ArticlePDF Available

Abstract and Figures

A combined experimental and numerical study is performed to investigate the flow field and associated aerodynamic forces on a cambered airfoil. The Reynolds number is low enough to ensure importance of viscous dynamics, and high enough so that instability and transition to turbulence can occur. The flow fields are complex and their correct description is essential in understanding the nonlinear curves describing the variation of lift and drag coefficients with angle of attack, α. As α is increased from 0, the flow states go through a number of qualitatively distinct phases. At low to moderate α, the laminar boundary layer separates before the trailing edge, and as the separation point moves forward, instabilities of the detached shear layer form coherent vortices over the upper (suction) surface. At a critical angle, αcrit, instabilities in the shear layer grow fast enough to transition to turbulence, which then leads to reattachment before the trailing edge. In this flow state, lift is increased and drag decreases. Hence, in order to understand the aerodynamics at this scale, we need to understand the viscous dynamics of the boundary layer, as elegantly described and analyzed by Frank White.
Content may be subject to copyright.
Phys. Fluids 33, 033326 (2021); https://doi.org/10.1063/5.0039787 33, 033326
© 2021 Author(s).
Lagrangian models of particle-laden
flows with stochastic forcing: Monte
Carlo, moment equations, and method of
distributions analyses
Cite as: Phys. Fluids 33, 033326 (2021); https://doi.org/10.1063/5.0039787
Submitted: 06 December 2020 . Accepted: 22 February 2021 . Published Online: 26 March 2021
Daniel Domínguez-Vázquez, Gustaaf B. Jacobs, and Daniel M. Tartakovsky
COLLECTIONS
Paper published as part of the special topic on In Memory of Edward E. (Ted) O’Brien
ARTICLES YOU MAY BE INTERESTED IN
Deep-learning of parametric partial differential equations from sparse and noisy data
Physics of Fluids 33, 037132 (2021); https://doi.org/10.1063/5.0042868
Overstable rotating convection in the presence of a vertical magnetic field
Physics of Fluids 33, 034130 (2021); https://doi.org/10.1063/5.0035555
Direct numerical simulation of bidisperse inertial particles settling in turbulent channel flow
Physics of Fluids 33, 023305 (2021); https://doi.org/10.1063/5.0035656
Lagrangian models of particle-laden flows with
stochastic forcing: Monte Carlo, moment
equations, and method of distributions analyses
Cite as: Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787
Submitted: 6 December 2020 .Accepted: 22 February 2021 .
Published Online: 26 March 2021
Daniel Dom
ınguez-V
azquez,
1,a)
Gustaaf B. Jacobs,
1,b)
and Daniel M. Tartakovsky
2,c)
AFFILIATIONS
1
Department of Aerospace Engineering, San Diego State University, San Diego, California 92182, USA
2
Department of Energy Resources Engineering, Stanford University, Stanford, California 94305, USA
Note: This paper is part of the special topic, In Memory of Edward E. (Ted) O’Brien.
a)
Electronic mail: ddominguezvazquez@sdsu.edu
b)
Author to whom correspondence should be addressed: gjacobs@sdsu.edu
c)
Electronic mail: tartakovsky@stanford.edu
ABSTRACT
Deterministic Eulerian–Lagrangian models represent the interaction between particles and carrier flow through the drag force. Its analytical
descriptions are only feasible in special physical situations, such as the Stokes drag for low Reynolds number. For high particle Reynolds and
Mach numbers, where the Stokes solution is not valid, the drag must be corrected by empirical, computational, or hybrid (data-driven)
methods. This procedure introduces uncertainty in the resulting model predictions, which can be quantified by treating the drag as a random
variable and by using data to verify the validity of the correction. For a given probability density function of the drag coefficient, we carry out
systematic uncertainty quantification for an isothermal one-way coupled Eulerian–Lagrangian system with stochastic forcing. The first three
moment equations are analyzed with a priori closure using Monte Carlo computations, showing that the stochastic solution is highly non-
Gaussian. For a more complete description, the method of distributions is used to derive a deterministic partial differential equation for the
evolution of the joint PDF of the particle phase and drag coefficient. This equation is solved via Chebyshev spectral collocation method, and
the resulting numerical solution is compared with Monte Carlo computations. Our analysis highlights the importance of a proper approxi-
mation of the Dirac delta function, which represents deterministic (known with certainty) initial conditions. The robustness and accuracy of
our PDF equation were tested on one-dimensional problems in which the Eulerian phase represents either a uniform flow or a stagnation
flow.
Published under license by AIP Publishing. https://doi.org/10.1063/5.0039787
I. INTRODUCTION
Particle- and droplet-laden flows occur in many anthropogenic
and natural environments. For example, the mixing of liquid fuel spray
and/or solid fuel particles with turbulent gas flows determines the effi-
ciency of many propulsion and energy systems. Environmental pollu-
tion is affected by the dispersion of aerosol particles in environmental
carrier air flow.
The Eulerian–Lagrangian (EL) method provides a natural frame-
work for the modeling of particle- and droplet-laden flows. It uses
Eulerian continuum models to describe the dynamics of the ambient
flow and tracks individual particles along their Lagrangian paths.
1
A “first-principle” EL approach accounts for the surface of a
finite-size particle and computes the flow over each particle. The
excessive computational cost involved in these high-resolution (aka
high-fidelity) simulations limits the number of particles that can be
dealt with on modern computational infrastructure to hundreds.
2–4
Reduced-complexity models, such as the force coupling method
5
and
filtered particle method,
6
can reduce the computational cost and
increase the computationally feasible number of particles. The accu-
racy of these methods deteriorates near walls, which limits their range
of applicability.
Process-scale problems involve a number of particles on the
order of millions to billions. To handle such numbers, it is common to
treat each particle as a point of zero volume, e.g., according to the
Particle-Source-In-Cell (PSIC) method.
1
This method models par-
ticles’ interaction with the carrier flow via singular point source terms
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-1
Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf
that account for the momentum and energy exchange. The momen-
tum and energy exchange between the point particles and the flow is
determined by the (drag) force and heat transfer between the particle
and the carrier phase. Analytical expressions for the force and heat
transfer coefficient are available only for a few physical conditions,
with limited application range.
7,8
Thus, the widely used Stokes drag
law
9
is valid for steady incompressible flows at low Reynolds numbers
and for spherical particles. The Maxey-Riley (MR) relation
10
accounts
for unsteady effects, but it applies only to low Reynolds number and
spherical particles. More general flow conditions with arbitrary
Reynolds and Mach numbers require empirical corrections to the
Stokes drag,
11,12
number density, slip coefficients,
13
or viscosity ratios
for droplets.
14
Treatment of the energy exchange between a point par-
ticle and the ambient flow is analogous. For low Reynolds number
flows over a spherical particle, the heat transfer coefficient is computed
analytically; to account for high Reynolds, Mach, and/or Nusselt num-
bers, one has to resort to a correction factor.
7,15
The empirical corrections depend on plethora of parameters,
such as particle shape and the Reynolds and Mach numbers, that have
nonlinear effects on the flow around a particle. This naturally trans-
lates into a prediction error of a point-particle model as the momen-
tum and energy exchange are only known within certain bounds. The
twin goals of alleviating this modeling inaccuracy and expanding its
applicability range have inspired multiscale and data-driven modifica-
tions to the PSIC method. Thus, the multiscale method
2,3,16–19
connects accurate high-resolution simulations with the reduced point-
particle method through surrogate models. The latter approximates
the interphase force and heat ux with a surrogate model in a wide
parameter space using high-resolution simulations in a data-driven
manner. In regions of the parameter space with a large uncertainty,
additional high-resolution simulations are conducted to improve the
accuracy and/or validity range of the surrogate model. As more high-
resolution simulations become available, the multiscale method is
updated via a Bayesian procedure. This procedure yields the stochastic
forcing with a computable PDF.
The presence of such a stochastic forcing in both the empirical and
data-driven approaches renders solutions to the corresponding PSIC
model random as well. These solutions are given in terms of a joint
probability density function (PDF) of system states or their ensemble
moments. Monte Carlo (MC) simulations are often used to obtain such
solutions. They are easy to implement, “embarrassingly” parallel, and
free of distributional assumptions; their only approximation stems from
the practical need to rely on a finite number of MC realizations, N
s
,to
compute the sample statistics. A drawback of the MC method is its slow
convergence: its sampling error decays as 1=ffiffiffiffi
Ns
p.ThiscanmakeMC
simulations prohibitively expensive if each realization takes a long time
to compute. Various modifications of the standard MC, e.g., multilevel
MC, pseudo-MC, or combinations thereof, can significantly accelerate
this convergence rate, but their performance is not guaranteed especially
when the goal is to compute the full PDF rather than its moments.
20,21
Other sampling-based methods, such as stochastic collocation, require
nontrivial modifications
22
in order to handle hyperbolic problems with
discontinuities. When the stochastic dimension and/or the noise
strength become large, such methods might become slower than the
standard MC even for problems with smooth solutions.
23
Direct numerical alternatives to sampling techniques include
methods based on (generalized) polynomial chaos expansions. These
methods represent uncertain inputs and state variables by orthogonal
polynomials of standard random variables and often exhibit spectral
accuracy. Of direct relevance to high-speed compressible flows
described by the Euler equations with shocks is the multi-element gen-
eralized polynomial chaos method,
24–26
which accommodates the
presence of discontinuities in the stochastic space. Its computational
cost might become comparable to sampling methods,
27
specifically
when the stochastic dimension is large. Like their sampling-based
counterparts, the direct simulation methods do not provide physical
insight into the probabilistic behavior of a system, e.g., the spatiotem-
poral nonlocality of the statistical moments
28
and PDFs
29
of the sys-
tem states.
The method of moments (MoM) alleviates some of these disad-
vantages by deriving deterministic equations for the statistical
moments of a system state. Since the MoM is free of polynomial
expansions, it does not suffer from the “curse of dimensionality,” but
it often requires closure approximations to be computable. It has been
used to derive moment equations for high-speed flows interacting
with a particle phase;
17
the closure terms were learned from the MC
simulations. Practical considerations limit the MoM to the derivation
andsolutionofequationsforthefirsttwomomentsmeanand
(co)variance—of a system state, which limits its usefulness for highly
non-Gaussian phenomena. Specifically, the MoM cannot capture rare
events occuring in such problems, which are characterized by fat-
tailed PDFs.
The method of distributions (MoD) overcomes this limitation by
deriving a deterministic equation for either the joint PDF or the joint
cumulative density function (CDF) of the system states. Having its ori-
gins in the statistical theory of turbulence,
30
the MoD has since been
used as an efficient uncertainty quantification technique.
31
It retains
all the advantages of the MoM, but it might require closure approxi-
mations. The MoD yields a closed-form PDF/CDF equation for non-
linear flows in the absence of a shock, e.g., those described by the
inviscid Burgers equation
32
and the shallow-water equations.
33
Within
the MoD framework, shocks and discontinuities can be treated either
analytically, as was done for the Buckley–Leverett equation
34
and
water hammer equations,
35
or by adding the PDF/CDF equation a
kinetic defect-like source term that generally has to be learned from
Monte Carlo runs.
36
Numerical solutions of PDF/CDF equations
admit high-order spectral accuracy
32
and can be up to two orders of
magnitude faster than the standard MC.
37
We deploy the MoD to describe isothermal particle-laden flows
driven by stochastic forcings. The underlying flow model relies on the
Lagrangian point-particle formulation with one-way coupling between
fluid flow and particle transport. The drag on a particle is modeled as
a random variable with a prescribed PDF. The MoD yields a closed-
form partial differential equation for the joint PDF of a particle’s posi-
tion and velocity. We consider two canonical flow scenarios, both in
one spatial dimension: a uniform carrier flow and a stagnation carrier
flow. These are important in their own right and can be used as build-
ing blocks of more general and multi-dimensional flows; for example,
the stagnation flow is a central component to the dynamic description
of attractors and repellers in dynamic systems.
38
Our PDF solutions
are validated against high-fidelity MC simulations and compared with
solutions of the moment equations.
17
ThehyperbolicPDFequationis
solved via the Chebyshev collocation method.
32
Discontinuities in its
solution are captured using the filtering techniques.
39,40
A key result of
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-2
Published under license by AIP Publishing
our analysis is the derivation of analytical expressions for the position
and velocity of a particle moving in deterministic uniform and stagna-
tion flows. These expressions allow us to generate sufficiently large
numbers of MC realizations. In both flow regimes, the PDF solutions
are non-Gaussian and their moments can increase or decrease
depending largely on the time-dependent increase or decrease in the
inter-phase velocity. Moreover, the stochastic solution can develop dis-
continuities at inflection points of the inter-phase velocity.
In Sec. II, we formulate an Eulerian–Lagrangian stochastic model
of particle-laden flows. The formulation is presented in one-
dimension but applies to multiple spatial dimensions as well. Its statis-
tical treatment is presented in Sec. III via the MoD (Sec. III A)andthe
MoM (Sec. III B). Our numerical strategy for solving the resulting
PDF equation is described in Sec. IV. We apply our methodology to
the uniform and stagnation flows, for which the analytical solutions
are derived in Sec. V.InSec.VI, we use these two canonical settings to
analyze the accuracy and robustness of the MoD and MoM solutions.
Section VII contains concluding remarks and future work.
II. LAGRANGIAN PROBLEM FORMULATION
Dynamics of an isothermal collisionless particle phase in a one-
way coupled unidimensional Eulerian–Lagrangian system with the
point-particle approximation is governed by
1,41
dxp
dt¼up;(1a)
dup
dt¼/uup
sp
:(1b)
Here, tis the non-dimensional time, x
p
is the non-dimensional particle
position, and u
p
is the non-dimensional particle velocity. The non-
dimensional particle response time s
p
is a measure of the response of
the particle to a change in the carrier velocity u. It is expressed as
sp¼d2
pRe=ð18eÞ,wheredp¼d
p=Lis the non-dimensional particle
diameter, Re ¼U1L= is the Reynolds number, Lis a characteristic
length, U1is the reference velocity, and e¼q=qpistherelativeden-
sity ratio of the two phases.
The function /is used to correct the Stokes drag force for flow
conditions outside of Stokes assumptions. Such models for the cor-
rected drag coefficient /are empirical and, therefore, can only be
determined within an uncertainty bound.
2,3,42
For the sake of general-
ity, we postulate
32,42
that /depends on the relative velocity uup
/¼agðuupÞ;(2)
without specifying the functional dependence of the function gðÞ.
This function can be expanded in terms of several random modes.
Here, we consider the first of those random modes and take /¼a.
The random coefficient awith a given PDF faðAÞaccounts for the
uncertainty in /stemming from a broad range of sources, such as
uncertainty in the particle size, shape, or inexactness/empicism of
the drag force, and renders the system of ordinary differential equa-
tions (1) stochastic. Its solution is given in terms of the joint PDF
faxpupðA;Xp;Up;tÞ.
Our model formulation ignores inter-particle collisions. This is
justified if the particle phase is dilute, specifically in one spatial
dimension.
43
III. SOLUTION STRATEGIES
A. Method of distributions
When applied to Eq. (1), the MoD yields an exact PDF equation
(see Appendix A)
@faxpup
@tþ@
@Xp
Upfaxpup

þ@
@Up
AgðUUpÞ
spðUUpÞfaxpup
"#
¼0;
(3)
with A,X
p
,andU
p
denoting the deterministic versions of the stochas-
tic variables a,x
p
,andu
p
.Equation(3) describes the evolution of the
joint PDF of the particle phase and drag coefficient,
faxpupðA;Xp;Up;tÞ,inthephasespaceXspanned by coordinates
ðXp;Up;AÞ. This space can be either infinite or bounded,
X¼½X0
p;X1
p½U0
p;U1
p½A0;A1. In the latter case, (3) is subject to
boundary conditions for the independent variables X
p
and U
p
faxpupðA;Xb
p;Up;tÞ¼fXp
axpupðA;Up;tÞ;(4)
faxpupðA;Xp;Ub
p;tÞ¼fUp
axpupðA;Xp;tÞ:(5)
The boundary functions on the right hand side of these expressions
are specified according to the corresponding marginal distributions;
and using the characteristic velocities of (3) defined as in Eqs. (14) and
(15),Xb
p¼X0
por Xb
p¼X1
pfor CXðX0
pÞ>0andCXðX1
pÞ<0, respec-
tively. Similarly, Ub
p¼U0
por Ub
p¼U1
pfor CUðU0
pÞ>0and
CUðU1
pÞ<0, respectively. The PDF equation (3) is also subject to the
initial condition
faxpupðA;Xp;Up;0Þ¼f0
axpupðA;Xp;UpÞ:(6)
The function form of f0
axpupðA;Xp;UpÞis determined by the degree of
certainty in the initial state of the system, ðxp0;up0Þ. If it is known with
certainty, i.e., deterministic, then
faxpupðA;Xp;Up;0Þ¼faðAÞdðXpxp0ÞdðUpup0Þ;(7)
where dðÞ is the Dirac delta function. We will refer to this as a deter-
ministic initial condition (dIC). If the initial condition is not known
with certainty, then we refer to it as stochastic (sIC).
Once faxpupðA;Xp;Up;tÞis computed from (3)–(6),thePDFs
fxpupðXp;Up;tÞ;fxpðXp;tÞ,andfupðUp;tÞare computed as its margin-
als via numerical integration over the respective variables (see
Appendix A).
B. Method of moments
Solutions of the moment equations have been used to elucidate
many salient features of stochastically forced particle-laden flows.
17
We summarize that analysis and extend it to derive third-moment
equations in order to understand the degree of non-Gaussianity. The
derivation starts by using the Reynolds decomposition to represent all
parameters and state variables as the sums of their respective ensemble
means (denoted by the overbar) and zero-mean fluctuations (denoted
by the prime), e.g., xp¼
xpþx0
pwith x0
p¼0. Substituting these
decompositions into (1) and taking the ensemble average, we obtain
equations for the means
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-3
Published under license by AIP Publishing
d
xp
dt¼
up;(8)
sp
d
up
dt¼
/ð
u
upÞþ/0u0/0u0
p;(9)
for the variances, r2
xp¼x02
pand r2
up¼u02
p,
dr2
xp
dt¼2x0
pu0
p;(10)
sp
2
dr2
up
dt¼
/ðu0u0
pr2
upÞþ/0u0
pð
u
upÞþ/0u0u0
p/0u02
p;(11)
and for the third central moments, sxp¼x03
pand sup¼u03
p,
dsxp
dt¼3x02
pu0p;(12)
sp
3
dsup
dt¼
/ðu0u02
psupÞþ/0u02
pð
u
upÞ
r2
upð/0u0/0u0
pÞþ/0u0u02
p/0u03
p:(13)
As opposed to the exact PDF equation (3), these moment equa-
tions are not closed since they contain unknown mixed, higher-order
moments. To render them computable, one has to introduce closure
approximations such as the aprioriclosure
41,42
used to analyze the
first two statistical moments or a posteriori closure as used in Eulerian
formulations.
43,44
The moment equations (9)–(13) provide insight into the devia-
tion of the stochastic solution from its deterministic counterpart and/
or general dynamics of the moments. For example, the mean dynam-
ics, described by (8) and (9), differs from the solution of the determin-
istic problem (1) with the mean value of the random parameter
a.The
difference between the deterministic equation and the averaged equa-
tion is the correlation /0u0and /0u0
p. In some special cases, e.g., when
thecarrierflowisconstantandtherelativevelocityiszero(u¼u
p
),
the mean of the solution is the same as the deterministic solution for
a; the velocity deviation decreases since the right hand side of (11)
contains only the damping term.
Finally, we note that the moment equations (9)–(13) are related
to the PDF equation (3) as they represent the evolution of the first
three moments of fxpand fup.
C. Singularities in the stochastic solution
The characteristic velocities which can be directly inferred
from (3),
CX¼dXp
dt¼Up;(14)
CU¼dUp
dt¼AgðUUpÞ
spðUUpÞ;(15)
can be different which can lead to a crossover of the characteristics’
paths at certain values of A. In general, hyperbolic systems when char-
acteristics cross, a discontinuity is expected to appear in the solution.
Depending on the sign of the relative velocity, UUp, we identify
two settings in which the resulting discontinuities appear in the joint
PDF, faxpupand its marginals. First, for a positive (and constant) rela-
tive velocity [Fig. 1(a)], we consider a cloud of N
a
particles with
uniformly spaced different drag coefficients A
i
(i¼1…Na)suchthat
Aiþ1>Ai. The particle with the greatest forcing, ANa(rightmost par-
ticles), moves fastest, whereas the leftmost particle with a slower
response is left behind. As a result the cloud expands. For a nonlinear
relative velocity, the characteristics could steepen and cross in the
expansion, yielding discontinuities (not exhibited in the graph).
For a second setting, the same initial cloud is considered but for a
negative (and constant) relative velocity, which causes the cloud of
particles to compress. At some point, the leftmost particles overtake
the rightmost particles and the cloud concentrates in a reduced region
or even in a singular point. At that instant, the characteristics of the
hyperbolic system (3) cross. If all of them cross in a single point, then
the PDF solution becomes the Dirac delta distribution. After this sin-
gular event, the cloud expands and asymmetry can reemerge, resulting
in steepening of the left side of the PDF fxpand in its discontinuity, as
it did for the positive relative velocity.
Consistent with the formation of discontinuities in the marginal
PDF, fxp, discontinuities also arise in the marginal PDF of the particle
velocity, fup.InSec.V, we illustrate these phenomena by analyzing the
uniform and stagnation carrier flows with stochastically forced particle
dynamics.
IV. NUMERICAL IMPLEMENTATION
The discontinuities and sharp gradients that can appear in the
solution of the PDF equation (3) require special numerical treatment.
We use a low-dispersive/diffusive Chebyshev collocation method to
approximate the derivatives with respect to X
p
and U
p
.Suchspectral
treatment was shown to be effective or even necessary to solve similar
moment equations in Ref. 17. We also deploy the filtering and
FIG. 1. Evolution of the PDF of x
p
of a cloud of particles initially distributed uni-
formly in space and traveling at the same initial velocity (from left to right) with dif-
ferent drag coefficients A
i
such that Aiþ1>Ai. Under the influence of positive
relative velocity in (a) and negative relative velocity in (b).
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-4
Published under license by AIP Publishing
regularization techniques designed to capture discontinuities and regu-
larize singularities in a spectral solution while preserving
accuracy.
17,40,45,46
A. Chebyshev collocation method and time
integration
The Chebyshev collocation method, extensively described in the
textbooks,
47,48
is briefly summarized below for the sake of complete-
ness. We do so for one spatial dimension; the multi-dimensional for-
mulation builds upon that as it is defined along lines on a tensorial
grid. In the Chebyshev collocation method, a function y(x) is approxi-
matedbyaChebyshevinterpolantas
yNxðxÞ¼X
Nx
j¼0
yðxjÞljðxÞ;ljðxÞ¼ Y
Nx
k¼0;kj
xxk
xjxk
:(16)
Here j¼0;;Nx;ljðxÞis the Lagrange polynomial of degree N
x
.The
collocation points are chosen as the Gauss-Lobatto quadrature points
ni¼cos ðip=NxÞ;i¼0;;Nx(17)
such that the L1norm of the interpolant is minimized on the interval
½1;1.
The derivative of the function y(x) at points x
i
is approximated
by
@y
@xðxiÞX
Nx
j¼0
yðxjÞl0
jðxiÞ;(18)
with l0
jthe derivative of the corresponding Lagrange polynomial. This
approximated derivative is recast in the matrix-vector form
y0¼Dy;(19)
where the differentiation matrix Dhas components Di;j¼l0
jðxiÞ.
The multi-dimensional PDF equation (3) is discretized on a ten-
sorial grid that spans X
p
,U
p
,andAin the domain X.Thespectral
approximation of the distribution function ~
f¼~
fNANXpNUpðA;Xp;UpÞ
on this grid is governed by the semi-discrete equation
d~
f
dtþDXpFXpþDUpFUp¼0;(20)
where the entries of the flux arrays are given by
FXp
i;j;k¼Upi;j;k~
fi;j;k;(21)
FUp
i;j;k¼Ai;j;kgðUi;j;kUpi;j;kÞ
spðUi;j;kUpi;j;kÞ~
fi;j;k;(22)
with counters i,j,kalong the tensors. The matrices DXpand DUpare
the scaled versions of the matrix Dwith the following entries:
DXp
ðÞ
m;j¼@n
@Xp
Dm;j
ðÞ
;DUp
ðÞ
m;k¼@n
@Up
D
ðÞ
k;m(23)
with @n=@Xp¼2=ðXmax
pXmin
pÞand @n=@Up¼2=ðUmax
pUmin
pÞfor
the one-dimensional case. The matrix-vector multiplication DXpFXp
and DUpFUpis performed along grid lines with the counters jand k,
respectively, in (21) and (22). The carrier flow velocity Uis specified at
the particle locations. The semi-discrete system is integrated in time
with the total variation diminishing (TVD) Runge-Kutta scheme.
49
To obtain the marginals, fxpand fupare obtained via the numeri-
cal integration of faxpupalong Aand either U
p
or X
p
,respectively.This
is done via Clenshaw-Curtis quadrature in U
p
and X
p
and via the trap-
ezoidal rule in A. Because the distribution equation does not have
terms with derivatives respect to A, the spectral approximation is not
necessary in this direction.
B. Regularization of Dirac delta function
The numerical solution of the PDF equation (3) with the deter-
ministic initial state (7) requires an approximation of the Dirac delta
function dðÞ. We rely on the kernel that regularizes dðÞ with a class
of high-order, compactly supported polynomials,
45
dm;k
eðxÞ¼ e1Pm;kðx=eÞ;x2e;e
½
0;otherwise;
(24)
where e>0 is the support width or scaling parameter. On the com-
pactly supported interval, the regularized delta function integrates to
unity (i.e., the zeroth moment is one). The polynomial Pm;kis designed
to have the first up to the m
th
moment vanished and to have kcontin-
uous derivatives at the endpoints of the compact support. For it to be
possible for the moments to vanish the regularized delta is permitted
to have negative values on its supported interval. The vanishing
moments ensure that the regularized Dirac delta kernel (a so-called
delta sequence) converges to the exact Dirac delta function at a rate of
Oðemþ1Þ.Thismomentpropertyisnecessaryfortheconstructionof
high-order approximations of singular Dirac delta source terms in
spectral approximations of PDEs as was shown in Ref. 45. To preserve
high-order spatial accuracy, it was further shown that the optimal
value for the compact support must be e¼Nk=ðmþkþ2Þ
x.Thecompact
kernel dm;k
eðxÞin (24) has a maximum at its center. To achieve high-
order accuracy, one has to relax positivity of the kernel, leading to the
undershoots in Fig. 2.
For the approximation of the initial Dirac delta distribution func-
tion in (7), the vanishing moments of the regularized delta function
yield an accurate representation of the zero moments of the determin-
istic initial state. Thus, in that case the regularized Dirac delta provides
both spatial accuracy and the correct statistical properties of the distri-
bution function at the initial time.
A naive alternative is to approximate dðxÞvia a Gaussian PDF
dðxÞ 1
ffiffiffiffiffi
2p
prx
exp x2
2r2
x
"#
;(25)
with small variance r2
x. The Gaussian PDF, however, has no vanishing
moments and can thus not yield high-order approximations to the
Dirac delta. If the initial state is random, than the Gaussian distribu-
tion does correctly represent uncertainty in the initial state of the
system.
C. Filtering for capturing discontinuities
Since (3) admits singularities, we have to regularize these singu-
larities in numerical approximations to avoid numerical instabilities.
To this end, we once again resort to the regularized Dirac delta kernel
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-5
Published under license by AIP Publishing
(24). This time the kernel serves as a convolution filter kernel as dis-
cussed in Ref. 40 to smoothen a function y(x) as follows:
~
yðxÞ¼ðxþe
xe
yðsÞdm;k
eðxsÞds:(26)
Using a quadrature rule for approximation of the convolution integral,
the interpolant yNxis filtered as
~
yNxðxÞ¼ðxþe
xe
dm;k
eðxsÞX
Nx
i¼0
yðxiÞliðsÞds¼X
Nx
i¼0
yðxiÞSiðxÞ:(27)
ThediscretefilterS
i
is defined as
SiðxÞ¼ðxþe
xe
liðsÞdm;k
eðxsÞds:(28)
In vector notation, (27) and (28) take the form
~
y¼Sy:(29)
The extension to tensorial form is straightforward. This convolution
filter was shown in Ref. 40 to smoothen shock discontinuities while
providing high-order accurate resolution away from shocks. In some
cases, a weak exponential filter
32
is needed to remove high-frequency
numerical noise that appears in regions near the boundaries.
D. Monte Carlo simulations
The PDF and the moments of the PDF can be computed with a
MC approach. In MC, realizations of x
p
and u
p
are computed by solv-
ing (1) with random coefficients adrawn from a given PDF faðAÞ.
Here, we use analytical solutions that will be discussed in Sec. Vand
that allow for a computationally efficient determination of a significant
number of MC realizations, N
s
. In all the tests considered, we found
Ns¼105realizations to be sufficiently accurate yielding a normalized
error of the third moment less than 0.1%. The kernel density estima-
tion, implemented in the Matlab 2019b subroutine kdensity,
determines the PDFs fxpðXp;tÞ;fupðUp;tÞ,andfxpupðXp;Up;tÞ.The
unknown correlation terms in the moment equations (8)–(13) are
closed using MC realizations. The resulting a priori closed moment
equations are integrated in time via the fourth-order Runge-Kutta
(RK4) scheme.
V. TWO CANONICAL PARTICLE-LADEN FLOWS
We consider two one-way coupled particle-laden flows—a uni-
form flow and an inviscid stagnation flow—for which the carrier phase
velocity is described by analytical expressions. These are both impor-
tant in their own right and serve as building blocks for more complex
flows. Both flows admit analytical solutions for the corresponding
particle-laden flow with constant deterministic forcing, /¼constant.
While this particle solution for the uniform carrier flow is well known,
we are not aware of an analytical solution to the particle-laden stagna-
tion flow. Analytical solutions are derived for both ows in Secs. VA
and VB.
A. Uniform flow
By its definition, a uniform carrier flow is characterized by a con-
stant velocity field u. To derive the analytical solution, we cast the par-
ticle transport equations (1) with the constant uand the initial
conditions xpð0Þ¼xp0and upð0Þ¼up0into the following linear sys-
tem of ODEs:
d
dt
xp
up

¼01
0b

xp
up

þ0
bu

;ba
sp
:(30)
The analytical solution of this system is
xpðtÞ¼xp0þut þ1
bðuup0Þebt 1
ðÞ
;(31)
upðtÞ¼uþðup0uÞebt:(32)
Details of the derivation are provided in Appendix B.
The solution is plotted in Fig. 3 and shows that the response of
the particle initially at rest to a fluid velocity is slower with increasing
b,i.e.,withincreasingeffectiveinertia.Hence,foragivens
p
,higher
values of the correction parameter adecrease the particle’s time
response. At long times on the order of Oð1=bÞ,theparticlevelocity
becomes equal to the carrier flow velocity u. When the relative velocity
between the particle and the carrier phase (also called interphase veloc-
ity) becomes zero, the particle position increases linearly at its constant
advection rate u.
B. Stagnation flow
The stagnation carrier velocity eld, u¼ðu;vÞ>, is given by the
Hiemenz analytical solution for an inviscid, irrotational ow
50
in the
domain x1;0as follows:
u¼kx;v¼ky;
where yis the coordinate perpendicular to the flow direction, and kis
a constant. (The viscous boundary layer solution near a wall at x¼0is
available as well.
51
It predicts the boundary layer thickness of
d¼ffiffiffiffiffiffiffi
=k
p, too thin to affect the particle dynamics.)
Along the centerline y¼0, the flow is one-dimensional with a
stagnation point at x¼0andvelocity
FIG. 2. Regularization of the Dirac delta function, dm;k
e, in comparison with a
Gaussian PDF.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-6
Published under license by AIP Publishing
u¼kx:(33)
With the carrier velocity at the particle location x
p
is u¼kxp(1) can
be cast into a linear dynamic system,
d
dt
xp
up

¼01
kb b

xp
up

:(34)
The analytical solution to this system is derived in Appendix B and is
characterized by the eigenvalues of the 22 matrix in (34)
k1¼bffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
bðb4kÞ
p2;k2¼bþffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
bðb4kÞ
p2:(35)
Their real and imaginary parts are plotted in Fig. 4,forb0;8and
k¼1. For 0 <b<4k, the eigenvalues are imaginary with negative
real part. In that case, the solution of the system (34), i.e., the particle
phase solution, is well-known to tend toward an inward spiral in the
phase plane as plotted in Fig. 5(c). Before it reaches the spiral singular-
ity, however, the particle will have crossed the x¼0 line where the
wall is located. This is, of course, not possible and the particle trace has
to terminate at x¼0. Alternatively, we can interpret the solution as a
physical solution of a particle trajectory in an opposed jet carrier flow.
For b>4k, both eigenvalues are real and negative, in which case the
particle moves toward an inward node in the phase space xpvp.A
bifurcation in particle dynamics from a spiral to a node occurs at
b¼4k.Figure 5(c) shows the particle phase when the stagnation point
is an inward node and an inward spiral for two different initial
conditions.
The analytical solutions for the particle’s position and velocity,
xpðtÞand upðtÞ,vstimeareplottedinFigs. 5(a) and 5(b).Theparticle
reaches the stagnation point for any forcing b¼a=sp.Thecollisionof
the particle with the wall for the stangation flow case is indicated by
the red dot in the graphs 5(a),5(b),and5(c).
C. Impact of stochastic forcing
The effect of a stochastic forcing on the particle-laden uniform
and stagnation flows is studied for the cases and parameters collated
in Table I. For both flow regimes, we consider particles initialized at
rest. For the stagnation flow, we also consider the particles initialized
according to the carrier flow velocity.
For each of these cases, we consider three PDFs, faðAÞ,for
the random variable ain the drag correction factor defined in (2)
including a uniform, normal, and beta distribution, all with the same
mean l
a
andstandarddeviationr
a
(Fig. 6). For the stagnation flow,
faðAÞis selected to have a non-zero probability in the interval
0<a=sp<4kto ensure that all particles reach the wall at a finite
time (according to the deterministic solution). Also investigated is the
effect of deterministic vs stochastic initial conditions.
VI. SIMULATION RESULTS AND DISCUSSION
A. Uniform flow: Monte Carlo results
The PDFs fxpðXp;tÞand fupðUp;tÞobtained via MC solution of
(1) for the uniform flow with the uniform forcing distribution faðAÞ
and deterministic initial conditions are depicted in Fig. 7. Starting
from the deterministic Dirac delta initial distribution, both fxpðXp;tÞ
and fupðUp;tÞfirst widen over time, while showing a skewness, i.e., a
FIG. 3. Time dependence (a) and phase space (b) of the particle dynamics in the
constant uniform carrier flow.
FIG. 4. Imaginary and real part of the eigenvalues k
1
and k
2
in (35) for b0;8.
The circle corresponds to b¼0, and the diamond and square correspond to b¼8
for k
1
and k
2
, respectively.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-7
Published under license by AIP Publishing
bias, toward the upper range of the X
p
and U
p
values, where more par-
ticles accumulate. This bias reflects the particles’ asymptotic behavior
in the limit of an infinite response time, sp!1in which case all the
particles congregate on a step function in time. After the initial widen-
ing, the velocity distribution narrows with time as the particles’ veloc-
ity settles to the uniform carrier flow velocity. The temporal evolution
of the PDFs has a characteristic time scale on the order Oðsp=
aÞ.At
later times, the velocity distribution returns to the Dirac delta and the
corresponding position distribution is advected at constant velocity u
without changes in time.
The means
xpðtÞand
upðtÞ, plotted with their corresponding two
standard deviation bandwidths in Fig. 8, tell a similar story. The mean
particle velocity
upðtÞincreases from its zero initial state and settles to
the constant carrier velocity at t!1. Associated with the accelera-
tion and settling is an initial increase in the velocity bandwidth that
then returns to zero at later times. Consistent with the velocity band-
width, the position bandwidth grows initially and then remains con-
stant when the particles settle.
Per definition, and as confirmed by Fig. 8, the mean of the solu-
tion must be contained in the interval of deterministic limit trajecto-
ries. Moreover, because xpðaÞand upðaÞare monotonically increasing
with a, it follows that
xp2min
axpðaminÞ;xpðamax Þ

;max
axpðaminÞ;xpðamax Þ

hi
;
up2min
aupðaminÞ;upðamax Þ

;max
aupðaminÞ;upðamax Þ

hi
;
where amin >0andamax denote a minimum and maximum value of a.
This suggests that
xpxpð
aÞand
upupð
aÞ,i.e.,themeansolution
is equal to the deterministic solution at the mean stochastic forcing.
The moment equations provide further insight. Because of the
correlation terms /0u0and /0u0
p,thegoverningequationsforthe
FIG. 5. Solutions for particles released at rest (a) and at flow condition (b) in a
stagnation flow with k¼1 for different values of parameter b. The phase space plot
in (c).
TABLE I. Flow regimes and parameter values considered in the simulations.
Test case xp0up0us
p
Uniform flow, particle launched at rest (UF) 0 0 1 0.25
Stagnation flow, particle launched at rest (SFR) –1 0 xp1
Stagnation flow, particle launched at flow
conditions (SFF)
–1 1 xp1
FIG. 6. Uniform, normal, and beta (U;N;B) PDFs selected for the random
parameter a. All three PDFs have the same mean la¼1 and standard deviation
ra¼0:2, i.e., aU½1ffiffiffiffi
12
p=2;1þffiffiffiffi
12
p=2;aN½1;0:2, and
aB½2;3þ0:6.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-8
Published under license by AIP Publishing
mean position and velocity in (8) and (9), respectively, are different
from the deterministic equations (1) with a¼
a. But the term /0u0is
zero for the uniform flow case because u0¼0. Moreover, the correla-
tion term /0u0
pin (9) is negligible, but not zero. Thus, to a first approx-
imation,
xpxpð
aÞand
upupð
aÞ. For a random solution with a
uniform stochastic forcing distribution, the root mean square differ-
ence over the time interval is 0.0073 for the position and 0.0062 for the
velocity.
With a zero carrier phase velocity perturbation, u0¼0, many of
the correlation terms in the second central moment or variance of the
velocity are also zero or negligible. Significant terms that remain are a
damping term
/r2
upand the source term, /0u0
pð
u
upÞ.Thelatter
is positive because the relative velocity is positive, ð
u
upÞ>0, and
because /0¼a0and u0
phave the same sign since the particle velocity
u
p
is monotonically increasing with respect to the forcing /¼a.The
positive source term is maximum initially and decreases as the particle
velocity settles to the flow conditions. The damping term reduces the
velocity variance to zero in the limit t!1. Correspondingly, the
PDF fuptends to the Dirac delta distribution [Fig. 7(b)]. The combina-
tion of the temporal damping and forcing by the positive source leads
to a maximum variance at times that are on the order of Oðsp=
aÞ.The
particle position variance depicted in Fig. 9(a) shows an initial increase
consistent with the increasing velocity variance and an increased
spreading of random particle trajectories. When the particles settle to
the constant carrier flow condition, all trajectories are advected at con-
stant velocity. After that time, the particle variance no longer changes.
In Fig. 9(a), the variance of the particle velocity and position are
plotted vs time for three different forcing distributions f
a
(uniform,
normal, and beta). The temporal trends for the different stochastic
forcing are very similar because the mean forcing and its variance are
chosen to be the same for the three forcing distributions. The damping
term in the velocity variance equation, which depends on the mean
forcing and velocity variance only, is therefore not affected by the
shape of the forcing distribution. The source correlation term, how-
ever, is directly dependent on the forcing fluctuations, /0,whichleads
to differences in the velocity and position variances for different shapes
of the forcing distribution of up to 15%.
The third central particle position and velocity moments evolve
in a similar way as the variances [Fig. 9(b)]. The third velocity
moment, sup, experiences a negative growth followed by an asymptotic
decay to zero (or the Dirac delta in the PDF sense as observed above).
The third position moment, sxp, first decreases and then asymptotically
evolves to a constant value. Both the minimum in supand the plateau
in sxpoccur at slightly later times as compared to the minimum in rup
and rxp. The difference in the factors 2/=spand 3/=spin Equations
(11) and (13) is assumed to be at least partially responsible for causing
this shift in the maximum. The similarity in variance and skewness
trends would suggest that the third moment dynamics might also be
primarily affected by a positive sourcing and a damping. To verify this,
the correlation terms in (13) are plotted vs time in Fig. 10.Clearly,the
damping term
/suphas a major influence on the long term response.
However, there is no single dominant source term. While the term
/0u02
pð
u
upÞplays a similar role as the positive source term in the
variance equation, the other correlation terms are not negligible and
contribute also. Surprisingly perhaps is that the term with fourth order
correlations, /0u03
p, is dominant, an indication that the tail behavior
of the solution PDF and tail behavior of the forcing function has a
FIG. 7. PDF of particle position (a) and velocity (b) for the UF test case carried out
with MC with a uniform forcing distribution.
FIG. 8. Two standard deviation interval along the mean for the test case UF with a
uniform forcing distribution. In dashed green, the particle velocity computed with
MoD and in black with MC. In dashed red, the particle position computed with MoD
and in blue with MC. Dark colors indicate dIC, whereas light ones indicate sIC.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-9
Published under license by AIP Publishing
considerable impact on the higher central moment solution of the
solution. This is confirmed by the deviation in the third central
moment evolution of up to 200% for different forcing distributions.
We plan to report further on the tail behavior of the PDF in the
near future.
The solution with stochastic initial condition (sIC) is plotted
in Figs. 8 and 9. It shows that while the trends in the position and
velocity mean and variance are similar to those determined with a
deterministic initial condition (dIC), the sIC solution is offset as
compared to the dIC solution. The offset is according to the initial
position deviation of rx¼0:05. Time integration of (10) from time
zero to a time tconfirms that exactly this term rxðt¼0Þappears at
the initial time, t¼0. The offset in Fig. 9 does not change signifi-
cantly over the time interval ½0;t, which implies that the term x0
pu0
p
in (10) is small. MC simulations confirm this and show that the
term has a maximum value of 0.002 over the time interval. Because
of the damping term, the velocity variance and third central
moment goes to zero in the asymptotic time limit for both deter-
ministic and stochastic initial conditions.
B. Uniform flow: Method of distributions
The solution of the governing equation for the PDF in (3) is grid-
resolved for the uniform flow case using a spectral grid with NXp
NUp¼300 300 collocation points and a uniform grid in Adirec-
tion with N
A
¼200. The CFL condition is set to 0.8. The Dirac delta
distribution function for dIC is regularized according to dk;m
ein (24)
with an optimal scaling e¼0:05, and m¼5 zero vanishing moments
and smoothness k¼2.
Figure 11 shows snapshots of fxpup(contours), fxp(left and bottom
axes), and fup(right and top axes) at three consecutive times. For refer-
ence, the mean of the particle phase solution (black line) is superposed
in the contour plot. At time t¼0, the marginals are initialized accord-
ing to the regularized Dirac delta as shown in Fig. 11(a).Atalater
time, t¼0.54, the joint PDF fxpuphas traveled along the mean in the
XpUpcoordinate system and has widened and deformed [contours
in Fig. 11(b)]. The marginal fxpand fupshow that the particles have a
bias toward the larger values of the position and the velocity. That is
consistent with the observations in the moments discussed previously;
because the particles with smaller response time, sp=A, travel a dis-
tance greater than the slower responding particles, they cluster at large
X
p
. Those fast particles furthermore reach their terminal settling veloc-
ity faster and thus there is a similar clustering in fup.Theconvexityof
the PDFs is an indication that the clustering is more pronounced
toward larger values. The schematic in Fig. 12 underscores this and
shows how the characteristic paths with non-constant advection veloc-
ity for different A
i
leads to a convex probability density.
At time t¼1.6, the velocity PDF has evolved toward a Dirac delta
function represented by a narrowly supported distribution centered at
U
p
¼1. The numerical solution successfully captures this PDF behav-
ior despite showing some minor fluctuations caused by Gibbs oscilla-
tions. The accuracy of the MoD solution at this time relies largely on
the number of vanishing moments mof the regularization of the Dirac
delta function at t¼0. Because the number of vanishing moments is
specified to be greater than five, the first up to the fifth moment is
accurately preserved even at times when the distribution function
FIG. 9. Computations for the UF case with MC in color lines and with the MoM in
black dots of the (a) second and (b) third central moments for the particle position
and velocity and the three PDFs considered for a(see Fig. 6) and dIC. It is included
the case of sIC for the uniform distribution. The legend in (b) is valid for (a) as well.
FIG. 10. Terms in Eq. (13) vs time tfor the UF case with dIC.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-10
Published under license by AIP Publishing
tends to the singular Dirac delta distribution. This accuracy preserva-
tion is confirmed by the results in Fig. 8 that compares the time evolu-
tion of the mean and variance determined with the MoD and MC
approach and that shows no discernible difference between the solu-
tions of the two approaches.
A few remarks on the accuracy of the numerical solution of Eq. (3):
Remark 1: Consistent with the findings in Ref. 17,theuseofhigh
order methods is necessary to compute an accurate solution of the
joint PDF faxpupsuch that the marginals determined according to
(A15)–(A17) are in good comparison with MC results. To underscore
the importance of numerical discretization, we compare the
Chebyshev spectral discretization with a first and a second order
upwind nite difference (FD) schemes. Figure 13(a) shows that the FD
schemes are overly dissipative as compared to the spectral method if
the same number of grid points are used. The root mean square error
(RMSE) between resolved MC results and the spectral solution is
0.064, whereas the RMSE for the first and second order FD method is
significantly larger at 0.280 and 0.180, respectively. To mitigate the dis-
sipation and inaccuracy, FD requires an excessive resolution for the
computation of the PDF after marginalization.
Remark 2: The spectral solution shows dispersion errors in the
form of high-frequency oscillations in the distribution function. These
are induced by the high-order approximation of the steep gradients in
the PDF that in turn are a result of the steep gradient in the uniform
forcing distribution f
a
. These dispersion errors, however, average out
and turn out to have no significant effect on the numerical accuracy of
the first three moments [see Fig. 13(b) for the third central moment].
The second order FD scheme also exhibits dispersion errors, but the
FD’s oscillations do not average out and the moments are not accu-
rately captured using this discretization.
Remark 3: For deterministic initial conditions, the regularization
of the Dirac delta is necessary to accurately compute the moments of
the evolving distribution function. Particularly, the vanishing moment
condition ensures that the evolution of the third moment [Fig. 13(b)]
is not affected by the numerical approximation as compared to Dirac
delta regularization with only two vanishing moment m¼2in(24),or
a Gaussian distribution function.
C. Stagnation flow
In the particle-laden stagnation flow, the relative (interphase)
velocity is not only affected by the evolution of the particle phase as is
FIG. 11. Marginals fxpand fupat t¼½0;0:54;1:6in (a), (b), and (c) respectively,
for the UF test case with dIC and the uniform distribution for f
a
. Contour plots of the
joint PDF fxpupsuperposed with the mean of the particle phase solution.
FIG. 12. Schematic evolution in time of the particle phase PDF fxpupfor the UF test
case.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-11
Published under license by AIP Publishing
the case in the uniform flow but also by the evolution of the carrier
phase velocity along the particle’s path. The temporal development of
the random particle position and velocity, therefore, displays a consid-
erably more complex behavior as compared to the uniform flow.
Because of the spatial dependence of the carrier flow, the particle solu-
tion is furthermore non-trivially dependent on its initial condition.
We consider two initial conditions described in Sec. I, one with the
particles starting at rest (Case SFR) and another withthe initial particle
velocity specified at the carrier flow’s velocity conditions (Case SFF).
We discuss the MC and MoD solutions for each case below.
1. SFR case: Monte Carlo results
The mean trends with two standard deviation bandwidth deter-
mined with the MC approach for a uniform forcing distribution are plot-
ted in Fig. 14. To discuss the SFR case, three stages of development are
identified. In the first stage (t<0.6) each particle identified with a coun-
ter iaccelerates in positive x-direction at a rate aiðkxpikupiÞ=sp.
Similar to the uniform flow case, the velocity and position variance both
increase in this stage with varying acceleration of the stochastically forced
particles. The second central moment plotted vs time in Fig. 15(a) con-
firms the growth of the particle variance in this stage.
In a second stage (0:6<t<1:6), the particle with the smallest
response time sp=amax—the fastest responding particle with accelera-
tion rate amax ðkxpmax kupmax Þ=sp—has accelerated to the carrier
velocity (t¼0.6). After that, the flow velocity continues to decrease
(stagnates) along this particle’s path. Because of the particle’s inertial
response, the particle’s velocity, however, does not decrease equally
fast along the particle’s path. Effectively, the relative velocity of this
particle therefor becomes negative. In other words, the particle starts
to decelerate. As more particles with larger response time reach the
carrier flow conditions, more particles start to decelerate until all par-
ticles have a negative relative velocity. During this second stage, the
velocity variance of the particle phase decreases to a minimum at t
1:6[Fig. 15(a)].
In a third stage t>1.6, when the relative particle velocity is smaller
than zero (up>uðxpÞ) for all the particles, the cloud decelerates to a
decreasing carrier velocity and the particle velocity variance increases.
The variance increase mechanism is similar to the first stage and the
uniform flow, in which a time varying carrier velocity in combination
with a random forcing leads to a variance increase in the particle veloc-
ity. In the stagnating flow, the random particle cloud compresses with a
decreasing position variance before the wall is reached.
As opposed to the uniform flow case the carrier flow’s velocity
fluctuation, u0, for the stagnation ow is non-zero which affects several
terms in the moment equations. Even with these extra terms, just like
for the uniform flow, the mean stagnation flow solution described by
(8) and (9) can also be approximated by the solution of the determinis-
tic equation for a¼
a. The latter position and velocity solution have a
root mean square deviation of 0.0073 for
xpand 0.0053 for
upas com-
pared to the former. Both the terms /0u0and /0u0
pturn out to be neg-
ligible in (9).
The evolution of the velocity variance as governed by its moment
equation (11) is affected by the second term on the right hand side,
FIG. 13. Comparison between the spectral discretization and finite difference
upwind discretization with first and second order for (a) fupat t¼0.24 and (b) sup.
Both figures are for the UF test case with dIC with a uniform forcing distribution.
FIG. 14. Two standard deviation interval along the mean for the test case SFR with
a uniform forcing distribution. In dashed green, the particle velocity computed
with the MoD and in black with MC. In dashed red, the particle position computed
with MoD and in blue with MC. Dark colors indicate dIC, whereas light ones indicate
sIC.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-12
Published under license by AIP Publishing
i.e., /0u0
pð
u
upÞ.Att1:6, the relative velocity ð
u
upÞin this
term changes sign when the particle phase begins to decelerate after its
initial acceleration. The sign of /0u0
pin this term changes at t1:6
also as follows: in the first stage u
p
is monotonically increasing for all
the forcing values of /¼aaccording to the analytical velocity solu-
tion (32); in stage two, some particles are accelerating and others are
decelerating which yields different signs for dup=d/depending on /.
Upon ensemble averaging, it turns out that the mean of dup=d/is
positive prior to t1:6 and negative after. In the third stage, u
p
is
monotonically decreasing with respect to /. So, the correlation term
/0u0
pchanges sign at t1:6 and thus the term /0u0
pð
u
upÞ>0.
In addition to the damping term that was discussed for the uni-
form flow case, the first term on the right hand side in the velocity var-
iance equations (11) also involves the term
/u0u0
pfor the stagnation
flow case. Similar to the sign change of /0u0
pat t1:6, the sign of
/u0u0
pis the same as
uupbecause of a comparable behavior of
dup=du and dup=d/. As a result,
/u0u0
pis negative before t1:6and
positive thereafter. The values of the term u0u0
pare between –0.002
and 0.002 and are thus of the same order as the velocity variance r2
up
[see Fig. 15(a)]. The term
/u0u0
ptherefor has a significant effect on the
variance dynamics. At early times, it reduces the growth of the vari-
ance, and at later times it enhances growth as compared to the uni-
form flow where the term is zero.
The third order correlation terms (the third and fourth term in
(11)) are observed to have a negligible contribution to the particle vari-
ance evolution. In comparing the maximum magnitude of each of the
terms in the right hand side of (11) with respect to the left hand side
over the time interval, we find that the terms /0u0u0
pand u0u02
phave at
most a 3.0% and 1.0% contribution, whereas the first and second
terms have a significant 120% and 152% contribution to the “variance
acceleration.”
Figures 15(a) and 15(b) include the variance evolution for several
distribution functions of the forcing f
a
. As in the uniform flow case,
the effect of the shape of f
a
is small on the order of 5% in the velocity
variance and slightly more (order of 10%) in the position variance.
The general trends are not affected by the shape of the forcing PDF.
The third position moment is negative throughout the time inter-
val considered [Fig. 15(b)], indicating a non-symmetric position distri-
bution that is skewed toward larger values of the particle coordinate.
To understand the evolution of the particle velocity’s third moment,
we differentiate between two stages; rst, when the mean interphase
velocity
u
upis positive and the skewness shows a bias toward
higher velocity values similar to the uniform flow case as also illus-
trated in Fig. 1. Second, when
u
up0atfirstafterwhichit
becomes negative, i.e.,
u
up<0 with a near zero skewness first and
decreasing after showing a bias toward small values of the particle
coordinate when
u
up<0 right before the particles hit the wall.
This second stage can be also understood through the evolution of the
PDF that consists of the formation of the singular Dirac delta distribu-
tion for which the skewness is zero and its consequent behavior as
illustrated in Fig. 1(b) with a change of the bias in the PDF.
Like in the uniform flow case, the evolution of the third central
moment is affected by many different terms in the velocity skewness
equation (13) as shown in Fig. 16. The fourth order correlation terms
are important in the stagnation flow also, but because the velocity fluc-
tuation is non-zero, u0 0, the evolution of terms that involve u0are
non-trivial and require a separate and more in-depth analysis. We feel
this is outside the scope of the current paper and we plan to report on
the skewness behavior in more detail in future work.
Stochastic initial conditions do not only alter the evolution of the
mean of the stagnation flow solution with dIC by a constant offset as
wasthecasefortheuniformflow(seeFig. 14), but the difference
between the solutions with dIC and sIC changes considerably over the
time interval and specifically at early times. The variance of the particle
position and velocity is initially offset according to its initial values as
shown in Fig. 15(a), but then the difference with respect to the dIC
casedecreasesastimeevolves.Thisreductioncanbeunderstoodby
considering the stagnation flow solution where particles can nonphysi-
cally cross the wall (i.e., an opposed jet flow). For this ow, all particles
move toward the same nal state with x
p
¼0andu
p
¼0intheasymp-
totic time limit, t!1, and thus the position and velocity variance
tend to zero.
Between the initial time and the infinite time, the terms x0
pu0
pand
u0u0
pare responsible for the reduced variances. The contribution of
u0u0
pwhich is negative for t<1.6, particularly, causes a greater increase
in the damping term for sIC as compared to dIC at early times. When
FIG. 15. Computations for the SFR case with MC in color lines and with the MoM
in black dots of the (a) second and (b) third central moments for the particle position
and velocity and the three distributions considered for a(see Fig. 6) and dIC. It is
included the case of sIC for the uniform distribution. The legend in (a) is valid for
(b) as well.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-13
Published under license by AIP Publishing
the interphase velocity changes sign at t1:6, this term becomes pos-
itive and it will have the opposite effect. A physical interpretation is as
follows: a more energetic initial state with higher velocity variance is
more resistant to changes induced by stochastic forcing resulting in
greater damping at early times. The term x0
pu0
pis positive in the accel-
eration stage and negative in the deceleration stage and its magnitude
is greater for dIC as compared to sIC consistent with greater values of
u0
pfor sIC.
Another considerable difference between the dIC and sIC is
that the minimum in the velocity variance at t1:6isnon-zero
for the stochastic case, while it is nearly zero for the deterministic
case. As a consequence, the singularity in the distribution function
when the relative velocity changes sign can be expected to be less
significant and the PDF can be expected to have a broader
support.
2. SFR case: Method of distributions
Using the same grid as was used for the uniform flow case, the
PDF solution for the SFR case with a uniform forcing distribution,
f
a
, and dIC are computed and plotted for three instances in Fig. 17
at t¼1.22, t¼1.60, and t¼2.15. At time t¼0, the initial condition
is identical to the uniform flow case plotted in Fig. 11(a), and it is
therefore not repeated in Fig. 17. The MC results are also plotted
in Fig. 17, and they are in excellent agreement with the MoD
results.
During the first stage (t<0.6), the joint PDF fxpupdeforms along
the mean of the particle trajectory (depicted by the black solid line),
showing a non-linear clustering of the particles in the XpUpplane
toward high values. During the second stage (0:6<t<1:6), some
particles accelerate and others decelerate leading to the near singular
Dirac delta distribution at t1:6[Fig. 17(b)]. At later times (t>1.6),
the PDF of the particle velocity increases on the left front [Fig. 17(c)],
confirming a bias toward lower velocities in a deceleration field as dis-
cussed in the MC results.
The position PDF solution has an increasing bias toward the
large value of X
p
which is consistent with the asymptotic infinite time
behavior of the nonphysical solution where particles are permitted to
cross the wall and where both the particle velocity and position distri-
bution evolve to a Dirac delta centered at X
p
¼0andU
p
¼0.
FIG. 16. Terms in Eq. (13) vs time tfor the SFR case with dIC and a uniform forc-
ing distribution.
FIG. 17. Marginals fxpand fupat t¼½1:22;1:60;2:15in (a), (b), and (c), respec-
tively, for the SFR test case with dIC and the uniform distribution for f
a
. Contour plots
are of the joint PDF fxpupsuperposed with the mean of the particle phase solution.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-14
Published under license by AIP Publishing
3. SFF case: Monte Carlo results
In a final test, the particle velocity is initialized with the carrier
phase velocity at the particle position. The MC results for the mean
with a two standard deviation bandwidth are plotted vs time in
Fig. 18. In the SFF case, the particle phaseonly decelerates which yields
an evolution that is opposite to the uniform flow evolution as plotted
in Fig. 1(a) or an evolution that is very similar to the third
“deceleration” stage of the SFR case for t>1.6. The mean velocity
decreases monotonically when the mean particle position increases
toward the wall. This evolution is accompanied by an increase in the
variances of both x
p
and u
p
.
Because the SFF case is similar to the other two cases, the
moment evolution results do not shed any additional light on the evo-
lution of the stochastically forced particle phase. It is therefore omitted
here. For completeness, we include the plots for evolution of the
moments in Appendix C (Figs. 20 and 21).
4. SFF case: Method of distributions
The results for the uniform distribution forcing, f
a
,fortheSFF
case are also very similar to the deceleration stage of the SFR case.
Rather than reiterating that discussion, we choose a different sto-
chastic forcing according to a beta distribution for f
a
in Fig. 6
which does not have steep gradients in f
a
like the uniform distribu-
tion. For a grid with the same size as described before, the distribu-
tion results for two different times are shown in Figs. 19(a) and
19(b) with a deterministic initial condition. Clearly, the solution
does not show Gibbs oscillations and the MC results and the MoD
are in excellent agreement.
As time evolves, the PDF of the particle position is advected with
a positive characteristic velocity (14) and the particle velocity with a
negative velocity according to (15).ThePDFswidenintimeasthe
response times of random particles is different for different stochastic
forcing leading to variations in the particles velocities and positions.
Both the position and velocity PDF display a non-Gaussian (non-sym-
metric) trend that is more subtle than for the uniform forcing.
VII. CONCLUDING REMARKS
Several techniques and models including a Monte Carlo
approach, a method of moments, and a method of distributions are
developed and compared for analysis of particle dynamics with sto-
chastic forcing in one-way coupled Eulerian–Lagrangian formula-
tions. Random solutions of two canonical flow problems are
discussed including a particle phase accelerated in a uniform car-
rier flow and a particle phase released in a stagnation carrier flow
with two initial conditions, one at rest and one initialized at the
carrier flow velocity.
Starting from the Lagrangian particle equations for position and
velocity with stochastic forcing, a closed PDF formulation is derived.
A single hyperbolic partial differential equation, whose characteristic
advection velocities are non-constant, governs the evolution of the
PDF solution. In a single spatial dimension, the PDF depends on three
FIG. 18. Two standard deviation interval along the mean for the test case SFF with
a uniform forcing distribution. In dashed green, the particle velocity computed with
the MoD and in black with MC. In dashed red, the particle position computed with
MoD and in blue with MC. Dark colors indicate dIC, whereas light ones indicate
sIC.
FIG. 19. Marginals fxpand fupat t¼½0:87;1:17in (a) and (b), respectively, for the
SFR test case with dIC and a uniform forcing distribution. Contours of the joint PDF
fxpupsuperposed with the mean of the particle phase solution.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-15
Published under license by AIP Publishing
variables at a given time, including the position, the velocity, and a
forcing coefficient.
A high-order spectral method with discontinuity regulariza-
tion is necessary for the accurate solution of the hyperbolic partial
differential equation that admits discontinuities. A polynomial reg-
ularization of a Dirac delta function with mvanishing moments is
shown to accurately capture the first mmoments of the PDF solu-
tion in time.
Moment equations are derived for the first three moments of the
particle position and velocity, representing the mean, variance, and
skewness of the PDF. Monte Carlo results are used to determine corre-
lations terms and to close the system of moment equations.
Analytical solutions are derived for the system of two linear
ODEs that govern the dynamics of particles with a deterministic forc-
ing in a one-dimensional uniform flow and stagnation flow. The parti-
cle solution in the stagnation flow has its nal state with a zero velocity
at the wall. Depending on the relative forcing, the particle manifolds in
the phase space (position/velocity space) tend to either a node or
spiral.
Themeansolutionwithrandomforcingcanbeapproximated
within 1% using a mere single deterministic solution at the mean forc-
ing for all flow cases considered.
In flows where all randomly forced particles accelerate or deceler-
ate, the velocity variance increases driven by a single correlation source
term. A damping terms counters this source term. When the particle
velocity settles, the velocity variance reduces to zero because of this
damping. Higher-order correlation terms are generally negligible in
the velocity variance equation. The position variance increases in
accelerating flows and decreases in decelerating flows, i.e., the random
cloud expands and compresses, respectively. When the relative velocity
changes sign, the particle variance approaches zero and the PDF has a
very narrow support.
The skewness of the distribution function has a bias toward the
carrier velocity to which the particle accelerates or decelerates. The
bias of the distribution function is non-linear and more significant
toward to the tail ends of the distribution function. The skewness
equation is driven by a sourcing and a damping similar to the variance
equation, but with different response times. High-order correlation
terms are significant, suggesting a complicated tail behavior of the
PDF.
In near future work, we intend to report on the tail behavior of
the PDF solution of the particle phase. We also plan to develop distri-
bution function models for two-way couple Eulerian–Lagrangian
formulations.
ACKNOWLEDGMENTS
This work was supported in part by the Air Force Office of
Scientific Research under Award Nos. FA9550-19-1-0387 and
FA9550-18-1-0474.
APPENDIX A: DERIVATION OF THE PDF EQUATION
USING THE METHOD OF DISTRIBUTIONS
Here, a PDF equation is developed to solve the stochastic sys-
tem defined by (1) with the so-called method of distributions. First,
we define the fine-grained JPDF Pas
PðA;a;Xp;xp;Up;up;tÞ¼dðAaÞdðXpxpðtÞÞdUpupðtÞ

;
(A1)
where A,X
p
, and U
p
are deterministic magnitudes and dðÞ is the
Dirac delta function. Taking the derivative of Pwith respect to
time and using the chain rule and the Dirac delta properties, we
find
@P
@t¼dxp
dt
@P
@xpþdup
dt
@P
@up¼dxp
dt
@P
@Xpdup
dt
@P
@Up
;(A2)
where we can make use of (1) for writing
@P
@tþup
@P
@Xpþag u up
ðÞ
spðuupÞ@P
@Up¼0;(A3)
that in conservative form is
@P
@tþ@
@Xp
upP

þ@
@Up
agðuupÞ
spðuupÞP
"#
¼0:(A4)
Defining the ensemble mean of an integrable function
hða;xp;upÞwith the joint PDF faxpupas
Ehða;xp;upÞ

¼ððð1
1
hðA0;X0
p;UpÞfaxpupðA0;X0
p;U0
p;tÞdA0dX0
pdU0
p;
(A5)
the ensemble of the function Pfor a particular set of the determin-
istic variables is obtained as
EP
½
¼ððð1
1
PðA;A0;Xp;X0
p;Up;U0
p;tÞ
faxpupðA0;X0
p;U0
p;tÞdA0dX0
pdU0
p
¼ððð1
1
dðAA0ÞdðXpX0
pÞdðUpU0
pÞ
faxpupðA0;X0
p;U0
p;tÞdA0dX0
pdU0
p
¼faxpupðA;Xp;Up;tÞ:(A6)
This procedure suggests how to obtain a partial differential equation
for faxpupfrom (A4) taking the ensemble mean each term. To do so,
we need to use the property that allows us to exchange expectation
with derivatives respect to deterministic variables. For example, for
the deterministic variable time, one has
E@hða;xp;upÞ
@t

¼@Ehða;xp;upÞ

@t:(A7)
Using this property, the ensemble mean of the first term in (A4) is
trivial
E@P
@t

¼@faxpup
@t:(A8)
The second is calculated as
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-16
Published under license by AIP Publishing
E@upP

@Xp
"#
¼@
@Xp
Eu
pP
 (A9)
¼@
@Xpððð1
1
U0
pPðA;A0;Xp;X0
p;Up;U0
p;tÞ
faxpupðA0;X0
p;U0
p;tÞdA0dX0
pdU0
pi(A10)
¼@
@Xpððð1
1
U0
pdðAA0ÞdðXpX0
pÞdðUpU0
pÞ
faxpupðA0;X0
p;U0
p;tÞdA0dX0
pdUpi(A11)
¼@
@Xp
Upfaxpup

:(A12)
In the same way, the third term is
E@
@Up
ag u up
ðÞ
spðuupÞP
!"#
¼@
@Up
Ag U Up
ðÞ
spðUUpÞfaxpup
!
;(A13)
where for coherence we make use of Uas the deterministic value of
the carrier flow. In this study, the carrier flow is considered deter-
ministic and therefore u¼U.
Finally, the deterministic equation that governs the joint prob-
ability density function of the solution is
@faxpup
@tþ@
@Xp
Upfaxpup

þ@
@Up
AgðUUpÞ
spðUUpÞfaxpup
!
¼0:
(A14)
This equation has to be solved with deterministic or stochastic ini-
tial conditions defined by (7) or (6) and also be marginalized
according to
fxpupðXp;Up;tÞ¼ð1
1
faxpupðA;Xp;Up;tÞdA;(A15)
fxpðXp;tÞ¼ð1
1
fxpupðXp;Up;tÞdUp;(A16)
fupðUp;tÞ¼ð1
1
fxpupðXp;Up;tÞdXp:(A17)
We also define here the n
th
moment about cof a continuum ran-
dom variable xwith PDF fxðxÞas
ln¼ð1
1
xc
ðÞ
nfxðxÞdx;(A18)
where if n¼1 and c¼0, we obtain the mean l
x
, and if cis selected
to be the mean, we find the n
th
central moment of f
x
.
APPENDIX B: DETERMINISTIC ANALYTICAL
SOLUTION OF THE TEST CASES
1. Uniform flow
For the uniform flow where uis constant, the system of Eq. (1)
can be expressed as a first order ODE system with constant coeffi-
cients for xpðtÞand upðtÞ
d
dt
xp
up

¼01
0b

xp
up

þ0
bu

;(B1)
a second order ODE for xpðtÞ
d2xp
dt2þbdxp
dt ¼bu;(B2)
a first order ODE for upðtÞ
dup
dt þbup¼bu;(B3)
or as a differential equation of separable variables for the phase
space upðxpÞas the system (1) becomes autonomous
dup
dxp¼buup
ðÞ
up
:(B4)
The analytical solution of xpðtÞand upðtÞis trivially obtained solv-
ing any of the above options as
xpðtÞ¼xp0þut þ1
bðuup0Þebt 1
ðÞ
;(B5)
upðtÞ¼uþðup0uÞebt:(B6)
The time can be removed combining the last two equations to find
the solution in the particle phase as upðxpÞ.
2. Stagnation flow
For the centerline of the stagnation flow y¼0, the carrier flow
is u¼kx ¼kxpafter interpolating at the particle location
according to Hiemenz solution.
50
The system (1) can be described
as a first ODE system of constant coefficients
d
dt
xp
up

¼01
kb b

xp
up

;(B7)
a second order ODE for xpðtÞ
d2xp
dt2þbdxp
dt þkbxp¼0;(B8)
or an integrable equation for the particle phase upðxpÞ
dup
dxp¼bkx
pþup

up
:(B9)
Writing the system (B7) as z0ðtÞ¼BzðtÞ, with the initial condition
zð0Þ¼z0, where
z0ðtÞ¼ xpðtÞ
upðtÞ

;z0¼xp0
up0

;B¼01
bk b

;(B10)
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-17
Published under license by AIP Publishing
the analytical solution is given by
zðtÞ¼eBt z0:(B11)
Using the eigen decomposition B¼SKS1, the exponential matrix
can be obtained as
eBt ¼SeKtS1;(B12)
with
S¼1þffiffiffiffiffiffiffiffiffiffiffiffiffi
b4k
p
2kffiffi
b
p1ffiffiffiffiffiffiffiffiffiffiffiffiffi
b4k
p
2kffiffi
b
p
11
2
43
5;K¼k10
0k2

;
(B13)
and
k1¼bffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
bb4k
ðÞ
p2;k2¼bþffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
bb4k
ðÞ
p2:(B14)
Finally, the exponential matrix is
eKt¼ek1t0
0ek2t

;(B15)
and the analytical solution can be expressed as
xpðtÞ¼ebt
2bxp0þ2up0
csinh tc
2

þxp0cosh tc
2

;(B16)
and
upðtÞ¼1
2cetcþb
ðÞ
2up0cectþ1
ðÞ
be
ct1
ðÞ
ð2kxp0þup0Þ
hi
;
(B17)
where c¼ffiffi
b
pffiffiffiffiffiffiffiffiffiffiffiffiffi
b4k
p.
Just for completion, in the case b¼4k, the matrix Sis singular,
and the Jordan decomposition B¼MJM1is required
52
to find the
exponential matrix defined as eBt ¼MeJt M1with
M¼1=ð2kÞ1=ð4k2Þ
10

;J¼2k1
02k

;(B18)
so that one has
eJt ¼e2kt e2ktt
0e2kt

;(B19)
and the analytical solution changes to
xpðtÞ¼e2kt xp0þt2kxp0þup0

;(B20)
upðtÞ¼e2kt up02kt 2kxp0þup0

:(B21)
APPENDIX C: MOMENT RESULTS FOR THE SFF
CASE
FIG. 20. Computations for the SFF case with MC in color lines and with the MoM in
black dots of the (a) second and (b) third central moments for the particle position
and velocity and the three distributions considered for a(see Fig. 6) and dIC. The
case of sIC for the uniform distribution is included. The legend in (b) is valid for (a)
as well.
FIG. 21. Terms in (13) vs time tfor the SFF with dIC.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-18
Published under license by AIP Publishing
DATA AVAILABILITY
The data that support the findings of this study are available
from the corresponding author upon reasonable request.
REFERENCES
1
C. T. Crowe, M. P. Sharma, and D. E. Stock, “The particle-source-in cell
(PSI-CELL) model for gas-droplet flows,” J. Fluids Eng. 99(2), 325–332
(1977).
2
O. Sen, S. Davis, G. B. Jacobs, and H. S. Udaykumar, “Evaluation of con-
vergence behavior of metamodeling techniques for bridging scales in
multi-scale multimaterial simulation,” J. Comput. Phys. 294, 585–604
(2015).
3
O. Sen, N. J. Gaul, K. K. Choi, G. B. Jacobs, and H. S. Udaykumar, “Evaluation
of kriging surrogate models constructed from mesoscale computations of
shock interactions with particles,” J. Comput. Phys. 336, 235–260 (2017).
4
S. Davis, O. Sen, G. B. Jacobs, and, and H. S. Udaykumar, “Coupling of micro-
scale and macro-scale Eulerian-Lagrangian models for the computation of
shocked particle-laden flows,” In ASME International Mechanical Engineering
Congress and Exposition (American Society of Mechanical Engineers, 2013),
Vol. 56314, p. V07AT08A011.
5
S. Lomholt and M. R. Maxey, “Force-coupling method for particulate two-
phase flow: Stokes flow,” J. Comput. Phys. 184, 381–405 (2003).
6
G. S. Shallcross, R. O. Fox, and J. Capecelatro, “A volume-filtered description
of compressible particle-laden flows,” Int. J. Multiphase Flow 122, 103138
(2020).
7
C. T. Crowe, M. Sommerfeld, and Y. Tsuji, Multiphase Flows with Droplets and
Particles (CRC Press LLC, Boca Raton, FL, 1998).
8
J. D. Anderson, Modern Compressible Flow: with Historical Perspective,
Number Book (McGraw-Hill, 1990).
9
G. G. Stokes, On the Effect of the Internal Friction of Fluids on the Motion of
Pendulums (Pitt Press Cambridge, 1851), Vol. 9.
10
M. R. Maxey and J. J. Riley, “Equation of motion for a small rigid sphere in a
nonuniform flow,” Phys. Fluids 26, 883–889 (1983).
11
V. M. Boiko and S. V. Poplavskii, “Drag of nonspherical particles in a flow
behind a shock wave,” Combust. Explos. Shock Waves 41(1), 71–77 (2005).
12
E. Loth, “Compressibility and rarefaction effects on drag of a spherical particle,”
AIAA J. 46(9), 2219–2228 (2008).
13
G. Tedeschi, H. Gouin, and M. Elena, “Motion of tracer particles in supersonic
flows,” Exp. Fluids 26(4), 288–296 (1999).
14
Z. G. Feng, E. E. Michaelides, and S. Mao, “On the drag force of a viscous
sphere with interfacial slip at small but finite Reynolds numbers,” Fluid Dyn.
Res. 44(2), 025502 (2012).
15
W. E. Ranz and W. R. Marshall, Jr., “Evaporation from drops, Part II,” Chem.
Eng. Prog. 48, 173–180 (1952).
16
O. Sen, N. J. Gaul, K. K. Choi, G. B. Jacobs, and H. S. Udaykumar, “Evaluation
of multifidelity surrogate modeling techniques to construct closure laws for
drag in shock–particle interactions,” J. Comput. Phys. 371, 434–451 (2018).
17
G. B. Jacobs and H. S. Udaykumar, “Uncertainty quantification in Eulerian-
Lagrangian simulations of (point-) particle-laden flows with data-driven and
empirical forcing models,” Int. J. Multiphase Flow 121, 103114 (2019).
18
W. C. Moore, S. Balachandar, and G. Akiki, “A hybrid point–particle force
model that combines physical and data–driven approaches,” J. Comput. Phys.
385, 187–208 (2019).
19
A. Seyed-Ahmadi and A. Wachs, “Microstructure–informed probability–driven
point–particle model for hydrodynamic forces and torques in particle–laden
flows,” J. Fluid Mech. 900, A21 (2020).
20
S. Taverniers and D. M. Tartakovsky, “Estimation of distributions via multi-
level Monte Carlo with stratified sampling,” J. Comput. Phys. 419, 109572
(2020).
21
S.Taverniers,S.B.M.Bosma,andD.M.Tartakovsky,“AcceleratedmultilevelMonte
Carlo with kernel-based smoothing and Latinized stratification,” Water Resour. Res.
56(9), e2019WR026984, https://doi.org/10.1029/2019WR026984 (2020).
22
X. Ma and N. Zabaras, “An adaptive hierarchical sparse grid collocation algo-
rithm for the solution of stochastic differential equations,” J. Comput. Phys.
228(8), 3084–3113 (2009).
23
D. A. Barajas-Solano and D. M. Tartakovsky, “Stochastic collocation methods
for nonlinear parabolic equations with random coefficients,” SIAM/ASA J.
Uncertainty Quantif. 4(1), 475–494 (2016).
24
X. Wan and G. E. Karniadakis, “An adaptive multi-element generalized poly-
nomial chaos method for stochastic differential equations,” J. Comput. Phys.
209(2), 617–642 (2005).
25
X. Wan and G. E. Karniadakis, “Multi-element generalized polynomial chaos
for arbitrary probability measures,” SIAM J. Sci. Comput. 28(3), 901–928
(2006).
26
D. Venturi, X. Wan, and G. E. Karniadakis, “Stochastic bifurcation analysis of
Rayleigh-B
enard convection,” J. fluid mechanics 650, 391–413 (2010).
27
B. Debusschere, Intrusive Polynomial Chaos Methods for Forward Uncertainty
Propagation (Springer International Publishing, Cham, 2017), pp. 617–636.
28
M. Ye, S. P. Neuman, A. Guadagnini, and D. M. Tartakovsky, “Nonlocal and
localized analyses of conditional mean transient flow in bounded, randomly
heterogeneous porous media,” Water Resour. Res. 40, W05104, https://
doi.org//10.1029/2003WR002099 (2004).
29
T. Maltba, P. Gremaud, and D. M. Tartakovsky, “Nonlocal PDF methods for
langevin equations with colored noise,” J. Comput. Phys. 367, 87–101 (2018).
30
T. S. Lundgren, “Distribution functions in the statistical theory of turbulence,”
Phys. Fluids 10(5), 969–975 (1967).
31
D. M. Tartakovsky and P. A. Gremaud, “Method of distributions for uncer-
tainty quantification,” in Handbook of Uncertainty Quantification, edited by R.
Ghanem, D. Higdon, and H. Owhadi (Springer, New York, 2015), pp.
763–783.
32
R. J. Rutjens, G. B. Jacobs, and D. M. Tartakovsky, “Method of distributions for
systems with stochastic forcing,” arXiv:1909.01774 (2019).
33
P. Wang and D. M. Tartakovsky, “Uncertainty quantification in kinematic-
wave models,” J. Comput. Phys. 231(23), 7868–7880 (2012).
34
P. Wang, D. M. Tartakovsky, Jr. K. D. Jarman, and A. M. Tartakovsky, “CDF
solutions of Buckley-Leverett equation with uncertain parameters,” Multiscale
Model. Simul 11(1), 118–133 (2013).
35
A. Alawadhi, F. Boso, and D. M. Tartakovsky, “Method of distributions for
water-hammer equations with uncertain parameters,” Water Resour. Res.
54(11), 9398–9411, https://doi.org/10.1029/2018WR023383 (2018).
36
F. Boso and D. M. Tartakovsky, “Data-informed method of distributions for
hyperbolic conservation laws,” SIAM J. Sci. Comput. 42(1), A559–A583
(2020).
37
H.-J. Yang, F. Boso, H. A. Tchelepi, and D. M. Tartakovsky, “Probabilistic fore-
cast of single-phase flow in porous media with uncertain properties,” Water
Resour. Res. 55(11), 8631–8645, https://doi.org/10.1029/2019WR026090
(2019).
38
G. Haller, “Lagrangian coherent structures,” Annu. Rev. Fluid Mech. 47,
137–162 (2015).
39
S. Jean-Piero and G. B. Jacobs, “Regularization of singularities in the weighted
summation of Dirac-delta functions for the spectral solution of hyperbolic con-
servation laws,” J. Sci. Comput. 72(3), 1080–1092 (2017).
40
B. W. Wissink, G. B. Jacobs, J. K. Ryan, W. S. Don, and E. T. A. van der
Weide, “Shock regularization with smoothness-increasing accuracy-conserving
Dirac-delta polynomial kernels,” J. Sci. Comput. 77(1), 579–596 (2018).
41
G. B. Jacobs and W. S. Don, “A high-order WENO-Z finite difference based
particle-source-in-cell method for computation of particle-laden flows with
shocks,” J. Comput. Phys. 228(5), 1365–1379 (2009).
42
S. L. Davis, G. B. Jacobs, O. Sen, and H. S. Udaykumar, “SPARSE-A subgrid
particle averaged Reynolds stress equivalent model: Testing with a priori
closure,” Proc. R. Soc. A: Math., Phys. Eng. Sci. 473(2199), 20160769 (2017).
43
B. Shotorban, G. B. Jacobs, O. Ortiz, and Q. Truong, “An Eulerian model for
particles nonisothermally carried by a compressible fluid,” Int. J. Heat Mass
Transfer 65, 845–854 (2013).
44
R. G. Patel, O. Desjardins, and R. O. Fox, “Three-dimensional conditional
hyperbolic quadrature method of moments,” J. Comput. Phys.: X 1, 100006
(2019).
45
S. Jean-Piero, G. B. Jacobs, and W. S. Don, “A high-order Dirac-delta regulari-
zation with optimal scaling in the spectral solution of one-dimensional singular
hyperbolic conservation laws,” SIAM J. Sci. Comput. 36(4), A1831–A1849
(2014).
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-19
Published under license by AIP Publishing
46
K. Sengupta, B. Shotorban, G. B. Jacobs, and F. Mashayek, “Spectral-based sim-
ulations of particle-laden turbulent flows,” Int. J. Multiphase Flow 35(9),
811–826 (2009).
47
D. Gottlieb and J. S. Hesthaven, “Spectral methods for hyperbolic problems,”
J. Comput. Appl. Math. 128(1-2), 83–131 (2001).
48
J. S. Hesthaven, S. Gottlieb, and D. Gottlieb, Spectral Methods for Time-
Dependent Problems (Cambridge University Press, 2007), Vol. 21.
49
S. Gottlieb and C. W. Shu, “Total variation diminishing Runge-Kutta schemes,”
Math. Comput. 67(221), 73–85 (1998).
50
K. Hiemenz, “Die Grenzschicht an einem in den gleichformigen
Flussigkeitsstrom eingetauchten geraden Kreiszylinder,” Dinglers Polytech. J.
326, 321–324 (1911).
51
L. Rosenhead, Laminar Boundary Layers: An account of the Development,
Structure, and Stability of Laminar Boundary Layers in Incompressible Fluids,
Together with a Description of the Associated Experimental Techniques (Dover
Publications, 1988).
52
B. Noble and J. W. Daniel, Applied Linear Algebra (Prentice-Hall Englewood
Cliffs, NJ, 1977), Vol. 3.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 033326 (2021); doi: 10.1063/5.0039787 33, 033326-20
Published under license by AIP Publishing
... In the range of high subcritical Re, Winslow et al. [2] noted that the laminar flow over the suction side of deflectors is subjected to a significant adverse pressure gradient close to the leading edge, resulting in the separation as the shear layer with low kinetic energies. Later, the separated shear layer would gain enough momentum from the freestream to reattach to the airfoil surface as a turbulent boundary layer, and the reattachment point is quite close to the trailing edge [8]. That hints at a more considerable separation bubble formed at the upper surface of the deflector, whereas the increase of Re can decrease the size of bubbles as a quicker reattachment is anticipated. ...
Article
Full-text available
The cambered deflectors in aquacultural facilities are applied to enhance hydro-dynamic efficiencies or enable flow fields to be fully developed. Given the anticipated improvements with the bio-inspired profiles or tandem configurations, the hydrodynamics of cambered deflectors with the above features are investigated at Re = 1 × 10 5. The relationship between force coefficients and local flow behaviors for both bionic and non-bionic isolated deflectors, as well as tandem deflectors, is revealed using k − ω SST simulation. The dependencies of force coefficients on gap (G), stagger (S), and inclination angles (θ) in tandem deflectors are illustrated using an updated metamodeling workflow with simulated data. It is demonstrated that the variations of force coefficients over angles of attack are related to flow physics in boundary-layer regions. The non-bionic isolated deflector with the θ = 10 • prevails as the decent performances of C L and γ globally, which is chosen in the following studies. Regarding tandem deflectors, θ plays a more vital role in drag coefficients (C D) and lift coefficients (C L), while the influence of S is not quite considerable compared to G. Aiming for cost minimizations and lift improvements, an optimized tandem case is obtained and justified with the superiorities in flow fields. This study has provided novel insights into the designs and optimizations of cambered deflectors in aquacultural engineering.
... These findings align with prior studies on cambered airfoils, where moderate camber and thickness contribute to low Cd at small AoA, but rapid drag escalation occurs beyond the stall angle [15]. Similar behavior has been reported for NACA 4412, which maintains a low drag coefficient up to +12° before experiencing a sudden increase [16]. ...
Article
Full-text available
Wind energy is a promising renewable source, with its efficiency largely dependent on the aerodynamic performance of wind turbine blades. The selection of an appropriate airfoil significantly influences key aerodynamic factors such as lift-to-drag ratio, power coefficient, and thrust coefficient, which determine the turbine's overall energy output. This study examines the aerodynamic behavior of the NACA 5518 airfoil using QBlade software to assess its feasibility for wind turbine applications. Computational simulations are conducted based on Blade Element Momentum (BEM) theory and the XFOIL solver, evaluating lift coefficient (Cl), drag coefficient (Cd), power coefficient (Cp), and thrust coefficient (Ct). The designed rotor operates at a tip speed ratio (TSR) of 8.0, with a wind speed of 10.0 m/s at hub height. The blade incorporates nonlinear twist distribution and varying chord length to enhance aerodynamic efficiency. Results indicate that the NACA 5518 airfoil demonstrates strong aerodynamic properties, including high lift generation and delayed stall. The power coefficient (Cp) peaks at 0.63, while the thrust coefficient (Ct) reaches 1.14. Simulations align well with theoretical predictions, affirming the reliability of the computational model. However, increased drag at higher angles of attack suggests a need for further blade design optimizations. These findings underscore the potential of the NACA 5518 airfoil in wind turbine applications, offering high efficiency and improved power output. However, future research should explore flow control strategies and experimental validation for real-world implementation.
... 4.1.1. Integrated forces on the foil Figure 3 compiles data from 3-D and 2-D DNS together with measurements from two different wind tunnel experiments (Choi 2020;Tank et al. 2021). Also included are calculations from the panel code Xfoil (Drela 1989) which uses a boundary integral method to estimate separation locations. ...
Article
Full-text available
The flow over a cambered NACA 65(1)–412 airfoil at Re=2×104Re\,=\,2\times 10^4 is described based on a high-order direct numerical simulation. Simulations are run over a range of angles of attack, α\alpha , where a number of instabilities in the unsteady, three-dimensional flow field are identified. The balance and competing effects of these instabilities are responsible for significant and abrupt (with respect to α\alpha ) changes in flow regime, with measurable consequences in time-averaged, integrated force coefficients, and in the far-wake footprint. At low α\alpha , the flow is strongly influenced by vortex roll-up from the pressure side at the trailing edge. The interaction of this large-scale structure with shear and three-dimensional modal instabilities in the separated shear layer and associated wake region on the suction side, explains the transitions and bifurcations of the the flow states as α\alpha increases. The transition from a separation at low α\alpha to reattachment and establishment of a laminar separation bubble at the trailing edge at critical α\alpha is driven by instabilities within the separated shear layer that are absent at lower angles. Instabilities of different wavelengths are then shown to pave the path to turbulence in the near wake.
... where is the fluid density, is the flow velocity, is the channel diameter, and is the dynamic viscosity [45]. Generally, flow is considered laminar for Re < 2000. ...
Article
Full-text available
This paper introduces a novel MEMS-based resonant airflow sensor designed specifically for respiration monitoring in application such as obstructive sleep apnea (OSA). The sensor operates based on the Time-of-Flight (ToF) technique and employs thin-film piezoelectric-on-substrate (TPoS) resonators. The resonators operate at ~25 MHz and are integrated into two oscillator circuits the frequency of which varies with temperature and humidity. By measuring the flight time of the airflow between sensing elements, the sensor can precisely calculate flow rates and velocities. A carefully designed laminar airflow channel is incorporated to maintain non-turbulent flow, which is essential for the accuracy of ToF measurements. Experimental validation demonstrates the sensor’s ability to measure flow rates from 0 to 10 L/min and velocities from 0 to 2 m/s with high precision and linearity with an accuracy as high as 97.85%. These results represent a significant advancement in respiratory monitoring technology, offering a non-invasive and cost-effective solution for home-based detection and management of OSA.
... A counter-rotating Table 1 Quantitative evaluation of the aerodynamic performance of the NACA 0012 wing at Re 10 4 and α 4 deg vortex S 2 , as the secondary structure originating from the trailing edge, was found to pair with the dominant clockwise structure S 1 in the recirculation zones. Similar structures were reported by Tank et al. [49] for the time-averaged flowfield of a cambered airfoil below the critical angle of attack. For the front wing, the trailing-edge separation bubble started approximately from the streamwise location of x∕c 0:4 to x∕c 1:1, where the separation point coincided with that of an isolated airfoil at the same attack angle and Reynolds number [42]. ...
Article
Full-text available
A three-dimensional direct numerical simulation was conducted to investigate the vortex–wing interactions through two NACA 0012 stationary wings placed in tandem at a low Reynolds number [Formula: see text]. The aerodynamic characteristics and three-dimensional flow structures were analyzed for the tandem wings. The back wing disturbed by the upstream vortices gained an evident increase in aerodynamic performance, where the advantage is related to the suppression of the large-scale vortex formation near the trailing edge. The Liutex method was applied to visualize the vortical structures for investigating the three-dimensional evolution and instability when interacting with the back wing. The upstream wake triggered dual-secondary vortices and intensified the secondary instability on the back wing. The induced vortices contributed to the lift enhancement because they provided an extra low-pressure region when propagating downstream on the suction side of the back wing. Because of the three-dimensional destabilization, the vortex interaction in the evolution process accelerated the transition and injection of the high-momentum flow into the boundary layer attached to the back wing, energizing the turbulent boundary layer and eliminating the large-scale separation near the trailing edge. This study provided a new perspective on the enhanced aerodynamic performance of tandem layout.
... These phenomena were also 280 observed in studies with similar geometries, e.g.,Souppez et al. (2022) studied flow around an arc and found that 281 at 14°≤ ≤ 20° sharply increases when varies from 5.4 × 10 5 to 6.8 × 10 5 . 282The significant changes in the aerodynamic loads due to changes in are called lift crisis and drag crisis, i.e., 283 force crisis,(Schmitz, 1970;Tank et al., 2021), which is a common phenomenon for blunt bodies like cylinders 284(Holloway et al., 2004). The major reason for the force crisis is the laminar-turbulent transition(Hoffman & 285 Johnson, 2007). ...
Preprint
Full-text available
Wingsails as a classical means for wind-assisted ship propulsion is booming today. Crescent-shaped wingsails have shown great potential considering their aerodynamics in thrust generation. However, there have been few studies on its structural responses exposed to unsteady aerodynamic loads. In this study, a crescent-shaped wingsail at a 1:100 scale is investigated by using wind tunnel tests. The wingsail with aspect ratios of 0.19 and 0.33, which are derived from practical wingsail design, are measured at the angle of attack (α\alpha) from 00^\circ to 9090^\circ, and at the wind speed from 20m/s20 m/s to 40m/s40 m/s. The wind loads, pressure distribution, and wingsail tip displacements are measured and discussed. A study of the Reynolds number sensitivity indicates that the trends of the force coefficients and critical α\alpha are changed when the Reynolds number (Re) is above 3.1×1053.1\times {10}^{5}. Besides, at Re=3.1×105Re=3.1\times {10}^{5} a lift crisis is found for α\alpha below 1010^\circ, and a drag crisis happens for α=0\alpha =0^\circ at Re=3.6×105Re=3.6\times {10}^{5}. Structural response analysis is conducted based on the wingsail tip displacement in the direction perpendicular to the chord line. The fundamental eigenfrequency of the structure, instead of the flow-induced frequencies, is found to dominate the fluid-structure interaction measured in the tests. Furthermore, the aerodynamic performance of a three-wingsail system is analyzed based on experiments. Notable interactions are found among the three wingsails.
... Transition models are employed to improve the precision of forecasting the initiation and flow separation in aerodynamic simulations. 26,27 By considering the effects of laminar-turbulent transition, these models provide more accurate predictions of the flow behavior near stall conditions, including the separation point location, flow reattachment, and the development of vortical structures. The research investigates the effectiveness of different control strategies to delay or mitigate stall using transition models. ...
Article
Full-text available
The phenomenon of rotating stall in centrifugal pumps is closely associated with the evolution of the blade boundary layer. Aiming to accurately predict the characteristics of the boundary layer, this study investigates the phenomenon of rotating stall in centrifugal pump impellers using the gamma (γ) transition model. The accuracy of the numerical simulation was confirmed by comparing its conclusions with the results of the testing. In calculations considering transition characteristics, the distribution of low-pressure areas inside the impeller is relatively discontinuous, while the pressure distribution is more uniform. However, in calculations without considering transition, the low-pressure regions in neighboring flow channels exhibit a tendency to be interconnected, resulting in a more variable pressure distribution, and the pressure contour at the outlet is closer to parallel. The dynamic characteristics of the centrifugal pump impeller rotating stall were obtained through the dynamic mode decomposition method, including the frequency, structure, and dynamic evolution process of the stall vortex. Through modal reconstruction, it was discovered that the impeller's rotation causes the stall vortex to undergo periodic fluctuations. The stall vortex is not stationary but moves synchronously with the rotation of the blades. At different time points, the stall vortex exhibits periodic changes. At the blade suction entrance, the stall vortex initially appears. Subsequently, multiple vortex structures resulted in channel blockage. After a period of development, the excess vortex structures merge to generate a typical “8” shaped vortex structure and move toward the exit. Finally, the exit stall vortex disappears, and a new vortex structure is generated at the inlet of the blade suction surface.
Preprint
Full-text available
Experiments were conducted in a low-turbulence wind tunnel to investigate the efficacy of localised acoustic forcing upon the dynamics and stability of the flow on a cambered, wall-bounded airfoil over a range of Reynolds numbers (Re) where the flow state can switch between two limits -- a low-lift state (SI) where separation continues beyond the trailing edge and a high-lift state (SII) where the separated flow is closed off to form a laminar separation bubble. The switching between SI and SII can occur close to a critical angle of attack (αcrit\alpha_{\textrm{crit}}) which varies with Re\textrm{Re}. The most effective forcing frequencies are found at a constant value of a rescaled Strouhal number, St=St/Re1/2=0.027\textrm{St}^* = \textrm{St}/\textrm{Re}^{1/2}= 0.027, which indicates that though the primary unstable modes of the separated shear layer are of the inviscid, Kelvin-Helmholtz type, these modes are seeded by length scales that originate in the laminar (viscous) boundary layer. The most effective chordwise forcing location varies with St/Re1/2\textrm{St}/\textrm{Re}^{1/2} and incidence angle, α\alpha, and is always upstream of the separation point. Although the boundary layer flows are far from two-dimensional, forcing at a fixed chord location across all spanwise locations is effective in controlling the SI -- SII transition. Strategies for active and passive feedback control are suggested.
Article
Full-text available
As computational resources have advanced, scale-resolving methods like large eddy simulation (LES) have gained popularity in the aerodynamic design particularly in predicting massively separated flow regions. The limitations of traditional Reynolds-averaged Navier–Stokes (RANS) are evident in this regime, given the importance of the intrinsic unsteadiness and the strong three-dimensionality of the flow. This research focuses on wall-resolved implicit large eddy simulation (ILES) for airfoil flows in deep stall with a focus on its dynamics. Moreover, it is also of interest to understand how the spanwise extent of the computational domain affects the resolved aerodynamic forces and flow dynamics. The simulations in this study analyzed various spanwise extents (lz ranging from 0.10c to 2c) under specific conditions: a chord Reynolds number Re=105, an asymptotic Mach number M∞=0.15, and an angle of attack α = 20 degrees. Significant differences were found in aerodynamic coefficients and flow dynamics when the smallest spanwise extent was used. Specifically, spectral proper orthogonal decomposition showed that a small span failed to capture critical low-frequency dynamics necessary for predicting adequate stall behavior. Additionally, the turbulence field in the small span case remained largely rod-like, unlike the three-dimensional turbulence seen in larger spans. The current work propose the nondimensional spanwise integral length scale ( SILS/lz) as a metric for determining the sufficiency of spanwise extent, using two-point correlations for the streamwise velocity. Findings suggest that a spanwise extent of lz=1c is sufficient, with minimal differences observed between lz=1c and lz=2c. The results emphasize the importance of adequate spanwise resolution to accurately capture deep-stall flow structures and their unsteady behavior.
Article
Full-text available
Heterogeneity and a paucity of measurements of key material properties undermine the veracity of quantitative predictions of subsurface flow and transport. For such model forecasts to be useful as a management tool, they must be accompanied by computationally expensive uncertainty quantification, which yields confidence intervals, probability of exceedance, and so forth. We design and implement novel multilevel Monte Carlo (MLMC) algorithms that accelerate estimation of the cumulative distribution functions (CDFs) of quantities of interest, for example, water breakthrough time or oil production rate. Compared to standard non‐smoothed MLMC, the new estimators achieve a significant variance reduction at each discretization level by smoothing the indicator function with a Gaussian kernel or replacing standard Monte Carlo (MC) with the recently developed hierarchical Latinized stratified sampling (HLSS). After validating the kernel‐smoothed MLMC and HLSS‐enhanced MLMC methods on a single‐phase flow test bed, we demonstrate that they are orders of magnitude faster than standard MC for estimating the CDF of breakthrough times in multiphase flow problems.
Article
Full-text available
Uncertainty about geologic makeup and properties of the subsurface renders inadequate a unique quantitative prediction of flow and transport. Instead, multiple alternative scenarios have to be explored within the probabilistic framework, typically by means of Monte Carlo simulations (MCS). These can be computationally expensive, and often prohibitively so, especially when the goal is to compute the tails of a distribution, that is, probabilities of rare events, which are necessary for risk assessment and decision making under uncertainty. We deploy the method of distributions to derive a deterministic equation for the cumulative distribution function (CDF) of hydraulic head in an aquifer with uncertain (random) hydraulic conductivity. The CDF equation relies on a self‐consistent closure approximation, which ensures that the resulting CDF of hydraulic head has the same mean and variance as those computed with statistical moment equations. We conduct a series of numerical experiments dealing with steady‐state two‐dimensional flow driven by either a natural hydraulic head gradient or a pumping well. These experiments reveal that the CDF method remains accurate and robust for highly heterogeneous formations with the variance of log conductivity as large as five. For the same accuracy, it is also up to four orders of magnitude faster than MCS in computing hydraulic head with a required degree of confidence (probability).
Article
The development of a separated bottom boundary layer in the footprint of a large-amplitude internal solitary wave of depression, propagating against an oncoming barotropic current, is examined in detail using high-resolution implicit large eddy simulation. The wave is supported by a continuous two-layer stratification. The Reynolds number based on the water column height is 1.6×105. This numerical simulation is the first to reproduce the self-sustained three-dimensional vortex shedding, resultant transition, and turbulence under an ISW, which have long been hypothesized to occur in field experiments. No artificial noise is inserted into the flow domain. Part I of this study focuses on a structural description of the sequence of flow regimes developing from a wave-induced, long, high-aspect-ratio, laminar separation bubble. Three illuminating topological features are identified. (a) The spatial development of the self-sustained turbulence is composed of three transitional stages: (i) spontaneous excitation of a global instability in the separation bubble that emanates trailing vortices, (ii) vortex breakup and degeneration into turbulent clouds, and (iii) relaxation to a spatially developing turbulent boundary layer. (b) In the separation bubble, there exists a three-dimensional linear global oscillator, which is primarily excited by the two-dimensional absolute instability of the separated shear layer. This global mode possesses a transverse coherent structure. The transverse perturbation subsequently excites an elliptic instability mode inside the shed vortex, resulting in an axial distortion of the vortex core. (c) A shortwave secondary instability is excited in the form of a series of coherent streamwise vortex streaks that wrap around each shed vortex, leading to rapid break up and burst of the vortex.
Article
We present a novel deterministic model that is capable of predicting particle-to-particle force and torque fluctuations in a fixed bed of randomly distributed monodisperse spheres. First, we generate our dataset by performing particle-resolved direct numerical simulations (PR-DNS) of arrays of stationary spheres in moderately inertial regimes with a Reynolds number range of 2 ≤ Re ≤ 150 and a solid volume fraction range of 0.1 ≤ 𝜙 ≤ 0.4. The key idea exploited by our model is that, while the arrangement of neighbours around each particle is uniform and random, conditioning forces or torques exerted on a reference sphere to specific ranges of values results in the emergence of significantly non-uniform distributions of neighbouring particles. Based on probabilistic arguments, we take advantage of the statistical information extracted from PR-DNS to construct force/torque-conditioned probability distribution maps, which are ultimately used as basis functions for regression. Given the locations of surrounding particles as input to the model, our results demonstrate that the present probability-driven framework is capable of predicting up to 85 % of the actual observed force and torque variation in the best cases. Since the precise location of each particle is known in an Eulerian-Lagrangian (EL) simulation, our model would be able to estimate the unresolved subgrid force and torque fluctuations reasonably well, and thereby considerably enhance the fidelity of EL simulations via improved interphase coupling.
Article
We design and implement a novel algorithm for computing a multilevel Monte Carlo (MLMC) estimator of the joint cumulative distribution function (CDF) of a vector-valued quantity of interest in problems with random input parameters or initial conditions. Our approach combines MLMC with stratified sampling of the input sample space by replacing standard Monte Carlo at each level with stratified Monte Carlo initialized with proportionally allocated samples. We show that the resulting stratified MLMC (sMLMC) algorithm is more efficient than its standard MLMC counterpart due to the additional variance reduction provided by the stratification of the random parameter's domain, especially at the coarsest levels. Additional computational cost savings are obtained by smoothing the indicator function with a Gaussian kernel, which proves to be an efficient and robust alternative to recently developed polynomial-based techniques.
Article
The robustness and accuracy of marginally resolved discontinuous Galerkin spectral element computations are evaluated for the standard formulation and a kinetic energy conserving split form on complex flow problems of physical and engineering interest, including the flow over a square cylinder, an airfoil and a plane jet. It is shown that the kinetic energy conserving formulation is significantly more robust than the standard scheme for under-resolved simulations. A disadvantage of the split form is the restriction to Gauss-Lobatto nodes with the inherent underintegration and lower accuracy as compared to Gauss quadrature used with the standard scheme. While the results support the higher accuracy of the standard Gauss form, lower numerical robustness and spurious oscillations are evident in some cases, giving the advantage to the kinetic energy conserving scheme for marginally resolved numerical simulations.
Article
In this work, we present a rigorous derivation of the volume-filtered viscous compressible Navier–Stokes equations for disperse two-phase flows. Compared to incompressible flows, many new unclosed terms appear. These terms are quantified via a posteriori filtering of two-dimensional direct simulations of shock-particle interactions. We demonstrate that the pseudo-turbulent kinetic energy (PTKE) systematically acts to reduce the local gas-phase pressure and consequently increase the local Mach number. Its magnitude varies with volume fraction and filter size, which can be characterized using a Knudsen number based on the filter size and inter-particle spacing. A transport equation for PTKE is derived and closure models are proposed to accurately capture its evolution. The resulting set of volume-filtered equations are implemented within a high-order Eulerian–Lagrangian framework. An interphase coupling strategy consistent with the volume filtered formulation is employed to ensure grid convergence. Finally PTKE obtained from the volume-filtered Eulerian–Lagrangian simulations are compared to a series of two- and three-dimensional direct simulations of shocks passing through stationary particles.
Article
An uncertainty quantification framework is developed for Eulerian–Lagrangianmodels of particle-laden flows, where the fluid is modeled through a system of conservation laws in the Eulerian frame and inertial particles are traced as points in the Lagrangian frame. The source of uncertainty in such problems is the particle forcing, which is determined empirically or computationally with high-fidelity methods (data-driven). The framework relies on the averaging of the deterministic governing equations with the stochastic forcing and allows for an estimation of the first and second moment of the quantities of interest. Via comparison with Monte Carlo simulations, it is demonstrated that the moment equations accurately predict the uncertainty for problems whose Eulerian dynamics are either governed by the linear advection equation or the compressible Euler equations. In areas of singular particle interfaces and shock singularities significant uncertainty is generated. An investigation into the effect of the numerical methods shows that low-dissipative higher-order methods are necessary to capture numerical singularities (shock discontinuities, singular source terms, particle clustering) with low diffusion in the propagation of uncertainty.