GQMOM Sept1
GQMOM Sept1
Abstract
The quadrature method of moments (QMOM) for a one-dimensional (1-D) population balance equation
was introduced by R. McGraw (Aerosol Science and Technology, 27, 255-265, 1997) to close the moment
source terms. QMOM is defined based on the properties of the monic orthogonal polynomials Qi of degrees
i = 0, 1, . . . , n that are uniquely defined by the set of 2n moments up to order 2n − 1. The moment of order
2n is fixed to the boundary of moment space such that the distribution function is approximated by a sum
of n Dirac delta functions. Using the recursion coefficients of the orthogonal polynomials for i > n ≥ 1,
the generalized quadrature method of moments (GQMOM) extends the quadrature representation to a sum
of N > n terms using the same moments as QMOM. In doing so, the known moments are preserved and
higher-order moments correspond to a distribution function in the interior of moment space. Here, GQMOM
closures for distributions on R, R+ , and (0, 1) are defined and analyzed. Generally speaking, GQMOM
provides a more accurate moment closure than QMOM without increasing the number of moments and at
nearly the same computational cost.
Keywords: population balance equation, quadrature-based moment methods, moment closures
1. Introduction
The quadrature method of moments (QMOM) (McGraw, 1997) is arguably the most successful and
widely used closure for finding the lower-order moments of the number density function (NDF) found from a
one-dimensional (1-D) population balance equation (PBE). There are numerous publications in the scientific
literature demonstrating its accuracy for treating particle growth, aggregation and breakage processes. In
many applications, only a relatively small number of moments are needed to attain sufficient accuracy. This
is because realistic growth, aggregation and breakage kernels are relatively smooth functions of the particle
size, which is a necessary condition for the Gaussian-quadrature approximation of integrals with respect to
the NDF to be accurate (Grosch et al., 2007). For example, most aerosol dynamics problems can be treated
accurately with between three to five quadrature nodes using QMOM (Marchisio and Fox, 2013; McGraw,
1997), which corresponds to six to ten moments. In this work, we use the ‘standard’ moments written as
integer powers of the phase-space variable. It is also possible to use generalized moments (Grosch et al.,
2007; Lage, 2011); however, the limitations associated with the QMOM that we wish to address remain the
same (Grosch et al., 2007).
Despite its many successes, the QMOM has an obvious shortcoming, i.e., in order to increase the number
of quadrature nodes, one must solve for a larger number of moments. In comparison, closures based on
reconstructing the NDF from a fixed set of moments do not suffer from this problem. For example, entropy
∗ Correspondingauthor
Email addresses: [email protected] (Rodney O. Fox), [email protected] (Frédérique Laurent),
[email protected] (Alberto Passalacqua)
Here we provide a brief overview of QMOM for a spatially homogeneous, 1-D PBE. Application to
spatially inhomogeneous cases is described in detail in Marchisio and Fox (2013).
In general, S(t, ξ) will be a functional of f (t, ξ) and its moments (Marchisio and Fox, 2013). The kth moment
of the NDF is
Z
Mk (t) = ξ k f (t, ξ) dξ (2)
B
for k ∈ N. Let M2n−1 := (M0 , M1 , . . . , M2n−1 ) be the moment vector of length 2n with n ∈ N.
The ordinary differential equation (ODE) system for the moment vector found from the PBE then has
the form
Z
dMk
= ξ k S(t, ξ) dξ = S k (t) (3)
dt B
where S k cannot generally be expressed exactly in terms of the moments of f contained in M2n−1 . QMOM
provides a closure for S k given M2n−1 (McGraw, 1997) of the form
n
X
S k (t) = wi Sk (t, ξi ) (4)
i=1
with non-negative weights wi (t) and abscissae ξi (t). QMOM has the property that if S(t, ξ) = f (t, ξ), then
Sk (t, ξi ) = ξik and S k (t) = Mk (t) for k = 0, 1, . . . , 2n−1; meaning that the weights and abscissae corresponds
to a Gauss quadrature. Computing these parameters from the moments is a well-known problem solved
thanks to the theory of orthogonal polynomials (Gautschi, 2004).
Then, this linear functional defines a scalar product (P, Q) 7→ hP Qi on R[X]n , with n = bN/2c1 , as soon as
MN is strictly realizable, i.e., is associated with a positive NDF f ∈ L2 (B) through eq. (2). In such cases,
we have
Z
∀P ∈ R[X]N hP iMN = P (ξ)f (ξ)dξ, (6)
R
1 brc is the largest integer less than or equal to the real number r
3
and we can define a sequence of monic orthogonal polynomials Qi for i = 0, . . . , n with Qi of degree i. This
sequence satisfied a three-term recurrence relation:
hXQ2i i hQ2i i
ai = , b i = . (8)
hQ2i i hQ2i−1 i
and
M1−d − M2−d M2−d − M3−d ... Mk − Mk+1
M2−d − M3−d M3−d − M4−d ... Mk+1 − Mk+2
H 2k+d = .. .. .. .. . (10)
. . . .
Mk − Mk+1 Mk+1 − Mk+2 ... M2k+d−1 − M2k+d
• If B = R+ : H k > 0 for k = 0, . . . , N .
• If B = (0, 1): H k > 0 and H k > 0 for k = 0, . . . , N .
On the boundary of moment space (i.e., for weakly realizable moments), one (or more) of the Hankel
determinants is null.
An equivalent characterization can be done using the coefficients of the three-term recurrence relation
of the orthogonal polynomials:
• If B = R: bi > 0 for i = 1, . . . , b N2 c.
• If B = R+ : existence of ζi > 0 for i = 1, . . . , N such that
i = 1, . . . , N 2−1
a0 = ζ1 , ai = ζ2i + ζ2i+1
(11)
bi = ζ2i−1 ζ2i i = 1, . . . , N2
4
• If B = (0, 1): existence of ζi > 0 for i = 1, . . . , N satisfying eq. (11) and existence of pi ∈ (0, 1) such
that
ζ1 = p1 , ζi = pi (1 − pi−1 ) i = 2, . . . , N. (12)
Moreover, in the last case, the pi are referred to as the canonical moments (Dette and Studden, 1997).
Hereinafter, we assume that the known moment vector M2n is strictly realizable. In any case, this
assumption can be verified by applying the Chebyshev algorithm, which computes the recurrence coefficients
ai and bi from the known moments.
The basic idea behind GQMOM is to retain the properties of QMOM while using N > n ≥ 1 quadrature
points found from M2n with the Chebyshev algorithm. The two principal desired properties are that the
source-term closure has the form
N
X
S k (t) = wi Sk (t, ξi ), (13)
i=1
i.e., the same as with QMOM but with n replaced by N ≥ n ≥ 1. Roughly speaking, this can be accomplished
by choosing realizable moments Mk for k = 2n+1, 2n+2, . . . , 2N −1, and applying the Chebyshev algorithm
to M2N −1 . This is equivalent to selecting recurrence coefficients ai for i = n, n + 1, . . . , N − 1; and bi > 0
for i = n + 1, . . . , N − 1 that satisfy the realizability constraints in section 2.3. It is important to recall that
if the recurrence coefficients are realizable, then we are guaranteed that the quadrature nodes lie in B, and
their weights are positive. Expressed in terms of the monic orthogonal polynomials Qi (x), this selection is
done as follows.
5
3.1. Definition of GQMOM using recurrence coefficients
With GQMOM, the abscissae correspond to the roots of the monic orthogonal polynomial QN with
N > n ≥ 1. Thus, we must define the unknown recurrence coefficients (ai , bi ) for i = n, n + 1, . . . , N . First,
we start by defining the monic orthogonal polynomials.
where the unknown recurrence coefficients depend on M2n and are formally defined by eq. (8). For example,
[M ] [M ]
bn 2n = bn is known, while an 2n depends on the closure for M 2n+1 (or vice versa). Hereinafter, an over-
[M ]
line is used to denote a closure for a higher-order moment given M2n . In general, the choice for bi 2n with
[M2n ]
i > n closes the moment M 2i . As long as bi > 0, this closure will be realizable for B = R (Hamburger,
1944), and similar constraints apply for B = R+ and B = (0, 1).
The unknown weight function is µ[M2n ] > 0, i.e., a distribution function on B with known moments
Z
Mk = ξ k dµ[M2n ] for k = 0, 1, . . . , 2n; (16)
B
where
[M2n ] H2i
γi = ≥0 (19)
H2i−2
[M ]
depends on moments up to M2i through the Hankel determinant H2i = H 2i . If, for some k, γk 2n = 0,
then bm = 0 for m ≥ k. This occurs when the Hankel determinant H2k = 0 and the moments M2n are on
[M ]
the boundary of moment space. For clarity, unless stated otherwise, we assume that γk 2n > 0, and thus
that M2n is in the interior of moment space for B = R. More generally, the GQMOM closure will depend
on the definition of the NDF, in particular on B.
aH
i = µ, bH 2
i = iσ . (20)
10-2
10-4
10-6
10-8
10-10
-8 -6 -4 -2 0 2 4 6 8
Figure 1: GQMOM quadrature weights versus abscissae for standardized Gaussian moments with n = 5, N = 101, and different
values of ν.
Definition 1 (GQMOM with moments M2n for n ≥ 1 from a NDF defined on R). The unknown re-
currence coefficients are selected as follows. For i = n in eq. (15), let
n−1
1X
a[M
n
2n ]
= ai , bn[M2n ] = bn . (21)
n i=0
q
2ν = 0 corresponds to Chebychev polynomials of the second kind for the distribution 1 − (ξ/ξmax )2 on the finite interval
√
(−ξmax , ξmax ) with ξmax = 2 b1 , which are the special case of the shifted Jacobi polynomials with α = β = 1/2.
7
100
10-5
10-10
10-15
0 10 20 30 40 50 60 70
Figure 2: Gamma–GQMOM quadrature weights versus abscissae for gamma moments with α = 10, N = 101, and different
values of n.
the GQMOM closure, we can select the additional ζi to produce the generalized Laguerre polynomials for
moments from a gamma NDF where f (x) ∝ xα e−βx and β > 0. The ζi in this case are, for i ≥ 1:
L i+α L i
ζ2i−1 = , ζ2i = . (24)
β β
Note that β is a scaling parameter, while α changes the shape of the NDF and is dimensionless. These
two parameters are fixed given the mean and variance of the NDF. These properties lead to the following
definition.
Definition 2 (Gamma–GQMOM with moments M2n for n ≥ 1 from a NDF defined on R+ ). Let
M12
α= − 1.
M2 M0 − M12
Then α > −1 and the unknown recurrence coefficients are selected as follows. For i > n ≥ 1, let
[M2n ] i+α [M ] i
ζ2i−1 = ζ2n−1 , ζ2i 2n = ζ2n . (25)
n+α n
Thus, for i > n ≥ 1, in eq. (15) the recurrence coefficients are
[M2n ] [M2n ] [M 2n] [M2n ] [M
2n ] [M2n ]
ai = ζ2i + ζ2i+1 , bi = ζ2i−1 ζ2i . (26)
In fig. 2, the weights and abscissae are shown for gamma–PDF moments with α = 10 and β = 1 and three
different values of n. In all cases, the abscissas are positive and do not depend on n. The NDF from EM
with the same zero- and first-order moments corresponds to α = 0 (i.e., an exponential NDF).
Another possibility is to select the additional ζi to produce the Stieltjes–Wigert polynomials for moments
from a lognormal NDF:
(ln(ξ) − µ)2
1
f (ξ) ∝ √ exp − . (27)
ξσ 2π 2σ 2
The ζi in this case are, for i ≥ 1, denoting η = exp(σ 2 /2) > 1 (Madadi-Kandjani and Passalacqua, 2015;
Wilck, 2001):
W
ζ2i−1 = eµ η 4i−3 , W
ζ2i = eµ η 2i−1 (η 2i − 1). (28)
8
100
10-2
10-4
10-6
10-8
10-10
10-12
10-14
10-16
10-18
10-20
100
Figure 3: Lognormal–GQMOM quadrature weights versus abscissae for lognormal moments with η = 1.01, N = 201, and
different values of n.
With η > 1, the recurrence coefficients for i > n ≥ 1 are found from eq. (26) using
2i
[M2n ] 4(i−n) [M2n ] 2(i−n) η −1
ζ2i−1 = η ζ2n−1 , ζ2i =η ζ2n . (30)
η 2n − 1
In fig. 3, the weights and abscissae are shown for three different values of n with lognormal moments and
η = 1.01. Again, the abscissae are positive and do not depend on n, but many have weights below 10−15
due to the long tail of the lognormal NDF.
ζi = pi (1 − pi−1 ) (31)
for 1 ≤ i ≤ 2n with p0 = 0. For this purpose, the Chebyshev algorithm with input M2n can be easily
modified to compute the pi from the ζi found from eq. (23). Here, the choice for pi produces the Jacobi
polynomials for moments from a beta NDF where f (x) ∝ xβ (1 − x)α , but also that the quadrature nodes
are always in the interval (0, 1), which is the condition for the strict realizability. The canonical moments
in this case are, for i ≥ 1:
β+i i
pJ2i−1 = , pJ2i = . (32)
2i + α + β 2i + 1 + α + β
The two parameters α and β are fixed given the mean and variance of the NDF. These properties lead to
the following definition.
Definition 4 (Beta–GQMOM with moments M2n for n ≥ 1 from a NDF defined on (0, 1)). The
unknown recurrence coefficients are selected as follows. Let
1 − p1 − 2p2 + p1 p2 p1 − p2 − p1 p2
α= , β= (33)
p2 p2
9
0.5 0.04
0.45
0.035
0.4
0.03
0.35
0.025
0.3
0.25 0.02
0.2
0.015
0.15
0.01
0.1
0.005
0.05
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Figure 4: Beta–GQMOM quadrature weights versus abscissae for beta-NDF moments in (0, 1) with N = 51 and different values
of n.
In fig. 4, the weights and abscissae are shown for moments from a beta NDF for different values of α, β > −1.
As can be observed, the abscissas remain bounded in (0, 1) and the quadrature points are the same for all
n. By a linear change of variables, beta–GQMOM on (0, 1) can be shifted to define a Gauss quadrature for
any finite interval (a, b).
10
3.3. Computation of weights and abscissas with GQMOM
As done with QMOM (McGraw, 1997; Wheeler, 1974), the N weights and N abscissae are found from
the eigenvectors and eigenvalues, respectively, of the following Jacobi matrix:
√
√a0 b1 √
b1 a1 b2
JN =
.. .. .. .
(36)
. . .
p p
bN −2 p aN −2 bN −1
bN −1 aN −1
Using the definitions in eq. (21) and eq. (22), it is straightforward to modify the Chebyshev algorithm in Fox
and Laurent (2022) to construct JN given the moments M2n , and then to find the weights and abscissae.
By construction, the GQMOM quadrature satisfies eq. (14). The source terms in the moment equations
eq. (3) are closed using eq. (13). Unlike with QMOM, with GQMOM the value of N can be very large
as long as n is not too large (n ≤ 10). Notwithstanding, increasing N with fixed n does not increase our
knowledge of the unknown NDF governed by the PBE. Thus, as with the secondary quadrature in EQMOM
(Yuan et al., 2012), increasing N is only needed to reduce the quadrature error on S k . In other words, for
fixed n, the value of N can be increased until S k satisfies a convergence criterion for all k.
In this section we provide numerical examples for problems that are difficult to treat using QMOM. First,
to illustrate the differences between QMOM and GQMOM, we show examples of NDF composed of three
modes. Recall that with n = 1, GQMOM corresponds to a known NDF (e.g., Gaussian) while for QMOM,
it corresponds to a Dirac delta function. By increasing n, it is possible to approximate a multi-modal NDF.
It is important to understand that QMOM with M2n−1 represents the NDF by the sum of n Dirac delta
functions, and, given the additional moment M2n , GQMOM provides a continuous NDF. Heuristically, for
fixed M2n−1 GQMOM spreads out the n Dirac delta functions found with QMOM as M2n increases from
its minimum value. Thus, with sufficiently large M2n , the GQMOM NDF will become mono-modal, losing
all resemblance with the QMOM NDF.
Generally speaking, if the time evolution of the moments of a NDF are adequately captured with QMOM,
then GQMOM will provide equivalent accuracy. Thus, in the numerical examples starting in section 4.2, we
focus on systems for which QMOM is known to have difficulties. While such systems are relatively rare in
real applications, we shall see that GQMOM offers a viable alternative to QMOM for such cases.
n
0.3 0.3 3
6
9
0.25 0.25
0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0 0
-6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6
Figure 5: Trimodal Gaussian NDF on R with n = 3, 6, 9. Left: Gaussian-GQMOM with N = 101. Right: QMOM. The
quadrature points are connected by lines to ease interpretation.
4.5
3.5
2.5
1.5
0.5
0
0 0.2 0.4 0.6 0.8 1
Figure 6: Trimodal NDF on (0, 1) with n = 6 and different N . Weights are scaled to sum to N .
points for each peak, while N = 60 covers the full domain with roughly equal-spaced points. In all cases,
because GQMOM reproduces the same moments as QMOM, using N > n does not produce a poorer quality
quadrature for fixed n. In practice, a NDF with up to five modes can be approximated with GQMOM to
reasonable accuracy.
Case β (ξ, ξ 0 ) a
((ξ) b (ξ|ξ 0 ) Mk (t = 0)
0 ξ=1
5 1 1, table 2 Mk = 1, k = 0, . . . , 5
0.02 ξ > 1
M0 = 1
M1 = 1.13
M2 = 1.294
(
0 2 2 02 0 ξ=1
8 (ξ + ξ ) |ξ − ξ | 2, table 2 M3 = 1.5
0.01ξ 6 ξ>1
M4 = 1.760
M5 = 2.087
M6 = 2.513
Z ∞
D̄a (t, ξ) = f (t, ξ) β(ξ, ξ 0 )f (t, ξ 0 ) dξ 0 , (39)
0
Z ∞
B̄ b (t, ξ) = a(ξ 0 )b(ξ|ξ 0 )f (t, ξ 0 ) dξ 0 , (40)
ξ
The application of eq. (2) to both sides of eq. (37) leads to the evolution equation for the moment Mk of
the NDF:
dMk (t)
= B̄ka (t) − D̄ka (t) + B̄kb (t) − D̄kb (t). (42)
dt
For consistency with prior work, in this example we use the moment vector M5 . The source terms for the
moments are defined as follows:
1 ∞
Z Z ∞
B̄ka (t) = f (t, ξ 0 ) β(ξ, ξ 0 )(ξ 3 + ξ 03 )k/3 f (t, ξ) dξ dξ 0 ,
2 0 0
Z ∞ Z ∞
a
D̄k (t) = k
ξ f (t, ξ) β(ξ, ξ 0 )f (t, ξ 0 ) dξ 0 dξ,
0 0
Z ∞ Z ∞ (43)
b k 0 0 0 0
B̄k (t) = ξ a(ξ )b(ξ|ξ )f (t, ξ ) dξ dξ,
Z0 ∞ 0
The integrals in eq. (43) are approximated as shown in eq. (4), when using GQMOM, by replacing n with N
to use all the additional nodes. The EQMOM integration strategy for eq. (42) is summarized in Appendix
A, and further details can be found in Madadi-Kandjani and Passalacqua (2015); Yuan et al. (2012). Table 1
summarizes the closure models used in the two test cases considered here, while the daughter distribution
functions b (ξ|ξ 0 ) are listed in table 2.
The time evolution of the volume-average diameter d43 = M4 /M3 is reported in fig. 7, where results are
shown for gamma–GQMOM, as well as for the lognormal–EQMOM (Madadi-Kandjani and Passalacqua,
2015), and the exact solution (Vanni, 2000). We consider the same cases to investigate the convergence of
13
Table 2: Daughter distribution functions and their moment transforms.
5 5
Lognormal EQMOM (Madadi and Passalacqua, 2015)
4.5
4.5
Rigorous solution (Vanni, 2000)
4
4 GQMOM
3.5
3.5 3
3 2.5
d43
d43
2
2.5
1.5
2
Lognormal EQMOM (Madadi and Passalacqua, 2015) 1
1.5 Rigorous solution (Vanni, 2000) 0.5
GQMOM
1 0
0 50 100 150 200 0 2 4 6 8 10
t (s) t (s)
Figure 7: Time evolution of d43 obtained with gamma–GQMOM, lognormal–EQMOM Madadi-Kandjani and Passalacqua
(2015), and the exact solution Vanni (2000).
4.5
3.5
3
d43 (μm)
2.5
2 N=4
N=5
1.5
N = 10
1 N = 50
N = 100
0.5
Validation data - Rigorous solution of Vanni (2000)
0
0 50 100 150 200
t (s)
14
3.5
2.5
d43 (μm)
2
N=4
1.5 N=5
N = 10
1 N = 50
N = 100
N = 150
0.5
QMOM (3 nodes)
Validation data - Rigorous solution of Vanni (2000)
0
0 2 4 6 8 10
t (s)
gamma–GQMOM with the number of nodes, while maintaining the number of QMOM quadrature nodes
constant and equal to n = 3 (6 moments), and increasing the number N of gamma–GQMOM nodes. We
consider N = 4, 5, 10, 50, 100 for Case 5 and N = 4, 5, 10, 50, 100, 150 for Case 8.
From the time evolution of d43 reported in fig. 8, no variation can be visually observed in the solutions
obtained with gamma–GQMOM using different N . A more interesting behavior is observed in Case 8, for
which the time evolution of d43 is shown in fig. 9. In this case, QMOM with n = 3 is insufficient to capture
the asymptotic behavior. Adding another gamma–GQMOM node for the same moments leads to correctly
predict the trend but not the asymptotic value. However, the gamma–GQMOM prediction appears to
converge for N > 50. Specifically, for the cases run with N = 100 and 150 the percentage difference between
the asymptotic values of d43 is 0.0462%.
∂f (t, ξ)
= ϕδξ0 (ξ) − ∂ξ (Gf (t, ξ)) − αf (t, ξ)I[ξ1 ,+∞) (ξ). (44)
∂t
where the identity function I[a,+∞) (x) is zero for x < a and 1 elsewhere. Let us remark that f t, ξξ−ξ
1
0
−ξ 0
is
the solution of a similar PBE with scaled parameters, in such a way that we can choose ξ0 = 0 and ξ1 = 1
without loss of generality. Finally, the values of the parameters used here are ϕ = 1, G = 2 and α = 10, in
this non-dimensional context. The main difficulty of this PBE lies in the filtration term, which is localized
in phase space so that it only affects particles with size ξ ≥ ξ1 . Indeed, quadrature-based methods usually
give very good results for the growth term, even with a non-constant rate G(ξ), and the nucleation term is
already closed in the moment equations.
The analytical solution of eq. (44), with the zero initial condition f (0, ξ) = 0, is
0 if ξ < ξ0
ϕ
f (t, ξ) = 1 if ξ0 ≤ ξ ≤ min(ξ1 , ξ0 + tG) , (45)
G α
exp − G (ξ − ξ1 ) if ξ1 ≤ ξ ≤ ξ0 + tG
15
0.7 0.5
0.45
0.6
0.4
0.5 0.35
0.4 0.3
M0
M1
0.25
0.3 0.2
0.2 exact 0.15 exact
QMOM-Radau 0.1 QMOM-Radau
0.1 GQMOM-Radau GQMOM-Radau
GQMOM 0.05 GQMOM
EQMOM EQMOM
0 0
0 1 2 3 4 5 0 1 2 3 4 5
time time
0.5 0.6
0.45
0.5
0.4
0.35
0.4
0.3
M2
M3
0.25 0.3
0.2
0.2
0.15 exact exact
0.1 QMOM-Radau QMOM-Radau
GQMOM-Radau 0.1 GQMOM-Radau
0.05 GQMOM GQMOM
EQMOM EQMOM
0 0
0 1 2 3 4 5 0 1 2 3 4 5
time time
0.7
0.6
0.5
0.4
M4
0.3
0.2 exact
QMOM-Radau
0.1 GQMOM-Radau
GQMOM
EQMOM
0
0 1 2 3 4 5
time
Figure 10: Test case with nucleation, growth and filtration: moments of order 0, 1, 2, 3 and 4 for the analytical solution,
QMOM–Radau with five moments (i.e., M4 ), gamma–GQMOM and gamma–GQMOM–Radau with five moments and N = 20
and gamma–EQMOM with 10 secondary quadrature points.
Using GQMOM with M2n , the equations for the moments are written, for k = 0, 1, . . . , 2n, as
Z +∞
dMk (t) k
= ϕξ0 + kGMk−1 (t) − α ξ k dµ[M2n ] . (47)
dt ξ1
For their resolution, an operator splitting is introduced. During one time step ∆t, the growth term is
first solved, followed by the resolution of the others terms. Each operator is solved analytically, using
the underlying representation of the moments as a sum of Dirac delta functions. Then, for the growth
operator, each abscissa of the quadrature is increased by G∆t. For the nucleation and filtration operators,
the weights corresponding to abscissa greater than ξ1 are multiplied by exp(−a∆t) and ϕξ0k ∆t is added to
the corresponding moment of order k, for k = 0 . . . , 2n. Moreover, a small enough time step (5×10−6 ) is
chosen to ensure time convergence.
The simulations are done using five moments M4 and N = 20 abscissas with gamma–GQMOM and
gamma–GQMOM–Radau. Here, Radau quadrature refers to a Gaussian quadrature where one abscissa is
16
0.7 0.5
0.45
0.6
0.4
0.5 0.35
0.4 0.3
M0
M1
0.25
0.3 0.2
exact exact
0.2 QMOM 0.15 QMOM
GQMOM-Radau 0.1 GQMOM-Radau
0.1 GQMOM GQMOM
EQ alpha=2 0.05 EQ alpha=2
EQ alpha=4 EQ alpha=4
0 0
0 1 2 3 4 5 0 1 2 3 4 5
time time
0.45 0.5
0.4 0.45
0.35 0.4
0.35
0.3
0.3
0.25
M2
M3
0.25
0.2
0.2
0.15 exact exact
QMOM 0.15 QMOM
0.1 GQMOM-Radau 0.1 GQMOM-Radau
GQMOM GQMOM
0.05 EQ alpha=2 0.05 EQ alpha=2
EQ alpha=4 EQ alpha=4
0 0
0 1 2 3 4 5 0 1 2 3 4 5
time time
Figure 11: Test case with nucleation, growth and filtration: moments of order 0, 1, 2 and 3 for the analytical solution, QMOM
with four moments (i.e., M3 ), gamma–GQMOM and gamma–GQMOM–Radau with four moments and entropic quadrature
(EQ) with multiplicative factor α = 2 and α = 4.
fixed (Gautschi and Li, 1991). These results are compared with the analytical solution in fig. 10. As seen
from these plots, gamma–GQMOM–Radau seems to be well adapted to this test case: compared to gamma–
GQMOM, an abscissa is fixed at ξ = 0, which is the nuclei size. And then, gamma–GQMOM–Radau gives
better results than gamma–GQMOM, even if gamma–GQMOM also has good results for a larger number of
abscissas (not shown here). These methods are also compared to QMOM–Radau, whose results for M0 are
quite good, but the error increases with the order of the moments. Gamma–EQMOM is also tested with
ten secondary quadrature points, so that the cost is quite similar to gamma–GQMOM. However, gamma–
EQMOM gives the worst results, even if these results can be improved by using a much larger number of
secondary quadrature points (not shown here).
Finally, to compare GQMOM with another method from the literature, entropic quadrature (EQ) de-
veloped by Böhmer and Torrilhon (2020), simulations using four moments M3 are done. In fig. 11 the
results are plotted with gamma–GQMOM and gamma–GQMOM–Radau using N = 20 quadrature points,
QMOM, and EQ with multiplicative factor α = 2 and α = 4 and the smallest abscissa fixed at ξ = 0. The
best results are obtained with gamma–GQMOM–Radau even if the moments are slightly overestimated,
whereas the other methods show an oscillatory behavior. In any case, it is very interesting to observe how
using Radau quadrature in the presence of nucleation significantly improves the predictions of GQMOM,
even with a relatively small number of moments. Also, comparing fig. 11 with fig. 10, we see that using
gamma–GQMOM–Radau with M4 is significantly more accurate than with M3 . This is consistent with
observations made with other systems, i.e., moment vectors terminating with an even-order moment usually
provide more accurate results.
17
2 0.9
n=2 n=2
1.8 n=3 n=3
n=4 0.85 n=4
1.6 n=5 n=5
sqrt(M3M1/M22-1)
n=7 n=7
1.4 n=10 0.8 n=10
n=15 n=15
M2/M1
1.2
0.75
1
0.8 0.7
0.6
0.65
0.4
0.2 0.6
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
time time
Figure 12: Fragmentation, Case 1: mean (left) and variance (right) obtained for simulations with QMOM with a number of
abscissas n equal to 2, 3, 4, 5, 7, 10, 15.
with the initial condition f (0, v) = f 0 (v). Moreover, the simple case g(v) = v suffices to illustrate the
difficulties. Using GQMOM with M2n−1 , the equations for the moments are written, for k = 0 . . . , 2n − 1,
as
dMk (t)
= (−1 + 21−k )Mk+1 (t). (49)
dt
Let us remark that, due to the simple choice for g, only M2n has to be closed. This set of equations is solved
using an adaptive time-step algorithm (Nguyen et al., 2016) based on embedded SSP explicit Runge–Kutta
methods, with a time-step selection designed both to control the error and to ensure the realizability of the
moment set.
For the initial distribution, two cases are considered:
18
2 0.71
QMOM, n=15 QMOM, n=15
1.8 n=2, N=3 n=2, N=3
n=3, N=4 0.7 n=3, N=4
1.6 n=4, N=5 n=4, N=5
sqrt(M3M1/M22-1)
n=5, N=6 0.69 n=5, N=6
1.4
0.68
M2/M1
1.2
1 0.67
0.8
0.66
0.6
0.65
0.4
0.2 0.64
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
time time
Figure 13: Fragmentation, Case 1: reference solution using QMOM and mean (left) and variance (right) obtained for simulations
with gamma–GQMOM with a number of moments 2n equal to 4, 6, 8, 10; and a number of quadrature points equal to n + 1.
1 2
reference reference
0.9 QMOM 1.8 QMOM
Gamma GQMOM Gamma GQMOM
0.8 Lognormal GQMOM 1.6 Lognormal GQMOM
sqrt(M3M1/M22-1)
0.7 1.4
M2/M1
0.6 1.2
0.5 1
0.4 0.8
0.3 0.6
0.2 0.4
0.1 0.2
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
time time
Figure 14: Fragmentation, Case 2: reference solution from Peterson et al. (2022) and converged mean (left) and variance (right)
obtained for simulations with QMOM, gamma–GQMOM, and lognormal–GQMOM.
In this work, we have generalized the widely used moment closure QMOM for solving moment systems
derived from a 1-D PBE to allow for an arbitrarily large number N of Gauss quadrature nodes. Like QMOM
does for M2n−1 , GQMOM exactly reproduces the input moment vector M2n where 2n is the order of the
highest-order known moment. Unlike QMOM, GQMOM closes the higher-order moments such that the
moment vector M2N is (almost always) in the interior of moment space, i.e., the moments correspond to a
continuous NDF unless M2n is on the boundary of moment space. When the latter occurs, the GQMOM
quadrature is the same as the one found with QMOM. This is because, by construction, unless bn = 0,
GQMOM will never produce a set of recurrence coefficients on the boundary of moment space. As we
have shown through examples, the principal advantage of using GQMOM versus QMOM is the ability to
increase the number of quadrature points from n to N without solving for more moments. This allows for
a more accurate evaluation of the moment source terms at nearly the same computation cost. Nonetheless,
as shown in the examples, the overall accuracy of the moment method is controlled principally by n (i.e.,
by the number of solved moments), and, hence, a much larger N is only needed for “difficult” source terms.
In comparison to other moment closures based on a continuous NDF (e.g., EQMOM, EM, EQ), GQ-
MOM is much easier to implement because it only requires a straightforward modification of the Chebyshev
algorithm to compute the additional recurrence coefficients (i.e., ai , bi for i = n, n + 1, . . . , N ). Furthermore,
because this process results in a Gauss quadrature on the support B, we are guaranteed that the abscissae,
being the roots of a monic orthogonal polynomial, lie inside B and are distinct. This is not the case, for
example, with EQMOM where the abscissae from the secondary quadrature can overlap, which causes severe
problems, for example, when one attempts to find conditional moments (Cheng et al., 2010; Yuan and Fox,
2011). Moreover, the weights of the Gauss quadrature from GQMOM will be positive as long as M2n lies
in the interior of moment space; otherwise, some weights will be null.
In general, moment closures that reside in the interior of moment space are preferable because they
can tolerate small numerical errors without becoming non-realizable. For example, for a PBE that includes
19
spatial transport with a known velocity (e.g., the gas velocity for an aerosol), constructing realizable finite-
volume schemes is challenging (Wright, 2007), even if some realizable second-order schemes were developed
(Laurent and Nguyen, 2017; Marchisio and Fox, 2013; Passalacqua et al., 2020; Shiea et al., 2020; Vikas et al.,
2011). In this respect, GQMOM may allow for use of higher-order spatial reconstruction of the moment
vector for cases where QMOM is limited to low order. The same issue is faced when solving the moment
source terms numerically (Nguyen et al., 2016), so that, generally speaking, we expect that GQMOM will
generate more robust, and more accurate, numerical solvers for the multi-scale, multi-physics codes used in
real-world applications (Bryngelson et al., 2020; Heylmun et al., 2021, 2019; Ilgun et al., 2021; Passalacqua
et al., 2018; Wick et al., 2017).
Many real-world applications are described by a NDF with more than one internal variable (e.g., for
aerosols, the droplet volume, temperature, composition, etc.). It is well known that the numerical solution
of a multi-dimensional PBE, even for spatially homogeneous cases, is very challenging due to the high
dimension of the phase space. One possible simplification is to condition the internal variables on the variable
with the largest variance (Marchisio and Fox, 2013; Yuan and Fox, 2011). In applications involving droplets
(aerosols), solid particles or bubbles, the mass (or size) of the particle is very often the most important
internal variable. As done in the examples in section 4, GQMOM can be applied with the moments M2n
of the size variable ξ. Then, the conditional mean of another variable φ given ξ is denoted by hφ|ξi, and
is computed using the joint moments hφξ k i for k = 0, 1, . . . , n − 1 by applying the conditional quadrature
method of moments (CQMOM) (Cheng et al., 2010; Yuan and Fox, 2011). When QMOM is replaced by
GQMOM, it will be necessary to approximate hφ|ξi using the same set of joint moments as CQMOM.
However, the straightforward application of CQMOM with N > n abscissae yields an under-determined
linear system. Thus, it is necessary to modify CQMOM to handle N > n abscissae. This can be done
successfully using an interpolation function, and we will describe the resulting algorithm for generalized
CQMOM in a future publication.
The authors declare that they have no known competing interests or personal relationships that could
have appeared to influence the work reported in this paper.
Differently from QMOM and GQMOM, in EQMOM the NDF is approximated as a weighted sum of
non-negative KDF (Yuan et al., 2012):
n
X
f (ξ) ≈ pn (ξ) = wα δσ (ξ, ξα ) (A.1)
α=1
where wα are the weights of each KDF δσ (ξ, ξα ) with standard deviation σ, ξα are the corresponding
abscissae, and n is the number of KDF being used. Based on this approximation, integrals involving
products of the NDF and a function of ξ are approximated as (Madadi-Kandjani and Passalacqua, 2015)
Z ∞ Z ∞ n
X X Nα
n X
g(ξ)pn (ξ) dξ = g(ξ) wα δσ (ξ, ξα ) dξ = wα wαβ g(ξαβ ). (A.2)
0 0 α=1 α=1 β=1
where, for fixed α, wαβ and ξαβ are the Nα Gaussian quadrature nodes corresponding to the KDF (Yuan
et al., 2012). In principle, Nα can be different for each α, but here we let Nα = 2n.
20
Consequently, the evolution equation of the moment Mk in aggregation and breakup processes discussed
in section 4.2 is rewritten as
n Nα n Nα
dMk 1 X X X X
= wα1 wα1 β1 wα2 wα2 β2 (ξα3 1 β1 + ξα3 2 β2 )k/3 βα1 β1 α2 β2
dt 2 α =1 α =1
1 β1 =1 2 β2 =1
n
X Nα
X n
X Nα
X
− wα ξαk 1 β1 wα1 wα1 β1 wα2 wα2 β2 βα1 β1 α2 β2
α1 =1 β1 =1 α2 =1 β2 =1
X Nα
n X Nα
n X
X
+ wα aαβ b̄kαβ wαβ − k
wα ξαβ aαβ wαβ (A.3)
α=1 β=1 α=1 β=1
where b̄kαβ = b̄k (ξαβ ). In the numerical implementation of EQMOM, σ is found from M2n using an iterative
procedure (Pigou et al., 2018; Yuan et al., 2012). Thus, since N ≈ nNα , the principal advantage of GQMOM
versus EQMOM is to eliminate the need for iterations to find σ before evaluating the right-hand side of
eq. (A.3).
References
Böhmer, N., Torrilhon, M., 2020. Entropic quadrature for moment approximations of the Boltzmann–BGK equation. Journal
of Computational Physics 401, 108992. URL: http://www.sciencedirect.com/science/article/pii/S0021999119306977,
doi:https://doi.org/10.1016/j.jcp.2019.108992.
Bryngelson, S.H., Colonius, T., Fox, R.O., 2020. QBMMLib: A library of quadrature-based moment methods. SoftwareX 12,
100615.
Cheng, J.C., Vigil, R.D., Fox, R.O., 2010. A competitive aggregation model for flash nanoprecipitation. J. Colloid Interface
Sci. 351, 330–342.
Dette, H., Studden, W.J., 1997. The Theory of Canonical Moments with Applications in Statistics, Probability, and Analysis.
Wiley Series in Probability and Statistics: Applied Probability and Statistics, John Wiley & Sons Inc., New York.
Fox, R.O., Laurent, F., 2022. Hyperbolic quadrature method of moments for the one-dimensional kinetic equation. SIAM J.
Appl. Math. 82, 750–771.
Gautschi, W., 2004. Orthogonal Polynomials: Computation and Approximation. Oxford University Press, Oxford, UK.
Gautschi, W., Li, S., 1991. Gauss–Radau and Gauss–Labatto quadratures with double end points. Journal of Computational
and Applied Mathematics 34, 343–360.
Grosch, R., Briesen, H., Marquardt, W., Wulkow, M., 2007. Generalization and numerical investigation of QMOM. AIChE
Journal 53, 207–227.
Hamburger, H.L., 1944. Hermitian transformations of deficiency-index (1,1), Jacobian matrices, and undetermined moment
problems. Amer. J. Math. 66, 489–552.
Heylmun, J.C., Fox, R.O., Passalacqua, A., 2021. A quadrature-based moment method for the evolution of the joint size–velocity
number density function of a particle population. Computer Physics Communications 267, 108072.
Heylmun, J.C., Kong, B., Passalacqua, A., Fox, R.O., 2019. A quadrature-based moment method for polydisperse bubbly
flows. Computer Physics Communications 244, 187–204.
Ilgun, A.D., Fox, R.O., Passalacqua, A., 2021. Solution of the first-order conditional moment closure for multiphase reacting
flows using quadrature-based moment methods. Chemical Engineering Journal 405.
Jaynes, E.T., 1957. Information theory and statistical mechanics. The Physical Review 106, 620–630.
Lage, P.L., 2011. On the representation of QMOM as a weighted-residual method – the dual-quadrature method of generalized
moments. Computers and Chemical Engineering 35, 2186–2203.
Laurent, F., Nguyen, T.T., 2017. Realizable second-order finite-volume schemes for the advection of moment sets of the particle
size distribution. Journal of Computational Physics 337, 309–338.
Madadi-Kandjani, E., Passalacqua, A., 2015. An extended quadrature-based moment method with log-normal kernel density
functions. Chemical Engineering Science 131, 323–339.
Marchisio, D.L., Fox, R.O., 2013. Computational Models for Polydisperse Particulate and Multiphase Systems. Cambridge
University Press, Cambridge, UK.
Marchisio, D.L., Pikturna, J.T., Fox, R.O., Vigil, R.D., 2003. Quadrature method of moments for population-balance equations.
AIChE Journal 49, 1266–1276.
McGraw, R., 1997. Description of aerosol dynamics by the quadrature method of moments. Aerosol Science and Technology
27, 255–265.
Mead, L.R., Papanicolaou, N., 1984. Maximum entropy in the problem of moments. J. Math. Phys. 25, 2404–2417.
Nguyen, T.T., Laurent, F., Fox, R.O., Massot, M., 2016. Solution of population balance equations in applications with fine
particles: mathematical modeling and numerical schemes. Journal of Computational Physics 325, 129–156.
21
Passalacqua, A., Laurent, F., Fox, R.O., 2020. A second-order realizable scheme for moment advection on unstructured grids.
Computer Physics Communications 248, 106993.
Passalacqua, A., Laurent, F., Madadi-Kandjani, E., Heylmun, J.C., Fox, R.O., 2018. An open-source quadrature-based
population balance solver for OpenFOAM. Chemical Engineering Science 176, 306–318.
Peterson, J.D., Bagkeris, I., Michael, V., 2022. A new framework for numerical modeling of population balance equations:
Solving for the inverse cumulative distribution function. Chemical Engineering Science 259, 117781.
Pigou, M., Morchain, J., Fede, P., Penet, M.I., Laronze, G., 2018. New developments of the extended quadrature method of
moments to solve population balance equations. Journal of Computational Physics 365, 243–268.
Randolph, A.D., Larson, M.A., 1988. Theory of Particulate Processes: Analysis and Techniques of Continuous Crystallization.
Academic Press.
Schmüdgen, K., 2017. The Moment Problem. volume 277 of Graduate Texts in Mathematics. Springer, Cham.
Shiea, M., Buffo, A., Vanni, M., Marchisio, D.L., 2020. A novel finite-volume TVD scheme to overcome non-realizability
problem in quadrature-based moment methods. Journal of Computational Physics 409, 109337.
Vanni, M., 2000. Approximate population balance equations for aggregation–breakage processes. J. Colloid Interface Sci. 221,
143–160.
Vikas, V., Wang, Z.J., Passalacqua, A., Fox, R.O., 2011. Realizable high-order finite-volume schemes for quadrature-based
moment methods. J. Comput. Phys. 230, 5328–5352.
Wheeler, J.C., 1974. Modified moments and Gaussian quadratures. Rocky Mt. J. Math. 4, 287–296.
Wick, A., Nguyen, T.T., Laurent, F., Fox, R.O., Pitsch, H., 2017. Modeling soot oxidation with the extended quadrature
method of moments. Proceedings of the Combustion Institute 36, 789–797.
Wilck, M., 2001. A general approximation method for solving integrals containing a lognormal weighting function. J. Aerosol
Sci. 32, 1111–1116.
Wright, D.L., 2007. Numerical advection of moments of the particle size distribution in Eulerian models. J. Aerosol Sci. 38,
352–369.
Yuan, C., Fox, R.O., 2011. Conditional quadrature method of moments for kinetic equations. J. Comput. Phys. 230, 8216–8246.
Yuan, C., Laurent, F., Fox, R.O., 2012. An extended quadrature method of moments for population balance equations. J.
Aerosol Sci. 51, 1–23.
22