Download as pdf or txt
Download as pdf or txt
You are on page 1of 52

Chemical and Biological Engineering Publications Chemical and Biological Engineering

2-2-2018

An open-source quadrature-based population


balance solver for OpenFOAM
Alberto Passalacqua
Iowa State University, albertop@iastate.edu

Frédérique Laurent
Université Paris-Saclay

Ehsan Madadi-Kandjani
Iowa State University

See next page for additional authors

Follow this and additional works at: https://lib.dr.iastate.edu/cbe_pubs


Part of the Complex Fluids Commons, Computer-Aided Engineering and Design Commons,
Engineering Physics Commons, and the Fluid Dynamics Commons
The complete bibliographic information for this item can be found at https://lib.dr.iastate.edu/
cbe_pubs/321. For information on how to cite this item, please visit http://lib.dr.iastate.edu/
howtocite.html.

This Article is brought to you for free and open access by the Chemical and Biological Engineering at Iowa State University Digital Repository. It has
been accepted for inclusion in Chemical and Biological Engineering Publications by an authorized administrator of Iowa State University Digital
Repository. For more information, please contact digirep@iastate.edu.
An open-source quadrature-based population balance solver for
OpenFOAM
Abstract
The extended quadrature method of moments (EQMOM) for the solution of population balance equations
(PBE) is implemented in the open-source computational fluid dynamic (CFD) toolbox OpenFOAM as part
of the OpenQBMM project. The moment inversion procedure was designed (Nguyen et al., 2016) to
maximize the number of conserved moments in the transported moment set. The algorithm is implemented
in a general structure to allow the addition of other kernel density functions defined on R+, and arbitrary
kernels to describe physical phenomena involved in the evolution of the number density function. The
implementation is verified with a set of zero-dimensional cases involving aggregation and breakage problems.
Comparison to the rigorous solution of the PBE provides validation for these cases. The coupling of the
EQMOM procedure with a CFD solver to address aggregation and breakage problems of non-inertial
particles is validated against experimental measurements in a Taylor-Couette reactor from the literature.

Keywords
Extended quadrature method of moments, Log-normal kernel density function, Population balance equation,
Aggregation and breakage, OpenFOAM, OpenQBMM

Disciplines
Complex Fluids | Computer-Aided Engineering and Design | Engineering Physics | Fluid Dynamics

Comments
This is a manuscript of an article published as Passalacqua, Alberto, Frédérique Laurent, E. Madadi-Kandjani,
J. C. Heylmun, and R. O. Fox. "An open-source quadrature-based population balance solver for OpenFOAM."
Chemical Engineering Science 176 (2018): 306-318. DOI: 10.1016/j.ces.2017.10.043. Posted with permission.

Authors
Alberto Passalacqua, Frédérique Laurent, Ehsan Madadi-Kandjani, Jeffrey C. Heylmun, and Rodney O. Fox

This article is available at Iowa State University Digital Repository: https://lib.dr.iastate.edu/cbe_pubs/321


An open-source quadrature-based population balance
solver for OpenFOAM
A. Passalacquaa,b,1,∗, F. Laurentb,c,1 , E. Madadi-Kandjania , J. C. Heylmuna ,
R. O. Foxd
a
Department of Mechanical Engineering, Iowa State University, Black Engineering
Building, Ames, IA 50011, USA
b
Laboratoire EM2C, CNRS, CentraleSupélec, Université Paris-Saclay, Grande Voie des
Vignes, 92295 Châtenay-Malabry cedex, France
c
Fédération de Mathématiques de l’Ecole Centrale Paris, FR CNRS 3487, France
d
Department of Chemical and Biological Engineering, Iowa State University, Sweeney
Hall, Ames, IA, 50011-2230, USA

Abstract

The extended quadrature method of moments (EQMOM) for the solution of


population balance equations (PBE) is implemented in the open-source com-
putational fluid dynamic (CFD) toolbox OpenFOAM as part of the Open-
QBMM project. The moment inversion procedure was designed (Nguyen
et al., 2016) to maximize the number of conserved moments in the trans-
ported moment set. The algorithm is implemented in a general structure to
allow the addition of other kernel density functions defined on R+ , and arbi-
trary kernels to describe physical phenomena involved in the evolution of the
number density function. The implementation is verified with a set of zero-
dimensional cases involving aggregation and breakage problems. Comparison
to the rigorous solution of the PBE provides validation for these cases. The


Corresponding author
Email address: albertop@iastate.edu (A. Passalacqua)
1
Equal contribution.

Preprint submitted to Chemical Engineering Science October 28, 2017


coupling of the EQMOM procedure with a CFD solver to address aggre-
gation and breakage problems of non-inertial particles is validated against
experimental measurements in a Taylor-Couette reactor from the literature.
Keywords: Extended quadrature method of moments, log-normal kernel
density function, population balance equation, aggregation and breakage,
OpenFOAM, OpenQBMM

1 1. Introduction

2 The spatial and temporal evolution of a discrete population of particles can


3 be described by a population balance equation (PBE) (Ramkrishna, 2000),
4 which is an evolution equation for the number density function (NDF) asso-
5 ciated to the particle population. The NDF can evolve due to discontinuous
6 phenomena such as nucleation, aggregation, breakage and evaporation, and
7 due to continuous phenomena such as convection and diffusion. Examples of
8 industrial processes, involving the evolution of a particle population include,
9 but certainly are not limited to, precipitation, polymerization and combus-
10 tion (Becker et al., 2014), sprays (Laurent and Massot, 2001) and aerosols
11 (McGraw, 1997; McGraw and Wright, 2003).
12 In this work, we concentrate on the case of a NDF with only one inter-
13 nal coordinate, representing the particle size. Approximate solutions of the
14 corresponding PBE can be determined using several approaches, including
15 Monte-Carlo methods (Lin et al., 2002; Meimaroglou and Kiparissides, 2007;
16 Rosner and Yu, 2001; Smith and Matsoukas, 1998; Zhao et al., 2007; Zhao and
17 Zheng, 2013), which, however, present challenges in practical applications due
18 to their computational cost. Some authors introduce a discretization along

2
19 the size variable, leading to the sectional method or the method of classes
20 (Alopaeus et al., 2006; Balakin et al., 2014; Bannari et al., 2008; Becker
21 et al., 2011; Hounslow et al., 1988; Hounslow, 1990; Kumar and Ramkrishna,
22 1996a,b; Muhr et al., 1996; Puel et al., 2003; Vanni, 2000). Similarly to
23 Monte Carlo methods, this approach is often computationally too demand-
24 ing when applied to large-scale problems of industrial interest, as observed
25 by Marchisio et al. (2003b). To overcome this issue, hybrid methods between
26 sectional and moment methods are developed (see Nguyen et al. (2016) and
27 references therein), but they are not the subject of this paper.
28 A widely adopted and sufficiently accurate approach to find approximate so-
29 lutions of the PBE for engineering applications is the quadrature method of
30 moments (QMOM), originally introduced by McGraw (1997), and extensively
31 applied to several problems in chemical engineering (see Gavi et al. (2007);
32 Marchisio and Fox (2013); Petitti et al. (2010) for examples). The QMOM
33 approach considers a discrete set of moments of the NDF, constituted by
34 an even number of moments. The NDF is then approximated with a dis-
35 crete weighted sum of Dirac delta functions, uniquely determined by means
36 of a moment inversion algorithm (Gautschi, 2004; Gordon, 1968; Wheeler,
37 1974). The extended QMOM (EQMOM) (Yuan et al., 2012) introduced the
38 capability of using a basis of non-negative kernel density functions (KDF)
39 to approximate the NDF in place of Dirac delta functions. This develop-
40 ment allows some of the limitations of QMOM, that appear when dealing
41 with problems that require the evaluation of the NDF at a particular value
42 of the internal coordinate (i.e. problems involving evaporation term or any
43 other continuous size decreasing term (Massot et al., 2010)), to be addressed.

3
44 Yuan et al. (2012) proposed the EQMOM procedure for β and Γ KDF, while
45 Madadi-Kandjani and Passalacqua (2015) considered log-normal KDF. The
46 EQMOM reconstruction can be done for every realizable moment set (i.e.
47 moments of a positive NDF), just not reproducing, eventually, the last mo-
48 ment. In particular, it can deal with the degenerate cases, encountered when
49 the moments are not strictly realizable: the only possible representation of
50 the NDF is then a sum of weighted Dirac delta functions, thus describing
51 a population of particles of only one or a few sizes, as in the case of nucle-
52 ation. Numerically, this possibility was achieved with the moment inversion
53 algorithm of Nguyen et al. (2016).
54 In this work we discuss the implementation of the EQMOM approach into the
55 open-source toolbox for computational fluid dynamics (CFD) OpenFOAM
56 (2015), as part of the OpenQBMM (2016a) project. We limit our attention
57 to a univariate PBE, where the internal coordinate of the NDF is the particle
58 size. We describe the implementation of EQMOM with log-normal KDF as
59 an example, but without loss of generality in the presentation of the compu-
60 tational framework, which was designed to accommodate any KDF defined
61 on the set of positive real numbers R+ . We then discuss the implementation
62 of realizable kinetic fluxes for advection, which guarantee the transported
63 moments are realizable if the step used for time integration satisfies a realiz-
64 ability condition similar to the Courant-Friedrichs-Lewy condition. Particu-
65 lar attention is put in detailing the implementation of the procedure used to
66 determine the approximate NDF, which always ensures that the maximum
67 possible number of moments is conserved (Nguyen et al., 2016). The PBE
68 solver is then verified considering aggregation and breakage problems stud-

4
69 ied by Vanni (2000), comparing the predicted results with both the rigorous
70 solution from Vanni (2000) and the numerical solution obtained with EQ-
71 MOM by Madadi-Kandjani and Passalacqua (2015). Finally, a case involving
72 spatial transport is considered for validation purposes, which consists of an
73 aggregation and breakage problem in a Taylor–Couette reactor. The system
74 was experimentally studied by Serra and Casamitjana (1998a,b); Serra et al.
75 (1997), considering the same test case discussed in Marchisio et al. (2003a).
76 Numerical results obtained with the CFD-PBE solver developed as part of
77 the present work are compared to experiments, showing satisfactory results.

78 2. The population balance equation

The PBE (Marchisio et al., 2003a; Marchisio and Fox, 2013; Ramkrishna,
2000; Randolph and Larson, 1988) accounting for the evolution of a univariate
NDF with internal coordinate ξ, representing the particle size, is

∂n (ξ, x, t)
+ ∇x · [n (ξ, x, t) U] − ∇x · [Γ∇x n (ξ, x, t)]
∂t
+ ∇ξ · [G(ξ)n (ξ, x, t)] = B̄ a (ξ, x, t) − D̄a (ξ, x, t)

+ B̄ b (ξ, x, t) − D̄b (ξ, x, t) + N (ξ, x, t) (1)

79 where n(ξ, x, t) is the NDF, U is the velocity of the carrier fluid, Γ is the
80 diffusivity, G(ξ) is the growth rate, B̄ (ξ, x, t) and D̄ (ξ, x, t) are, respectively,
81 the rate of change of n due to birth and death, in the aggregation process
82 when the a exponent is present and in the breakage process when a b expo-
83 nent is present, and N (ξ, x, t) the rate of change due to nucleation. Let us
84 notice that we assumed the particle size is sufficiently small to have negligible

5
85 influence on the carrier fluid. This allows the velocity U to be assumed equal
86 to the local fluid velocity, and independent of the particle size.
87 The diffusivity Γ, assumed to be independent from the particle size ξ, is
88 defined as the sum of a laminar and a turbulent contribution: Γ = Γl + Γt .
89 The turbulent diffusivity is calculated as the ratio of the turbulent viscosity
90 µt and the turbulent Schmidt number σt : Γt = µt /σt .
Following Marchisio and Fox (2013); Marchisio et al. (2003b); ?, the terms
describing aggregation and breakage phenomena are written in continuous
form as:

B̄ a (ξ, x, t)
 
1/3
ξ2 Z ξ β (ξ 3 − ξ 03 ) , ξ 0  
03 1/3
3
n (ξ 0 , x, t) dξ 0 , (2)

= 2/3
n ξ − ξ , x, t
2 0 3
(ξ − ξ ) 03

91
Z ∞
a
D̄ (ξ, x, t) = n (ξ, x, t) β (ξ, ξ 0 ) n (ξ 0 , x, t) dξ 0 , (3)
0

92
Z ∞
b
B̄ (ξ, x, t) = a (ξ 0 ) b (ξ|ξ 0 ) n (ξ 0 , x, t) dξ 0 , (4)
ξ

93

D̄b (ξ, x, t) = a (ξ) n (ξ, x, t) . (5)

94 Growth and nucleation terms were not considered in the example applica-
95 tions presented in this work to verify and validate the implementation of
96 the EQMOM procedure, however they have been implemented in the PBE
97 solver, and their testing is left to future work. These terms and their numer-
98 ical approximation in quadrature methods are kept in the description of the
99 theory presented in this work for completeness and as documentation of the

6
100 code implementation for the interested reader. Their numerical integration
101 is performed in the OpenQBMM framework in an identical manner to that
102 used for the source terms due to aggregation and breakup (Nguyen et al.,
103 2016).

104 3. The extended quadrature method of moments

105 The approximate solution of the PBE of Eq. (1) is obtained in this work by
106 solving transport equations for a finite set of the moments of the NDF. In
107 the case of a univariate NDF, the moments are defined as:
Z +∞
Mk (t) = n (ξ, x, t) ξ k dξ. (6)
0

The transport equation for the moment of order k is obtained by multiplying


the PBE (Eq. (1)) by ξ k and integrating over [0, +∞[. Under the previously
discussed assumptions on the velocity and the diffusivity, such transport
equation is (Marchisio and Fox, 2013; Marchisio et al., 2003b)
∂Mk (x, t)
+ ∇x · [Mk (x, t) U] − ∇x · [Γ∇Mk (x, t)] + Ḡk (x, t)
∂t
= B̄ka (x, t) − D̄ka (x, t) + B̄kb (x, t) − D̄kb (x, t) + N̄k (x, t) . (7)

108 The evaluation of the growth, aggregation and breakup terms in Eq. (7)
109 requires the NDF to be known, in addition to the kernel functions a and
110 β. However, because the NDF evolves in space and time, it is not known a
111 priori, and it has to be approximated from the transported moments.
112 This can be achieved by writing an approximant of the NDF as a weighted
113 sum of non-negative KDFs δσ (ξ, ξα ) (Yuan et al., 2012)
N
X
n (ξ) ≈ pN (ξ) = wα δσ (ξ, ξα ) (8)
α=1

7
114 where N is the number of KDFs used to approximate the NDF, the KDF
115 δσ (ξ, ξα ) is chosen to formally tend to the Dirac delta function δ (ξ − ξα )
116 when σ tends to zero, wα are the non-negative weights of each KDF (pri-
117 mary quadrature weights), and ξα are the corresponding quadrature abscis-
118 sae (primary abscissae). These parameters (wα , ξα )α=1,...,N and σ have to
119 be computed from (Mk )2N
k=0 in such a way that this moment set represents

120 the moments of the reconstructed NDF. We use log-normal KDFs (Madadi-
121 Kandjani and Passalacqua, 2015) in all the problems considered in this work
122 because the support of the NDF to be reconstructed in these problems is R+ :
123

1 (ln ξ−ξα )2
δσ (ξ, µ) = √ e− 2σ 2 , ξ, ξα , σ ∈ R+ , (9)
ξσ 2π
124 however, also the Γ KDF discussed in Yuan et al. (2012) is implemented in
125 the OpenQBMM PBE solver. The primary quadrature is determined with
126 the algorithm described in Nguyen et al. (2016), whose implementation is
127 detailed in Sec. 4.4. Once the NDF in Eq. (8) is reconstructed, it is used
128 to calculate integrals that appear in source terms for the moment transport
129 equations, as described in Yuan et al. (2012). To this purpose, a secondary
130 quadrature, with weights wαγ and abscissae ξαγ , is determined by consider-
131 ing the recurrence relation that defines the family of orthogonal polynomials
132 with respect to the measure defined from the KDF, and solving the eigen-
133 value problem associated to the Jacobi matrix defined by this relationship
134 (Gautschi, 2004). In performing this calculation, the Stieltjes-Wigert quadra-
135 ture is considered for log-normal KDF (Weisstein, 1998; Wilck, 2001), while
136 Laguerre quadrature is used for gamma KDF (Gautschi, 2004). While, in
137 principle, the case of a log-normal KDF could be addressed with a change

8
138 of variable, allowing Hermite quadrature to be adopted (Madadi-Kandjani
139 and Passalacqua, 2015), it is worth noticing that only the Stieltjes-Wigert
140 quadrature guarantees the correct preservation of the moments of the NDF,
141 and as such it is used in this work.
142 The source term describing the rate of change of the moments of the NDF
143 due to molecular growth (Marchisio and Fox, 2013) is
Z ∞
Ḡk (x, t) = −k ξ k−1 G(ξ)n (ξ, x, t) dξ. (10)
0

144 Substituting the approximation of the NDF in terms of the extended quadra-
145 ture leads to:
N
X Nα
X
k−1
Ḡk (x, t) ≈ −k wα wαγ ξαγ G(ξαγ , x, t). (11)
α=1 γ=1

Similarly, the source terms representing the rate of change of the moments
due to birth and death of particles because of aggregation and breakage are
(Madadi-Kandjani and Passalacqua, 2015; Marchisio et al., 2003b):
1 ∞
Z Z ∞
0
k/3
a
B̄k (x, t) = n (ξ , x, t) β (ξ, ξ 0 ) ξ 3 + ξ 03 n (ξ, x, t) dξdξ 0
2 0 0
N N α N Nα
1X X X X k/3
≈ w α1 wα1 γ1 w α2 wα2 γ2 ξα3 1 γ1 + ξα3 2 γ2 βα1 γ1 α2 γ2 , (12)
2 α =1 γ =1 α =1 γ =1
1 1 2 2

Z ∞ Z ∞
D̄ka (x, t) = k
ξ n (ξ, x, t) β (ξ, ξ 0 ) n (ξ 0 , x, t) dξ 0 dξ
0 0
N
X N
X α N
X Nα
X
≈ w α1 wα1 γ1 ξαk 1 γ1 w α2 wα2 γ2 βα1 γ1 α2 γ2 , (13)
α1 =1 γ1 =1 α2 =1 γ2 =1

Z ∞ Z ∞
B̄kb (x, t) = ξ k
a (ξ 0 ) b (ξ|ξ 0 ) n (ξ 0 , x, t) dξ 0 dξ
0 0
N
X Nα
X
≈ wα wαγ aαγ b̄kαγ , (14)
α=1 γ=1

9
146
Z ∞ N
X Nα
X
D̄kb (x, t) = ξ k a (ξ) n (ξ, x, t) dξ ≈ wα k
wαγ ξαγ aαγ , (15)
0 α=1 γ=1

147 where βα1 γ1 α2 γ2 is the aggregation kernel computed using the abscissae ξα1 γ1
148 and ξα2 γ2 , and b̄kαγ is the value of b̄k (Table 2) calculated for ξ = ξαγ .

149 4. The numerical solution procedure

150 The transport equations for the moments (Eq. (7)) are discretized using
151 a finite-volume procedure (Ferziger and Peric, 2002; LeVeque, 2002), with
152 kinetic-based spatial fluxes (Desjardins et al., 2008; Perthame, 1990) to en-
153 sure moment realizability.

154 4.1. Spatial fluxes

155 The discretization of the moment advection term is key to ensure moment
156 realizability, and achieve a robust solution procedure. The problem of mo-
157 ment realizability caused by convection schemes is well known in the lit-
158 erature (Desjardins et al., 2008; Wright, 2007), and it appears when using
159 conventional numerical schemes of order higher than one, since the appli-
160 cation of limiters to the individual transported moments is not a sufficient
161 condition to ensure the realizability of the transported moment set. Vikas
162 et al. (2011, 2013) proposed to implement kinetic schemes (Perthame, 1990)
163 where the convection flux for each moment is computed from the quadra-
164 ture approximation of the transported moment set. In these schemes, the
165 high-order reconstruction is applied only to the quadrature weights, while
166 the reconstructed quadrature abscissae are constant in the cell, hence the
167 name of quasi high-order schemes. This approach allows the realizability

10
168 of the moment set to be preserved if a constraint on the time-step size is
169 satisfied. This realizability constraint becomes the Courant-Friedrichs-Lewy
170 (CFL) condition, if weights and abscissae are reconstructed on cell faces with
171 a first-order scheme, which, therefore, always ensures moment realizability if
172 the CFL condition is satisfied.
173 The numerical scheme used to compute face values of a variable in Open-
174 FOAM can be selected independently for any variable, or specifying the
175 scheme as argument of the fvc::interpolate function provided by the
176 OpenFOAM framework. As a consequence, the details of the implemen-
177 tation of high-order reconstruction schemes are not discussed here, and our
178 description is limited to the general structure of kinetic-based schemes for
179 quadrature-based moment methods.
180 We assume the flow velocity vector U is known at cell centers. We indicate
181 with Uown and Unei the interpolated face values of such velocity on the owner
182 and neighbour side of the cell face, respectively. Similarly, we indicate with
183 Mk,own and Mk,nei the moments computed on the owner and neighbour side
184 of the cell face (see Fig. 1), defined as
Nf
X Nf
X
k k
Mk,own = wα,own ξα,own , Mk,nei = wα,own ξα,nei , (16)
α=1 α=1

185 where Nf is the number of quadrature nodes used to define the quadrature
186 approximation of the NDF used to compute the spatial fluxes, wα,own and
187 wα,nei are the face, and ξα,own , ξα,nei the face values of the quadrature ab-
188 scissae. If the quasi-second-order scheme of Vikas et al. (2011) is used, a
189 limited second-order reconstruction with the minmod limiter can be used to
190 obtain the face values of the quadrature weights, while the face values of the

11
Neighbour
Neighbour
Owner

Neighbour

Figure 1: Schematic representation of a computational cell and its neighbour cells, to


illustrate the concept of owner and neighbour used in the calculation of moment spatial
fluxes.

191 quadrature abscissae are equal to the values of the same abscissae at cell cen-
192 ters, since only a first-order reconstruction is allowed for the abscissae. This
193 scheme ensures the realizability of the convected moment set if the time step
194 satisfies, in each cell of the computational grid, the condition (Vikas et al.,
195 2011) !

∆t ≤ min P |Ω| , (17)
f wα,own max(Uown · Sf , 0)

196 where |Ω| is the cell volume, Sf the face surface area vector, and the sum-
197 mation at the denominator in Eq. (17) is computed only considering faces
198 with outgoing flux with respect to the owner cell of the face.
199 It is important to observe at this point that, following Laurent and Nguyen
200 (2017), spatial fluxes are calculated as a function of a different quadrature

12
201 approximation with respect to the one determined with the EQMOM pro-
202 cedure. If an odd number of moments is realizable, this quadrature approx-
203 imation is obtained using a N + 1 point Gauss-Radau quadrature formula,
204 which allows the 2N + 1 moments to be preserved, with one quadrature ab-
205 scissa set to zero. However, if the number of realizable moments is even, the
206 regular Gauss quadrature is used to define the face flux. This approach is
207 independent from the reconstruction of the NDF, and allows the fluxes to be
208 defined without relying on the secondary quadrature approximation, leading
209 to a more robust formulation of the numerical scheme, with the additional
210 benefit of a slightly lower computational cost due to the reduced number of
211 terms involved in the summations in Eq. 18 compared to EQMOM. Details on
212 the rationale of the definition of the advection schemes for moment methods
213 are provided in Laurent and Nguyen (2017), where fully high-order schemes
214 for moment methods are also developed for structured meshes. Their exten-
215 sion to unstructured grids and their implementation into the OpenQBMM
216 framework are not trivial, and will be the subject of future work.
217 The convective flux of the moment Mk is then found computing

ϕMk = Mk,nei min(Unei · Sf , 0) + Mk,own max(Uown · Sf , 0), (18)

218 where Sf is parallel to the unit normal vector to the cell face n, oriented in
219 the direction of the outgoing flux with respect to the owner cell. Once ϕMk is
220 calculated, the rate of change of the moment Mk due to convection is found
221 by integrating ϕMk over the cell faces.

13
222 4.2. Diffusion term

223 The discretization of diffusion term in the moment transport equations does
224 not present particular problems when a traditional second-order scheme (Ferziger
225 and Peric, 2002) is used, as explained in Vikas et al. (2013). As a consequence,
226 no special treatment is adopted to ensure moment realizability when the diffu-
227 sion term is discretized, because second-order accuracy is considered sufficient
228 for most practical applications. The diffusion term in the moment transport
229 equations is therefore discretized with the fvm::laplacian(gamma, Mk) op-
230 erator provided by OpenFOAM, where gamma is the diffusivity and Mk is
231 the moment of order k. It is worth observing that this approach suffices
232 for size-independent diffusion, however, if size-dependent diffusion needs to
233 be considered, the diffusion term will have to be computed as a function of
234 the quadrature approximation, and treated as an explicit source term in the
235 transport equations for the moments, as discussed in Vikas et al. (2013).

236 4.3. Source terms

237 The source terms for aggregation, breakage growth, and nucleation may be
238 introduced as explicit source terms in the moment transport equations, using
239 the standard mechanism available in OpenFOAM. This however typically
240 leads to significant restrictions on the global time step of the simulation. This
241 limitation was overcome in the present work using the multi-step adaptive
242 scheme proposed by Nguyen et al. (2016). This approach leverages the fact
243 that explicit Euler methods with small enough time steps can be used to
244 ensure the realizability of the moment set obtained after the integration of the
245 source terms. Any convex combination of Euler explicit time steps (Nguyen

14
246 et al., 2016), such as the strong-stability-preserving (SSP) explicit Runge-
247 Kutta methods (Gottlieb et al., 2011), which are high-order ODE solvers,
248 enjoy the same properties.

249 4.4. Moment inversion

250 A key aspect of the numerical solution procedure is the moment inversion to
251 find the primary quadrature weights wα , abscissae ξα and the parameter σ
252 of the KDF, in such a way that the moments of the reconstruction defined
253 by Eq. (8) are the considered moment set. The approach implemented in the
254 OpenQBMM-PBE solver is based on the work of Nguyen et al. (2016), which
255 represents an improvement of the method proposed in Yuan et al. (2012)
256 and also used in Madadi-Kandjani and Passalacqua (2015). The key differ-
257 ence between the method in Yuan et al. (2012) and the new one proposed
258 in Nguyen et al. (2016) consists in an efficient procedure to check the strict
259 realizability of the moment set, and the conservation of the largest possi-
260 ble subset of transported moments, which is not always achieved using the
261 approach of Yuan et al. (2012).
262 The algorithm is based on the possibility to use an efficient quadrature al-
263 gorithm based, for example, on the Wheeler (1974) algorithm. Introduc-
PN
264 ing the variables Mk∗ = k
α=1 wα ξα , the KDF is chosen in such a way

265 that, for a fixed value of σ, there exists a linear relation between the mo-
266 ments MK = (M0 , . . . , MK ) of the reconstruction and M∗K = (M0∗ , . . . , MK∗ ):
267 MK = AK (σ)M∗K . For a log-normal KDF, it is given by

2 σ 2 /2
Mk∗ = Mk e−k , (19)

268 in such a way that the matrix AK (σ) is diagonal. Then, from the transported

15
269 moment vector M2N = (M0 , . . . , M2N ), one can define the function M̄2N (σ)
270 in the following way: if M∗2N−1 (σ) = A2N−1 (σ)−1 M2N−1 is strictly realizable,
271 which corresponds to σ < σmax,2N , then one computes the corresponding
272 quadrature weights and abscissae (wα (σ), ξα (σ))α=1,...,N and M2N (σ) is the
273 moment of order 2N of the reconstructed NDF defined by Eq. (8) with the
274 parameters σ and (wα (σ), ξα (σ))α=1,...,N . When σ ≥ σmax,2N , the function
275 M̄2N (σ) is set to a large value2 L. For a log-normal KDF, this function is
276 given by

 N wα (σ)ξα (σ)2N z (2N)
2
P
σ < σmax,2N

α=1
M̄2N (σ) = (20)
L σ ≥ σmax,2N ,

2 /2
277 where z = eσ . This helper function is used to define the target function
278 J2N , whose roots are the values of the parameter σ we seek, with
M2N − M̄2N (σ)
J2N (σ) = . (21)
M2N
279 If Eq. 21 has no root, then the last moment will not be reproduced by the
280 reconstructed NDF. One then will have to minimize the error on it, by an
281 optimal choice of σ. In any case, at least all moments of orders 0 to 2N − 1
282 will be well reproduced by the reconstruction.
283 It can happen that the transported moment vector M2N is not strictly real-
284 izable (degenerate case), when it corresponds to a NDF that is necessarily a
285 sum of a few weighted Dirac delta functions. This is the case for example,
286 in problems involving nucleation, and at the beginning of aggregation pro-
287 cesses. Only one representation is then possible and the moment inversion

2
A value of 1.0 × 1016 was used in the implementation, which corresponds to the
definition of the GREAT constant in OpenFOAM.

16
288 has to be adapted in this case. The first step of the moment inversion pro-
289 cedure is then to study the realizability of the moment vector. Differently
290 from what done in Madadi-Kandjani and Passalacqua (2015); Yuan et al.
291 (2012), which directly evaluated the Hankel determinants associated with
292 the set of transported moments, this is achieved computing the values of the
293 ζk quantities defined as in Dette and Studden (1997), which allow the size
294 Nr ≤ 2N + 1 of the largest subset of strictly realizable moments MNr −1 to
295 be found. This approach is more computationally efficient than evaluating
296 the Hankel determinants directly, as briefly discussed later, and explained in
297 detail in Nguyen et al. (2016). The quantities ζk are directly related to the
298 Hankel determinants

M0 . . . Mk M1 . . . Mk+1

. .. .. ..

H 2k = .. . , H 2k+1 = . . , (22)


Mk . . . M2k Mk+1 . . . M2k+1

299 as follows3
H k+1 H k−2
ζk = , k = 0, 1 . . . , (23)
H k H k−1
300 with H −2 = H −1 = H 0 = 1. They are also related to the coefficients of the
301 recurrence relation that defines the family of orthogonal polynomials (Pk )k≥0
302 with respect to a measure associated to the moments:

Pk+1 (X) = (X − ak )Pk (X) − bk Pk−1 (X), (24)

303 with a0 = ζ1 and for k ≥ 1: bk = ζ2k−1 ζ2k and ak = ζ2k +ζ2k+1 . Thanks to this
304 relation, the values of ζk are computed using the Q-D algorithm described

3
Note that ζk is here re-defined to start the numbering from zero rather than from one,
for consistency with the implementation.

17
305 in Dette and Studden (1997), or, more effectively, with a modified version
306 of the Chebyshev algorithm (Wheeler, 1974) to include the calculation of ζk
307 (Nguyen et al., 2016). Following Dette and Studden (1997), the positivity
308 of ζk ensures the strict realizability of the moment set. Based on the largest
309 value of k for which ζk > 0, it is then possible to identify the subset of strictly
310 realizable moments in the transported moment set.
311 Depending on the value of strictly realizable moments Nr , the following cases
312 are possible:

313 1. If Nr < 2, the procedure terminates because the number of strictly


314 realizable moments is insufficient to define a quadrature formula.
315 2. If Nr is an even integer, the Wheeler (1974) algorithm is applied, and
316 Nr /2 primary quadrature weights and abscissae are calculated from the
317 Nr realizable moments. In this case, the parameter σ is set to zero, and
318 no attempt is made to reconstruct a continuum distribution because the
319 direct application of the standard QMOM procedure, relying on Dirac
320 delta functions, ensures the conservation of the Nr strictly realizable
321 moments.
322 3. If Nr is an odd integer, the complete EQMOM procedure is applied to
323 the Nr first Nr − 1 moments, i.e. from MNr −1 . This allows to determine
324 the (Nr −1)/2 primary quadrature weights and abscissas, the parameter
325 σ, and the secondary quadrature. The objective of the procedure is to
326 determine the value of σ ∈ R+ as a root of the non-linear function JNr −1
327 thus ensuring the conservation of the entire set of realizable moments,
328 when it exists, or to minimize the error on the last moment of the set,
329 whenever a value of σ that guarantees its exact conservation cannot be

18
330 found. The procedure is iterative, and operates as follows:
331 (a) An initial interval where to search for the value of σ is determined
332 considering the value of the σ parameter that ensures that M∗3 (σ)
333 is strictly realizable. We will indicate this value of σ as σmax,4 , and
334 consider I0 =]0, σmax,4 [ as initial interval. For a log-normal KDF,
335 we have
√ √
s s 
M0 M2 M1 M3 
σmax,4 = min  2 log , 2 log (25)
M1 M2

336 It is important to observe, however, that σmax,4 is not a valid


337 value for the σ parameter, because the condition on the strict
338 realizability of M∗k (σ) for k ≥ 4 would further reduce the right
339 bound of the interval of σ. This however does not represent a
340 problem in our procedure, since the interval we determine is only
341 used to initially start the search for σ.
342 (b) The function described in Eq. (21) changes sign in I0 by construc-
343 tion, as a consequence of the definition of M̄2N (σ) in Eq. (20).
344 We then apply Ridder’s algorithm (Ridders, 1979) to the func-
345 tion JNr −1 (σ) to find a root for the target function. If this root
346 represents an actual solution that allows the moment set to be
347 preserved, then the value of σ found from Ridder’s algorithm will
348 be used. If the computed value of σ does not preserve the mo-
349 ment set, a minimization procedure based on the golden search
350 algorithm (Press et al., 2007) is applied to JNr −1 (σ)2 , in order to
351 find the value of σ which minimizes the distance of JNr −1 (σ) from
352 zero.

19
353 (c) Once the value of the parameter σ is determined, the secondary
354 quadrature weights and abscissae are computed by means of the
355 standard eigenvalue problem used to calculate the roots of poly-
356 nomials orthogonal to the chosen KDF (Gautschi, 2004).

357 The steps of the EQMOM inversion algorithm are schematically represented
358 in Fig. 2, where NI indicates the number of invertible moments in a moment
359 set.

360 5. Code structure

361 The solution procedure for univariate PBE was implemented into Open-
362 FOAM creating a general data and implementation structure which allow
363 the extension to multivariate problems, and the straightforward addition of
364 sub-models such as kernels describing new physics. Additionally, the func-
365 tionality of OpenFOAM was leveraged to couple the implementation of the
366 PBE solver with fluid solvers, re-using the existing thermal and turbulence
367 models. In the following subsections we provide a brief description of the
368 code structures and implementation choices made during the development of
369 OpenQBMM-PBE.

370 5.1. Basic moment inversion

371 We call basic moment inversion the process of calculating N quadrature


372 weights and abscissae from a moment vector of dimension 2N. This pro-
373 cess is achieved by means of the algorithm proposed by Wheeler (1974).
374 A specific C++ object, called univariateMomentSet, was derived for the
375 uni-dimensional array of scalars object scalarDiagonalMatrix provided by

20
Start

Mk

Compute ζk

Performed by univariateMomentSet
Determine Nr
and NI

Yes Not enough moments to


NI < 2
define quadrature
No

Find NI/2 primary


weights and
abscissae

Yes Use QMOM and


Nr even
do not try to find σ

No Use Ridder’s
method to find σ
Compute σmax and
define I0

No Normalized
moment error
< momentTol

Minimize J2(σ)
over I0 to find σ Yes
Compute secondary
quadrature

End

Figure 2: Schematic representation of the EQMOM procedure.

21
376 the OpenFOAM library. The univariateMomentSet object combines the
377 use of ζk to verify the strict realizability of the moment set, with the Wheeler
378 (1974) algorithm to perform the moment inversion. The procedure is schemat-
379 ically represented in Fig. 3. For each moment set that has to be inverted, the
380 maximum even number of strictly realizable moments, which we call number
381 of invertible moments4 NI , is determined by checking the positivity of ζk and
382 taking the largest even integer smaller than the number of strictly realizable
383 moments Nr . Once NI is found, the procedure verifies that NI ≥ 2, which
384 is required to perform the inversion. If this condition is satisfied, the Ja-
385 cobi matrix (see Appendix A) is constructed and the associated eigenvalue
386 problem is solved to determine quadrature weights and abscissae.

387 5.2. Extended moment inversion

388 The extended moment inversion procedure is implemented in a general form,


389 in order to accommodate different types of KDF with support on R+ . The
390 type of KDF can be selected by the user at run-time as conventionally done
391 in OpenFOAM for model selection, by means of a keyword in an input file.
392 A base class named extendedMomentInversion implements the algorithm
393 to find the value of the parameter σ, and the general parts of the procedure
394 to define the secondary quadrature (Fig. 2). Specialized classes derived from
395 extendedMomentInversion implement the part of the algorithm specific to
396 a KDF such as the expression of the KDF itself, the relationship between
397 Mk and Mk∗ , and the upper extremum σmax of the interval I0 .

4
The number of invertible moments coincides with the number of strictly realizable
moments, if this is even, and it is equal to Nr − 1, if the Nr is odd.

22
Start

Mk

Already
inverted?
No

Determine ζk and find


number of invertible
moments NI

Yes Not enough realizable


NI < 2 moments to define
quadrature
No

Construct
coefficients of the
recurrence relation
(ai and bi)

Compute Jacobi’s
matrix and find its
eigenvalues and
eigenvectors Wheeler’s algorithm

Calculate
quadrature weights End
and abscissae

Figure 3: Schematic representation of the ζk -Wheeler procedure implemented in the


univariateMomentSet object.

23
398 5.3. Population balance solver

399 The PBE is solved in a separate class, which also uses the run-time selec-
400 tion mechanism, allowing future extensions to multivariate problems to be
401 implemented. A base class called populationBalanceModel implements the
402 generic structure required for run-time selection of the type of PBE to be
403 solved. The actual implementation of the PBE solver is contained in the
404 derived class univariatePopulationBalance, which provides methods to
405 calculate the convection, diffusion and source terms for the specific PBE.

406 6. Verification of the PBE solver

407 The PBE solver was verified considering a set of zero-dimensional problems
408 involving aggregation and breakage (Vanni, 2000). The solutions obtained
409 in these cases were compared to the numerical solution obtained with a
410 MATLAB code implementing EQMOM in Madadi-Kandjani and Passalac-
411 qua (2015), which relied on the solution approach detailed in Yuan et al.
412 (2012) extended to log-normal KDF, and to the rigorous solution of the
413 same problems reported by Vanni (2000).
414 In this situation, Eq. (1) reduces to Eq. (26), where only the source terms
415 due to aggregation and breakup remain, while diffusion and advection are
416 not considered:

∂n (ξ, x, t)
= B̄ a (ξ, x, t) − D̄a (ξ, x, t) + B̄ b (ξ, x, t) − D̄b (ξ, x, t) . (26)
∂t

417 The cases considered in this validation study are reported in Tab. 1, and
418 correspond to those examined in Madadi-Kandjani and Passalacqua (2015).
419 The table also indicates the number of primary and secondary quadrature

24
420 nodes used in each case. The expressions for the daughter distribution that
421 appear in the breakage kernel are summarized in Tab. 2.
422 The simulation results shown in Figs. 4–8 show the time evolution of the
423 value of M0 and of the average particle size d43 = M4 /M3 . The agreement
424 between the numerical solution obtained with MATLAB (Madadi-Kandjani
425 and Passalacqua, 2015), and the one obtained with the OpenFOAM im-
426 plementation of the EQMOM procedure is excellent: in all the cases under
427 examination, the two solutions match. This comparison serves as verification
428 of the EQMOM procedure implemented into OpenFOAM. The comparison
429 to the rigorous solution of Vanni (2000) shows good agreement in all the
430 cases under consideration, with the exception of case 3 (Fig. 6), in which
431 some deviation of the numerical solution from the rigorous solution can be
432 observed for t ∈ [1, 4] s.

433 7. Validation of CFD-PBE solver

434 The solution of the full PBE with an inhomogeneous velocity field is tested
435 and validated using the case studied by Marchisio et al. (2003a) in a Taylor–
436 Couette reactor made by two concentrical cylinders with diameter D1 =
437 193 mm, and D2 = 160 mm. The height of the system is H = 360 mm.

438 7.1. Kernels for the PBE

439 The aggregation kernel used in this case is given by the sum of the kernel for
440 Brownian motion (Eq. (27)) (Smoluchowski, 1917), combined with the kernel
441 proposed by Adachi et al. (1994) for particles whose size is smaller than the

25
Case N Nα β (ξ, ξ 0 ) a (ξ) b (ξ|ξ 0 ) Mk (t = 0)

0

ξ=1
1 2 7 1 1, Tab. 2 Mk = 1, k = 0, . . . , 6
0.02 ξ > 1


0

ξ=1
2 2 7 ξ 3 + ξ 03 1, Tab. 2 Mk = 1, k = 0, . . . , 6
0.02ξ 3

ξ>1

0

ξ=1
3 2 7 (ξ + ξ 0 )3 1, Tab. 2 Mk = 1, k = 0, . . . , 6
0.1 e0.01ξ3

ξ>1

M0 =1







M1 = 1.13








 

 M2 = 1.294
0

ξ=1


4 3 7 (ξ + ξ 0 )2 |ξ 2 − ξ 02 | 2, Tab. 2 M3 = 1.5
0.01ξ 6

ξ>1




M4 = 1.760






M5 = 2.087








M = 2.087
6


0 3
ξ< 5

5 4 10 (ξ + ξ 0 )3 √ 3, Tab. 2 Mk = 1, k = 0, . . . , 6
0.1 e0.01ξ3

ξ≥ 35

Table 1: Cases examined for the aggregation and breakage process.

26
1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0.8

0.6
M0

0.4

0.2

0
0 50 100 150 200
Time, t / [s]

(a) M0

4
d43 / [µm]

1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0
0 50 100 150 200
Time, t / [s]

(b) d43

Figure 4: Time evolution of M0 and of d43 in case 1.

27
1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0.8

0.6
M0

0.4

0.2

0
0 5 10 15 20
Time, t / [s]

(a) M0

4
d43 / [µm]

1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0
0 5 10 15 20
Time, t / [s]

(b) d43

Figure 5: Time evolution of M0 and of d43 in case 2.

28
1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0.8

0.6
M0

0.4

0.2

0
0 2 4 6 8 10
Time, t / [s]

(a) M0

5
d43 / [µm]

2
Madadi and Passalacqua (2015)
1 Rigorous solution (Vanni, 2000)
OpenQBMM
0
0 2 4 6 8 10
Time, t / [s]

(b) d43

Figure 6: Time evolution of M0 and of d43 in case 3.

29
1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0.8

0.6
M0

0.4

0.2

0
0 2 4 6 8 10
Time, t / [s]

(a) M0

4
d43 / [µm]

1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0
0 2 4 6 8 10
Time, t / [s]

(b) d43

Figure 7: Time evolution of M0 and of d43 in case 4.

30
1
Madadi and Passalacqua (2015)
Rigorous solution (Vanni, 2000)
OpenQBMM
0.8

0.6
M0

0.4

0.2

0
0 2 4 6 8 10
Time, t / [s]

(a) M0

5
d43 / [µm]

2
Madadi and Passalacqua (2015)
1 Rigorous solution (Vanni, 2000)
OpenQBMM
0
0 2 4 6 8 10
Time, t / [s]

(b) d43

Figure 8: Time evolution of M0 and of d43 in case 5.

31
No. Mechanism b (ξ|ξ 0 ) b̄k

2 ξ = ξ 0 /21/3

1 Symmetric fragmentation 2(3−k)/3 ξ k
0 otherwise


6ξ 2 /ξ 03 ξ ∈]0, ξ 0 [

2 Uniform 6ξ k /(k + 3)
0

otherwise

1/3
1 ξ = ξ 0 51






k/3
3 Mass ratio 1 to 4 0 4 1/3 ξ k 4 5k/3+1

 1 ξ = ξ 5



0 otherwise

Table 2: Daughter distribution functions and their moment transformations.

442 local Kolomogorov microscale (Eq. (28)):

0 2kB (ξ + ξ 0 )2
β(ξ, ξ ) = (27)
3µ ξξ 0

443 r
0 4 3πε 3
β(ξ, ξ ) = (ξ + ξ 0 ) . (28)
3 10ν
444 The breakage kernel of Luo and Svendsen (1996), which depends on the
445 kinematic viscosity ν, the turbulent dissipation rate ε and the particle size
446 ξ, was used:
a(ξ) = cBr ν p εq ξ r . (29)

447 Following Marchisio et al. (2003a), we set p = 3/4, q = −5/4 and r = 1,


448 and we adopt the symmetric fragmentation daughter distribution function
449 (Tab. 2). The coefficient cBr was set equal to 6.0 × 10−4 (Marchisio et al.,
450 2003a).

32
Figure 9: Schematic representation of the Taylor–Couette reactor and of the computational
domain used in the simulations.

451 7.2. Case setup

452 The Taylor–Couette reactor was modeled assuming symmetry, reducing the
453 computational domain to a rectangular region whose width is (D1 − D2 )/2 =
454 16.5 mm, and its height is H = 180 mm, as schematized in Fig 9. This do-
455 main was discretized using 18 cells in the horizontal direction, and 180 in the
456 vertical direction (Marchisio et al., 2003a), leading to a total of 3240 compu-
457 tational cells. Since the particle concentration in the cases examined in this
458 work is low, and particles are sufficiently small to consider them having low
459 inertia, the effect of the presence of particles on the flow was neglected. The
460 flow field was simulated using the steady-state incompressible solver in Open-
461 FOAM and the Reynolds stress turbulence model of Launder et al. (1975),
462 which is provided by OpenFOAM, with the generalized gradient diffusion

33
463 model of Daly and Harlow (1970). Default values of the coefficient of the
464 model were used. Convection terms for all the conservation equations were
465 discretized with a limited second-order central scheme with Sweby (1984)
466 limiter. Once the converged flow field was obtained, the solution of the PBE
467 was started, assuming the flow field was frozen.
468 The initial particle size distribution is assumed to be a Dirac delta function
469 (uniform size), with a mean particle volume fraction αp = 2.5 × 10−5 , and
470 an initial particle size ξ0 = 2 µm (Marchisio et al., 2003a). The first five
471 moments of the NDF are considered and, as a consequence, N = 2 primary
472 quadrature nodes are used. The number of secondary quadrature nodes was
473 set equal to Nα = 10. It is worth observing that the initial condition leads to
474 a quadrature representation with only one quadrature node, and null value
475 of the σ parameter. Therefore, the moment inversion procedure will use the
476 standard Wheeler (1974) algorithm to perform the inversion. However, the
477 EQMOM procedure is used as soon as the particle size distribution is not a
478 Dirac delta anymore, and some variance in the particle size appears. The
479 transition between the two algorithms is managed automatically, based on
480 the realizability of the moment set and of the number of moments required
481 to define the quadrature representation of the NDF.
482 Marchisio et al. (2003a) considered four rotational speeds (75, 125, 165 and
483 211 RPM), however we limit our study to two rotational speeds for brevity,
484 since the purpose of our investigation is to verify the robustness and accuracy
485 of the computational framework. We consider then ω1 = 75 RPM, and
486 ω2 = 165 RPM.

34
(a) 75 RPM (b) 165 RPM

Figure 10: Velocity magnitude (m/s) of the fluid in the Taylor–Couette reactor.

487 7.3. Results and discussion

488 The velocity field of the fluid phase in the Taylor–Couette reactor is reported
489 in Fig. 10 for the two cases considered in this work. The volume-averaged
490 turbulent dissipation rate calculated when the flow inside the Taylor–Couette
491 reactor reached steady state is 0.0048 m2 /s3 for the case with rotational speed
492 of 75 RPM, and of 0.0770 m2 /s3 for the case with rotational speed of 165
493 RPM. These values are consistent with those reported by Marchisio et al.
494 (2003a), who reported a value of 0.070 m2 /s3 for the case with rotational
495 speed of 165 RPM, slightly lower than the result obtained in the present work.
496 The difference may be potentially explained by differences in the turbulence
497 model and in the numerical setup between the simulations performed in this

35
498 work and those reported by Marchisio et al. (2003a).
499 The evolution in time of the normalized mean particle diameter d = d43 /ξ0
500 predicted by the simulations is compared to the experimental data of Serra
501 and Casamitjana (1998a,b); Serra et al. (1997) as reported in Fig. 11. The
502 numerical predictions obtained by Marchisio et al. (2003a) in the same case,
503 but using the standard QMOM procedure are reported for comparison. The
504 same value of the coefficient cBr = 6.0 × 10−4 , found to give the best match
505 with experimental results by Marchisio et al. (2003a), was used for all the
506 simulations.
507 Fig. 11(a) shows that the numerical simulation provides a good estimate of
508 the steady-state value of d43 for the case with rotational speed of 75 RPM,
509 which is however predicted to be smaller than what experimentally observed
510 both by the results obtained with the standard QMOM procedure in Marchi-
511 sio et al. (2003a) and by EQMOM in the present work. Both the EQMOM
512 and the QMOM approach predict the particle diameter reaches its steady-
513 state value significantly earlier (t ≈ 2000 s) than what observed in the ex-
514 periments (t ≈ 12000 s).
515 The EQMOM simulation slightly over-predicts the steady-state value of the
516 mean normalized particle diameter (≈ 11) compared to the experimental
517 results (≈ 10) in the case with rotational speed of 165 RPM. The results
518 obtained with EQMOM show a higher overprediction of the steady state
519 value of d43 compared to the results obtained by Marchisio et al. (2003a) with
520 QMOM. This difference may be due to possible differences in the mean flow
521 field and, in particular, in the spatial distribution of the turbulent dissipation
522 rate. The steady state value of the particle size is reached after about 1000 s

36
523 from the beginning of the simulation both with EQMOM and in the QMOM
524 results from Marchisio et al. (2003a), however experimental data show that
525 a slower evolution of the value of d43 , with a time required to reach steady
526 state of about 2000 s.
527 The simulations of the two cases discussed above were performed on a dual-
528 CPU workstation with Intel Xeon processors using 15 cores. The steady-state
529 solution of the flow field was reached after 46 minutes of simulation time, in
530 the case with rotational speed of 75 RPM, and after 25 minutes, in the case
531 with rotational speed of 165 RPM. The solution of the PBE with rotational
532 speed of 75 RPM required ≈ 4.39 days (precisely, 379,623 s) to complete
533 ≈ 4.16 hours (15,000 s) of flow time. This time slightly increased to ≈ 5
534 days (431,727 s) to complete ≈ 1.39 h (5000 s) of flow time in the case with
535 rotational speed of 165 RPM. The difference in computational time may be
536 explained by the explicit nature of the advection fluxes, which lead to a
537 time-step size controlled by the Courant-Friedrichs-Lewy condition.

538 8. Conclusions

539 The EQMOM was implemented into OpenFOAM to solve univariate popula-
540 tion balance equations. The implementation was successfully verified against
541 the numerical results of Madadi-Kandjani and Passalacqua (2015) involving
542 aggregation and breakage problems in a homogoeneous system. The results
543 were validated against the rigorous solutions obtained by Vanni (2000). The
544 PBE solver was then coupled to a CFD solver to enable the solution of the
545 PBE for inhomogeneous systems with transport of particles due to the fluid
546 motion. This was achieved by implementing kinetic-based fluxes for the mo-

37
547 ment transport equations. The implementation was validated by simulating
548 shear-driven aggregation and breakage in the Taylor–Couette reactor consid-
549 ered by Marchisio et al. (2003a). Satisfactory results were shown concerning
550 the prediction of the steady-state value of the mean particle diameter.

551 9. Code repository

552 The source code of the OpenQBMM-PBE software can be downloaded from
553 the git repository https://github.com/OpenQBMM, and is distributed under
554 the terms of the GNU General Public License version 3. The specific version
555 of the code, and the tutorials used to produce the results reported in this work
556 is OpenQBMM 2.0.0 (OpenQBMM, 2016b). The repository also contains all
557 the test cases mentioned in this article, and the corresponding validation
558 data.

559 10. Acknowledgments

560 The authors would like to gratefully acknowledge the support of the US
561 National Science Foundation under the SI2 –SSE award NSF–ACI 1440443
562 and the support of the French National Research Agency (ANR) under grant
563 ANR-13-TDMO-02 ASMAPE for the ASMAPE project.
564 A.P. would like to acknowledge the support of the ”Jean d’Alembert” fel-
565 lowship program Idex Paris-Saclay, which supported his stay at Laboratoire
566 EM2C, CNRS, CentraleSupélec, Université Paris-Saclay

38
567 Appendix A. Inversion algorithm

568 We briefly summarize here the moment inversion algorithm, based on the
569 work of Wheeler (1974), to perform the moment inversion using the same
570 notation adopted in the implementation. We assume the number of invert-
571 ible moments is NI . Let us denote with ai and bi , i ∈ {0, 1, . . . , NI /2} the
572 coefficients of the recurrence relation that define the family of orthogonal
573 polynomials with respect to a measure associated to the moments. We also
574 introduce the matrix S ∈ R2NI +1,2NI +1 , which is defined as follows in the code:
575
M1
a0 =
M0
b0 = 0

S1,i = Mi−1 , i = 1, . . . , NI ,

k = 2, . . . , NI /2 + 1, l = k, . . . , NI − k + 2, (A.1)
Sk,l = Sk−1,l+1 − ak−2 Sk−1,l − bk−2 Sk−2,l ,
Sk,k+1 Sk−1,k
ak−1 = − ,
Sk,k Sk−1,k−1
Sk,k
bk−1 = ,
Sk−1,k−1
576 The matrix z ∈ RNI /2,NI /2 is then defined as

zi,i = ai , i = 0, . . . , NI /2 − 1,
p
zi,i+1 = bi+1 , i = 0, . . . , NI /2 − 2, (A.2)

zi+1,i = zi,i+1 , i = 0, . . . , NI /2 − 1
577 Quadrature weights and abscissae are then found as
2
wi = M0 vz,1,i ,
(A.3)
ξi = λz,i ,

39
578 where λz,i is the i-th eigenvalue of z and vx,1,i is the first component of the
579 corresponding eigenvector.

40
25

20

15
d

10

5
Experiments (Serra et al., 1997)
OpenQBMM CBr = 0.6e-3
Marchisio et al. (2003) CBr = 0.6e-3
0
0 2000 4000 6000 8000 10000 12000 14000
Time [s]

(a) 75 RPM

12

10

6
d

2
Experiments (Serra et al., 1997)
OpenQBMM CBr = 0.6e-3
Marchisio et al. (2003) CBr = 0.6e-3
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time [s]

(b) 165 RPM

Figure 11: Time evolution of the volume-average normalized particle diameter in the
Taylor–Couette reactor.
41
(a) 75 RPM (b) 165 RPM

Figure 12: Steady-state field of the volume-averaged normalized particle diameter in the
Taylor–Couette reactor.

42
580 References

581 Adachi, Y., Stuart, M. A. C., Fokkink, R., Jul. 1994. Kinetics of turbulent
582 coagulation studied by means of end-over-end rotation. Journal of Colloid
583 and Interface Science 165 (2), 310–317.

584 Alopaeus, V., Laakkonen, M., Aittamaa, J., Oct. 2006. Solution of popu-
585 lation balances with breakage and agglomeration by high-order moment-
586 conserving method of classes. Chemical Engineering Science 61 (20), 6732–
587 6752.

588 Balakin, B. V., Hoffmann, A. C., Kosinski, P., 2014. Coupling STAR-CD
589 with a population-balance technique based on the classes method. Powder
590 Technology 257, 47–54.

591 Bannari, R., Kerdouss, F., Selma, B., Bannari, A., Proulx, P., 2008. Three-
592 dimensional mathematical modeling of dispersed two-phase flow using class
593 method of population balance in bubble columns. Computers & Chemical
594 Engineering 32 (12), 3224–3237.

595 Becker, P. J., Puel, F., Dubbelboer, A., Janssen, J., Sheibat-Othman, N.,
596 Sep. 2014. Coupled population balance–CFD simulation of droplet breakup
597 in a high pressure homogenizer. Computers & Chemical Engineering 68,
598 140–150.

599 Becker, P. J., Puel, F., Henry, R., Sheibat-Othman, N., 2011. Investigation of
600 discrete population balance models and breakage kernels for dilute emul-
601 sification systems. Industrial & Engineering Chemistry Research 50 (19),
602 11358–11374.

43
603 Daly, B. J., Harlow, F. H., 1970. Transport equations in turbulence. Physics
604 of Fluids (1958-1988) 13, 2634 – 2649.

605 Desjardins, O., Fox, R., Villedieu, P., 2008. A quadrature-based moment
606 method for dilute fluid-particle flows. Journal of Computational Physics
607 227 (4), 2514–2539.

608 Dette, H., Studden, W. J., 1997. The Theory of Canonical Moments with
609 Applications in Statistics, Probability and Analysis. John Wiley & Sons.

610 Ferziger, J. H., Peric, M., 2002. Computational Methods for Fluid Dynamics.
611 Springer.

612 Gautschi, W., 2004. Orthogonal Polynomials: Computation and Approxima-


613 tion. Oxford University Press.

614 Gavi, E., Rivautella, L., Marchisio, D. L., Vanni, M., Barresi, A. A., Baldi,
615 G., 2007. CFD modelling of nano-particle precipitation in confined im-
616 pinging jet reactors. Chemical Engineering Research and Design 85 (5),
617 735–744.

618 Gordon, R. G., 1968. Error bounds in equilibrium statistical mechanics. Jour-
619 nal of Mathematical Physics 9, 655–662.

620 Gottlieb, S., Ketcheson, D., Shu, C.-W., 2011. Strong Stability Preserving
621 Runge-Kutta and Multistep Time Discretizations. World Scientific.

622 Hounslow, M., Ryall, R., Marshall, V., 1988. Discretized population balance
623 for nucleation, growth, and aggregation. AIChE Journal 34 (11), 1821–
624 1832.

44
625 Hounslow, M. J., 1990. A discretized population balance for continuous sys-
626 tems at steady state. AIChE Journal 36 (1), 106–116.

627 Kumar, S., Ramkrishna, D., Apr. 1996a. On the solution of population bal-
628 ance equations by discretization – I. A fixed pivot technique. Chemical
629 Engineering Science 51 (8), 1311–1332.

630 Kumar, S., Ramkrishna, D., Apr. 1996b. On the solution of population bal-
631 ance equations by discretization – II. A moving pivot technique. Chemical
632 Engineering Science 51 (8), 1333–1342.

633 Launder, B. E., Reece, G. J., Rodi, W., 1975. Progress in the development
634 of a Reynolds-stress turbulence closure. Journal of Fluid Mechanics 68 (3),
635 537–566.

636 Laurent, F., Massot, M., 2001. Multi-fluid Modeling of Laminar Poly-
637 dispersed Spray Flames: Origin, Assumptions and Comparison of the Sec-
638 tional and Sampling Methods. Combustion Theory and Modelling 5 (4),
639 537–572.

640 Laurent, F., Nguyen, T. T., 2017. Realizable second-order finite-volume


641 schemes for the advection of moment sets of the particle size distribution.
642 Journal of Computational Physics 337, 309–338.

643 LeVeque, R. J., 2002. Finite Volume Methods for Hyperbolic Problems, 1st
644 Edition. Cambridge University Press.

645 Lin, Y., Lee, K., Matsoukas, T., 2002. Solution of the population balance
646 equation using constant–number Monte Carlo. Chemical Engineering Sci-
647 ence 57 (12), 2241–2252.

45
648 Luo, H., Svendsen, H. F., May 1996. Theoretical model for drop and bubble
649 breakup in turbulent dispersions. AIChE Journal 42 (5), 1225–1233.

650 Madadi-Kandjani, E., Passalacqua, A., Jul. 2015. An extended quadrature-


651 based moment method with log-normal kernel density functions. Chemical
652 Engineering Science 131, 323–339.

653 Marchisio, D., Dennis Vigil, R., O. Fox, R., Aug. 2003a. Implementation of
654 the quadrature method of moments in CFD codes for aggregation–breakage
655 problems. Chemical Engineering Science 58 (15), 3337–3351.

656 Marchisio, D. L., Fox, R. O., 2013. Computational Models for Polydisperse
657 Particulate and Multiphase Systems. Cambridge University Press.

658 Marchisio, D. L., Vigil, R., Fox, R. O., 2003b. Quadrature method of mo-
659 ments for aggregation-breakage processes. Journal of Colloid and Interface
660 Science 258 (2), 322–334.

661 Massot, M., Laurent, F., Kah, D., de Chaisemartin, S., 2010. A robust
662 moment method for evaluation of the disappearance rate of evaporating
663 sprays. SIAM Journal on Applied Mathematics 70 (8), 3203–3234.

664 McGraw, R., 1997. Description of aerosol dynamics by the quadrature


665 method of moments. Aerosol Science and Technology 27 (2), 255–265.

666 McGraw, R., Wright, D. L., 2003. Chemically resolved aerosol dynamics
667 for internal mixtures by the quadrature method of moments. Journal of
668 Aerosol Science 34 (2), 189–209.

46
669 Meimaroglou, D., Kiparissides, C., 2007. Monte Carlo simulation for the
670 solution of the bi-variate dynamic population balance equation in batch
671 particulate systems. Chemical Engineering Science 62 (1820), 5295–5299.

672 Muhr, H., David, R., Villermaux, J., Jezequel, P. H., 1996. Crystallization
673 and precipitation engineering-VI. Solving population balance in the case
674 of the precipitation of silver bromide crystals with high primary nucleation
675 rates by using the first order upwind differentiation. Chemical Engineering
676 Science 51 (2), 309–319.

677 Nguyen, T. T., Laurent, F., Fox, R. O., Massot, M., 2016. Realizable second-
678 order finite-volume schemes for the advection of moment sets of the particle
679 size distribution. Journal of Computational Physics 325, 129–156.

680 OpenFOAM, 2015. OpenFOAM User Guide. The OpenFOAM Foundation.


681 URL www.openfoam.org

682 OpenQBMM, 2016a. An open-source implementation of quadrature-based


683 moment methods.
684 URL www.openqbmm.org

685 OpenQBMM, 2016b. OpenQBMM 2.0.0 Stable for OpenFOAM 4.x. DOI:
686 10.5281/zenodo.161617.
687 URL https://github.com/OpenQBMM/OpenQBMM/releases/tag/
688 OpenQBMM_2.0.0_stable

689 Perthame, B., 1990. Boltzmann type schemes for compressible Euler equa-
690 tions in one and two space dimensions. SIAM Journal of Numerical Anal-
691 ysis 29 (1), 1–19.

47
692 Petitti, M., Nasuti, A., Marchisio, D. L., Vanni, M., Baldi, G., Mancini, N.,
693 Podenzani, F., 2010. Bubble size distribution modeling in stirred gas-liquid
694 reactors with QMOM augmented by a new correction algorithm. AIChE
695 Journal 56 (1), 36–53.

696 Press, W. H., Teukolsky, S. H., Vetterling, W. T., Flannery, B. P., 2007. Nu-
697 merical Recipes: The Art of Scientific Computing. Cambridge University
698 Press.

699 Puel, F., Fvotte, G., Klein, J. P., 2003. Simulation and analysis of indus-
700 trial crystallization processes through multidimensional population bal-
701 ance equations. part 1: a resolution algorithm based on the method of
702 classes. Chemical Engineering Science 58 (16), 3715–3727.

703 Ramkrishna, D., 2000. Population Balances: Theory and Applications to


704 Particulate Systems in Engineering. Academic Press, San Diego, CA.

705 Randolph, A. D., Larson, M. A., 1988. Theory of Particulate Processes:


706 Analysis and Techniques of Continuous Crystallization. Academic Press.

707 Ridders, C., Nov. 1979. A new algorithm for computing a single root of a real
708 continuous function. IEEE Transactions on Circuits and Systems 26 (11),
709 979–980.

710 Rosner, D. E., Yu, S., 2001. MC simulation of aerosol aggregation and simul-
711 taneous spheroidization. AIChE Journal 47 (3), 545–561.

712 Serra, T., Casamitjana, X., Aug. 1998a. Effect of the shear and volume frac-
713 tion on the aggregation and breakup of particles. AIChE Journal 44 (8),
714 1724–1730.

48
715 Serra, T., Casamitjana, X., Oct. 1998b. Structure of the aggregates during
716 the process of aggregation and breakup under a shear flow. Journal of
717 Colloid and Interface Science 206 (2), 505–511.

718 Serra, T., Colomer, J., Casamitjana, X., Mar. 1997. Aggregation and breakup
719 of particles in a shear flow. Journal of Colloid and Interface Science 187 (2),
720 466–473.

721 Smith, M., Matsoukas, T., 1998. Constant-number Monte Carlo simulation
722 of population balances. Chemical Engineering Science 53 (9), 1777–1786.

723 Smoluchowski, M. Z., 1917. Versuch einer mathematischen theorie der koag-
724 ulationskinetik kolloider losunger. Zeitschrift fur Physikalische Chemie 92,
725 129 – 142.

726 Sweby, P., Oct. 1984. High Resolution Schemes Using Flux Limiters for Hy-
727 perbolic Conservation Laws. SIAM Journal on Numerical Analysis 21 (5),
728 995–1011.

729 Vanni, M., 2000. Approximate population balance equations for aggregation-
730 breakage processes. Journal of Colloid and Interface Science 221 (2), 143–
731 160.

732 Vikas, V., Wang, Z., Passalacqua, A., Fox, R., 2011. Realizable high-order
733 finite-volume schemes for quadrature-based moment methods. Journal of
734 Computational Physics 230 (13), 5328–5352.

735 Vikas, V., Wang, Z. J., Fox, R. O., 2013. Realizable high-order finite-volume
736 schemes for quadrature-based moment methods applied to diffusion popu-
737 lation balance equations. Journal of Computational Physics 249, 162–179.

49
738 Weisstein, E. W., 1998. CRC Concise Encyclopedia of Mathematics. CRC
739 Press.

740 Wheeler, J. C., 1974. Modified moments and Gaussian quadratures. Rocky
741 Mountain Journal of Mathematics 4, 287–296.

742 Wilck, M., 2001. A general approximation method for solving integrals con-
743 taining a lognormal weighting function. Journal of Aerosol Science 32 (9),
744 1111–1116.

745 Wright, D. L., 2007. Numerical advection of moments of the particle size
746 distribution in eulerian models. Journal of Aerosol Science 38 (3), 352–
747 369.

748 Yuan, C., Laurent, F., Fox, R., 2012. An extended quadrature method of
749 moments for population balance equations. Journal of Aerosol Science 51,
750 1–23.

751 Zhao, H., Maisels, A., Matsoukas, T., Zheng, C., 2007. Analysis of four
752 Monte Carlo methods for the solution of population balances in dispersed
753 systems. Powder Technology 173 (1), 38–50.

754 Zhao, H., Zheng, C., 2013. A population balance-Monte Carlo method for
755 particle coagulation in spatially inhomogeneous systems. Computers &
756 Fluids 71, 196–207.

50

View publication stats

You might also like