 932–945. [CrossRef]
37. Wang, B.; Wei, W.E.; Zhang, H. A study on correlation moments of two-phase fluctuating velocity using direct numerical
simulation. Int. J. Mod. Phys. C 2013, 24, 1350068. [CrossRef]
38. Yu, K.F.; Lau, K.S.; Chan, C.K. Large eddy simulation of particle-laden turbulent flow over a backward-facing step. Commun.
Nonlinear Sci. Numer. Simul. 2004, 9, 251–262. [CrossRef]
39. Yu, K.F.; Lau, K.S.; Chan, C.K. Numerical simulation of gas-particle flow in a single-side backward-facing step flow. J. Comput.
Appl. Math. 2004, 163, 319–331. [CrossRef]
40. Yu, K.F.; Lee, E.W. Evaluation and modification of gas-particle covariance models by Large Eddy Simulation of a particle-laden
turbulent flows over a backward-facing step. Int. J. Heat Mass Transf. 2009, 52, 5652–5656. [CrossRef]
41. Armenio, V.; Fiorotto, V. The importance of the forces acting on particles in turbulent flows. Phys. Fluids 2001, 13, 2437–2440.
[CrossRef]
42. Wilcox, D.C. Turbulence Modeling for CFD, 2nd ed.; Printing (with Corrections); DCW Industries: La Cãnada, CA, USA, 1998.
43. Uijttewaal, W.S.J.; Oliemans, R.V.A. Particle dispersion and deposition in direct numerical and large eddy simulations of vertical
pipe flows. Phys. Fluids 1996, 8, 2590–2604. [CrossRef]
44. Yeh, F.; Lei, U. On the motion of small particles in a homogeneous isotropic turbulent flow. Phys. Fluids A Fluid Dyn. 1991, 3,
2571–2586. [CrossRef]
45. Minier, J.P.; Chibbaro, S.; Pope, S.B. Guidelines for the Formulation of Lagrangian Stochastic Models for Particle Simulations of
Single-Phase and Dispersed Two-Phase Turbulent Flows. Phys. Fluids 2014, 26, 113303. [CrossRef]
46. Sodja, J. Turbulence models in CFD. Master’s Thesis, University of Ljubljana, Ljubljana, Slovenia, 2007.
47. Kuczaj, A.K.; Geurts, B.J. Mixing in manipulated turbulence. J. Turbul. 2006, 7, N67. [CrossRef]
48. Simonin, O.; Deutsch, E.; Minier, J.P. Eulerian prediction of the fluid/particle correlated motion in turbulent two-phase flows.
Appl. Sci. Res. 1993, 51, 275–283. [CrossRef]
49. Benra, F.K.; Dohmen, H.J.; Pei, J.; Schuster, S.; Wan, B. A Comparison of One-Way and Two-Way Coupling Methods for Numerical
Analysis of Fluid-Structure Interactions. J. Appl. Math. 2011, 2011, 853560. [CrossRef]
50. Chen, S.F.; Lei, H.; Wang, M.; Yang, B.; Dai, L.J.; Zhao, Y. Two-way coupling calculation for multiphase flow and decarburization
during RH refining. Vacuum 2019, 167, 255–262. [CrossRef]
51. Kitagawa, A.; Murai, Y.; Yamamoto, F. Two-way coupling of Eulerian–Lagrangian model for dispersed multiphase flows using
filtering functions. Int. J. Multiph. Flow 2001, 27, 2129–2153. [CrossRef]
52. Ruetsch, G.R.; Meiburg, E. Two-way coupling in shear layers with dilute bubble concentrations. Phys. Fluids 1994, 6, 2656–2670.
[CrossRef]
53. Fessler, J.R.; Eaton, J.K. Turbulence modification by particles in a backward-facing step flow. J. Fluid Mech. 1999, 394, 97–117.
[CrossRef]
54. Pope, S.B. Ten questions concerning the large-eddy simulation of turbulent flows. New J. Phys. 2004, 6, 35. [CrossRef]
55. Gimenez, J.M.; Idelsohn, S.R.; Oñate, E.; Löhner, R. A Multiscale Approach for the Numerical Simulation of Turbulent Flows with
Droplets. Arch. Comput. Methods Eng. State Art Rev. 2021, 28, 4185–4204. [CrossRef]
56. Hardalupas, Y.; Taylor, A.M.K.P.; Whitelaw, J.H. Particle Dispersion in a Vertical Round Sudden-Expansion Flow. Philos. Trans.
Phys. Sci. Eng. 1992, 341, 411–442.
57. Dehbi, A. Stochastic Models for Turbulent Particle Dispersion in General Inhomogeneous Flows; Paul Scherrer Institut: Villigen PSI,
Switzerland, 2008.
58. Pope, S.B. Turbulent Flows. Meas. Sci. Technol. 2001, 12, 2020–2021. [CrossRef]
Fluids 2022, 7, 337 26 of 26

59. Peirano, E.; Chibbaro, S.; Pozorski, J.; Minier, J.P. Mean-field/PDF numerical approach for polydispersed turbulent two-phase
flows. Prog. Energy Combust. Sci. 2006, 32, 315–371. [CrossRef]
60. Durbin, P.A. A Reynolds stress model for near-wall turbulence. J. Fluid Mech. 1993, 249, 465. [CrossRef]
61. Hanjalić, K.; Launder, B.E. A Reynolds stress model of turbulence and its application to thin shear flows. J. Fluid Mech. 1972, 52,
609–638. [CrossRef]
62. Pope, S.B. Turbulent Flows; Cambridge University Press: Cambridge, MA, USA, 2015.
63. Kasper, R.; Turnow, J.; Kornev, N. Multiphase Eulerian–Lagrangian LES of particulate fouling on structured heat transfer surfaces.
Int. J. Heat Fluid Flow 2019, 79, 108462. [CrossRef]

You might also like