You are on page 1of 61

CHAPTER TWO

LBM for 2D and 3D chemical


reactors
Giacomo Falcucci*
Department of Enterprise Engineering “Mario Lucertini”, University of Rome “Tor Vergata”, Rome, Italy
John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA,
United States
*Corresponding author: e-mail addresses: giacomo.falcucci@uniroma2.it; falcucci@g.harvard.edu

Contents
1. Introduction 82
2. Prologue: The sparking question 84
3. First act—Training: Developing the numerical model 85
4. Second act—Fighting: Validating the numerical model 89
5. Third act—The good win: Answering the question 95
6. “First steps in a larger world” : np-Au hollow spheres 106
6.1 Results and discussion 108
7. Microbial fuel cells (MFC) 113
7.1 Results and discussion 115
7.2 Activity of bacteria strains 116
7.3 Ion motion within MFC reactor 117
7.4 Polarization and power curves 118
8. 3D Microbial fuel cells 122
9. Selective catalytic reactors (SCR) for diesel engine applications 127
9.1 Results and discussion 128
10. 3D SCR systems 130
11. Conclusions 135
Acknowledgments 136
References 137

Abstract
In this chapter, the main aspects related to the analysis of 2D and 3D chemical reactors
by means of the lattice Boltzmann method will be presented and discussed.
The numerical schematization of nano- and micro-scale catalytic process in porous
substrates represents a serious challenge for fluid dynamics, as the local values of
Knudsen number cannot be considered infinitesimal, thus posing a serious threat to
the reliability of the simulation results.
In the first sections, a novel, ad hoc developed approach to account for heteroge-
neous catalysis in nano-porous active media will be presented.

Advances in Chemical Engineering, Volume 55 # 2020 Elsevier Inc. 81


ISSN 0065-2377 All rights reserved.
https://doi.org/10.1016/bs.ache.2020.04.004
82 Giacomo Falcucci

Such a numerical scheme is grounded on two pillars: the effective, yet computa-
tionally light, reconstruction of the nano-porous media and the kinetic foundations
of the LBM that allow the accurate implementation of active boundary conditions at
mesoscopic physical scales.
Results for various applications of scientific and technological interest will be illus-
trated and discussed, highlighting the versatility and reliability of the proposed
approach, even for fluid flows marginally in the Knudsen regime. The comparison with
experimental and numerical works from the literature highlights that the proposed
methodology can provide a useful tool in the design of novel catalysts and in answering
open questions rising from experimental activities.

1. Introduction
Catalytic processes lay at the heart of a large number of phenomena
and production techniques that affect everyday life, from fuel synthesis to
commodity chemicals, from pollutant emission and control to fuel cells.
Such a broad range of applications make catalysis pivotal to reach the ambi-
tious goals of sustainable and green development for the future years (Duan
et al., 2015; Weaver et al., 2017). To this aim, it is crucial to improve
traditional catalytic processes and to develop novel catalysts: among the
new-generation materials, nano-structured precious elements (such as gold)
have shown remarkable properties (Wittstock et al., 2010). It is well
addressed in the literature, in fact, the impact that the progress in nanotech-
nology has achieved on the efficiency and performances of catalytic pro-
cesses (Bell, 2003): however, to predict the behavior of such complex
nano-materials, it is crucial to develop novel, multiscale methodologies to
account for the molecular interactions at reactant–substrate interface, where
the catalytic reactions take place (Falcucci et al., 2016; Kickelbick, 2007;
Krastev and Falcucci, 2018; Xu et al., 2011). The fundamental, quantum-
mechanical interactions, for instance, would require the accurate description
of electron-transfer phenomena (i.e., bond-breaking events), but such pro-
cesses take place on timescales of femtoseconds (1015 s), thus making com-
putationally impossible their tracking on experimental physical scales
(Bernaschi et al., 2010; Succi et al., 2002, 2019). This is the reason why a
number of scientific works has flourished, in recent years, to address the
need of multiscale techniques, connecting the atomic level to the meso-
and macroscopic scales (Falcucci et al., 2011a, b; Keil, 2012; Krastev
et al., 2019; Ma et al., 2015; MacMinn et al., 2015; Salciccioli et al.,
2011; Succi, 2002). In this context, the lattice Boltzmann method (LBM)
LBM for 2D and 3D chemical reactors 83

(Benzi et al., 1992b; Succi, 2001) has been successfully applied to numerous
complex phenomena of scientific and technological interest (Cali et al., 1992;
Chen et al., 1991; Di Ilio et al., 2018, 2019; Falcucci et al., 2007, 2013;
Krastev and Falcucci, 2019; Machado, 2012; Montessori et al., 2015;
Succi, 2001, 2002; Zhou et al., 2015), providing a reliable yet versatile
num-erical framework for multiscale catalysis simulation, as well (Falcucci
et al., 2016, 2017; Succi et al., 2019).
A major issue in applying LBM to the evaluation of nano- and micro-
scale chemical reactors lies in the characteristic physical scales involved in
fluid–solid interface phenomena, where the local value of the Knudsen
number (Kn) can approximate and even overcome the threshold value of
Kn ¼ 1 (Montessori et al., 2015) thus making the numerical predictions
of LBM inconsistent with hydrodynamic regimes (Kr€ uger et al., 2017;
Succi, 2001). Such a problem rises in traditional experimental apparata for
catalytic synthesis, such as the continuously stirred tank reactors (CSTRs)
and the plug-flow-reactor (PFRs)—used for the nonsteady temporal analysis
of products (TAP) (Feres et al., 2009b; Marin et al., 2019; Shekhtman et al.,
1999, 2003)—which work at high-vacuum conditions, typically 103 
104 torr. The analytical characterization of such operating conditions is still
an open issue and, consequently, efficient numerical methods to investigate
the nonhigh-vacuum regimes are mandatory (Falcucci et al., 2017).
Through the sections of this chapter, the main characteristics of 2D and
3D numerical approaches based on LBM to dissect the aspects related to
multicomponent, reacting flows within complex, nano- and micro-scale
porous media will be presented. The digital reconstruction of the nano-
and micro-scale active porous media will be illustrated; the validation of
the proposed methodology against experimental cases in the literature will
be presented. Moreover, we will assess the issues (and solutions) in dealing
with high Kn numbers, which are typical for reactive flows through com-
plex porous geometries, without using higher-order methods. In the liter-
ature, such regimes are generally attacked by means of high-order schemes
(Montessori et al., 2015), characterized by the drawback of a heavy compu-
tational burden, which, however, remains lighter compared to Monte Carlo
simulations (Bird and Brady, 1994; Di Staso et al., 2016): even so, the D2Q9
and D3Q19 schemes that will be portrayed in the present chapter remain by
far less demanding and onerous, mainly for the implementation of boundary
conditions (Falcucci et al., 2017; Montessori et al., 2017).
In the following the path that has led to the implementation of 2D (first)
and 3D (lately) approaches will be illustrated: these approaches will be
84 Giacomo Falcucci

applied to experimental phenomena of technological and scientific rele-


vance, providing results in good agreement with physical measurements.
Such encouraging results highlight that our LB model may provide a viable
computational tool for the optimal design of future nano-porous catalyst
substrates.

2. Prologue: The sparking question


Discussing with our friends and colleagues from the Chemistry and
Physics Departments at Harvard University, they turned us a question that
was puzzling them, at that moment. They were conducting experiments on
nano-porous gold catalytic ingots, for methanol oxidation reaction: such
samples were prepared by dealloying bulk Ag70Au30 alloy discs (Biener
et al., 2006) (200–300 μm thickness, 5 mm diameter) in 70% nitric acid
(Alfa Aesar) for 48 h. The sample was then washed with deionized water
and dried in static air, and then it was installed into a quartz tube micro-
reactor to perform the catalytic tests. Before the experiments, the catalyst
was activated according to the procedure described in Gordon et al.
(2003) and R€ ohe et al. (2013). A schematic representation of the reactor sys-
tem is shown in Fig. 1.
After few hours of operation at 150°C in a continuous flow of reac-
tant (6.5% MeOH–20% O2–73.5% He) with a volumetric flow rate of

He, Co, O2 Pressure


gauge

Syringe
pump

Quartz substrate and


Np-Au ingot
Bypass
Furnace

To GC Heated lines
Fig. 1 Schematic representation of the experimental reactor with np-Au catalytic ingot
for methanol oxidation reaction. The ingot is placed into an inert quartz porous sub-
strate within the gas line. The porosity of the support is several orders of magnitude
that of the ingot, thus the quartz substrate is neglected, in the following.
LBM for 2D and 3D chemical reactors 85

Fig. 2 Evolution of the morphology of np-Au ingot during operation. The panel reports
the SEM images of (A) the surface facing the reactant stream, and (B) the rear surface of
the porous catalytic slab: the change in the microstructure of np-Au is apparent (Falcucci
et al., 2016).

50 mL/min, scanning electron microscopy (SEM) analysis performed on


the ingot sample by means of a Zeiss Supra 55VP instrument revealed that
the microstructure of the np-Au ingot undertook a dramatic change in its
side facing the stream of reactants, compared to the opposite side, which
remained almost unchanged, as reported in Fig. 2.
So, our friends and colleagues left us with this puzzling, compelling ques-
tion: “Why such a superficial restructuring takes place?”

3. First act—Training: Developing the numerical model


Given the layout of the catalytic ingot used in the experiments, we first
focused on a 2D representation of the porous media. Due to the complex
morphology of the ingot micro- and nano-structure and the very small char-
acteristic physical scales, the lattice Boltzmann method provides an ideal tool
for this investigation (Falcucci et al., 2009, 2011b; Krastev and Falcucci,
2018; Kr€ uger et al., 2017; Succi, 2002, 2008).
This method, in fact, yields a robust and versatile numerical algorithm for
the analysis of complex fluid phenomena in the kinetic framework of statis-
tical mechanics, according to which the evolution in time at microscopic
scales is described by the probability of finding a particle (a molecule, a family
of molecules or, more generally, a quasiparticle), at the position x with molec-
ular velocity c, at given time t.
In our representation, we adopt a standard implementation of the LBM
equation for a multicomponent system,
86 Giacomo Falcucci

f αi ðx + ci , t + 1Þ  f αi ðx, tÞ ¼ ω½ f i ðx, tÞ  f αi ðx, tÞ,


eq,α
(1)

where f αi ðx, tÞ is the probability density function of finding a particle of spe-


cies α at site x at time t, streaming along the i-th lattice direction defined by
the discrete speeds ci, with i ¼ 0, …, b. For the sake of simplicity, we fix the
lattice time step Δt ¼ 1; index α runs over the involved chemical species: the
model is highly flexible, as there is no constraint on the number of chemical
species that can be taken into account. For our first attempt in disentangling
the catalytic processes inside the np-Au ingot, we chose to focus on a very
simple, ideal reaction (Falcucci et al., 2016),

R!P (2)
thus, three species have been considered: an inert carrier (C), a reactant (R),
and a product (P). It is worth stressing that such a simple, ideal chemical con-
version does not imply any loss of generality, as the model can be fully cus-
tomized and even complex reaction mechanisms can be included.
Eq. (1) describes the interaction between quasiparticles as a sequence of
streaming (left-hand side) and collision (right-hand side) events between
the fluid molecules in the bulk and fluid–solid molecules at the boundaries.
The collision event accounts for the characteristic relaxation time τ ¼ 1/ω,
eq,α
toward the local equilibrium f i ðx, tÞ, expressed as a low-Mach, second-
order expansion of the local Maxwellian (Succi, 2001),

eq, α ci  uðx,tÞ
fi ðx,tÞ ¼ wi ρα ðx, tÞ 1 +
cs2
 (3)
ðci  uðx, tÞÞ2 juðx, tÞj2
+  :
2 cs4 2 cs2

In a standard simulation of a continuum flow regime, the collisions


between fluid particles are orders of magnitude more frequent than fluid–
solid interactions. In our case, since we deal with gas diffusion through a
nano-porous solid, we face a mean free path between gas-gas collisions that
becomes comparable to the (average) pore characteristic dimension: thus, it
is fair to assume that the two interactions take place with approximately the
same frequency.
It is well known that such a finite-Knudsen regime can deeply mine the
reliability of LBM predictions (Falcucci et al., 2009; Kr€ uger et al., 2017;
Succi, 2001) but recent developments have shown that LBM Method can
be reliably employed even for this kind of simulations, once geared with
LBM for 2D and 3D chemical reactors 87

proper kinetic boundary conditions to correctly outline fluid-wall mass/


momentum transfer (Ansumali and Karlin, 2002; Montessori et al., 2015;
Niu et al., 2007; Succi, 2002). As a further step ahead, in our model, such
kinetic boundary conditions must be boosted with a catalytic plug-in at
solid–fluid interface: thus, a proper mechanism has to be calibrated to account
for the local exchange of populations between R and P species, in the pres-
ence of the solid walls of the pore. In the case of heterogeneous catalysis,
such an operation takes places when reactant gas populations hit the surface
of the porous catalyst, thus we assume that no reaction takes place in the fluid
region; moreover, the chemical conversion time scale is typically faster than
the characteristic time of the molecular collisions in the bulk (Falcucci
et al., 2016; Montessori et al., 2016). Upon colliding with a solid site, in fact,
the reactants R partially stick to the wall with probability pRS (see Fig. 3); the
sticked fraction reacts with probability p, developing products P at the
pore surface. Then, the new species reenters the pore fluid region with prob-
ability ð1  pPS Þ , where pPS is the sticking probability of the products.
Nonconverted reactant molecules return into the fluid domain pore with
probability ð1  pRS Þ.
The above-described “sputtering” boundary condition can be schema-
tized through the following equations for R and P populations (Falcucci
et al., 2016; Krastev et al., 2016a):

A B C
6 2 5

f6 f2 f5

f3 f1
3 1

f7 f4 f8

7 4 8

Fig. 3 (A) Sketch of the employed lattice stencil: a 2D D2Q9 lattice (Qian et al., 1992)
characterized by four near neighbors (1–4) and four second (5–8) nearest neighbor
sites. The numbers run over the discretized velocities; (B) and (C): sketch of the pro-
posed reactive boundary condition: the green arrow in (B) represents a quasiparticle
of R species hitting a solid site (black circle) in the pore bulk. The length of the arrow
sketches the magnitude of the corresponding Boltzmann’s discrete population and,
thus, the (local) mass of the R species; in (C), the green (reactant) and red (product)
arrows sketch the coast-line molecular distributions after the reaction and the
desorbing event (“sputtering” condition).
88 Giacomo Falcucci

X
bðxÞ
fi R, in ðxÞ ¼ ð1  pRS Þ Si, j fjR, out ðx  cj Þ (4)
j¼1

X
bðxÞ
fiP , in ðxÞ ¼ ðp pRS Þ Si, j fjR, out ðx  cj Þ
j¼1
(5)
XbðxÞ
+ ð1  pS Þ
P
Si, j fjP , out ðx  cj Þ,
j¼1

in which the superscript in and out stand for the population that enters and
exits the a lattice node. In Eqs. (4) and (5), Si, j is a random sputtering matrix
that renders the coast-line spreading of desorbed species. Si, j obeys the
PbðxÞ
conservation rule i¼1 Si,j ¼ 1: such a matrix provides the probability of
a molecule to leave the bulk along direction “j” and reenter along direction
“i”. Since coefficients pRS , pPS and p take into account sticking and reaction
events, the conservation rule for matrix Si, j is granted.
In our first schematization, the relaxation to local equilibrium takes place
on a time scale taken to be equal for all species (Falcucci et al., 2016): in our
simplified reaction model, in fact, it is not necessary to account for the dif-
ferent relaxation time of the singles species, even though our numerical
approach is able to consider a τα for each α chemical component.
Once Eq. (1) is solved, the macroscopic gas fields are recovered, namely
the density, ρα(x, t), and velocity uα(x, t) of each species, from the first two
moments of the distribution functions f αi ðx, tÞ:
X
b
ρα ðx, tÞ ¼ fiα ðx, tÞ, (6)
i¼0
X
b
ρα ðx, tÞuα ðx, tÞ ¼ ci fiα ðx, tÞ: (7)
i¼0

At local equilibrium, we assume that all species move with the common
barycentric velocity:
X X
uðx, tÞ ¼ ρα ðx, tÞ uα ðx, tÞ= ρα ðx, tÞ: (8)
α α

The weights wi in the equilibrium distribution function are chosen in order


to satisfy the lattice isotropy constraints (Qian et al., 1992):
X
b X
b X
b
wi ¼ 1, wi ci ¼ 0, wi ci ci ¼ cs2 1, (9)
i¼0 i¼0 i¼0
LBM for 2D and 3D chemical reactors 89

where 1 denotes the identity matrix and index b runs on the lattice direc-
tions (9 in 2D and 19 in 3D). The lattice speeds are ci ¼ (0;0), with a weight
wi ¼ 4/9, ci ¼ (1;0) and ci ¼ (0;1), with wi ¼ 1/9, and ci ¼ (1;1),
with wi ¼ 1/36. Finally, cs is the lattice speed of sound, which, for the
pffiffiffi
chosen stencil, is equal to 1= 3.
In the limit of small Knudsen numbers (defined as the ratio between the
mean free path and the characteristic length scale of the phenomenon under
investigation), Eq. (1) reproduces the Navier–Stokes equation for a Carrier
fluid C of viscosity ν ¼ c 2s ðτ  1=2Þ in lattice units. In a standard hydrody-
namic regime, typically τ  1/2 ≪ 1: in the present applications, which deal
with fluid flows characterized by Knudsen numbers Kn  1, we work in the
regime τ  1/2  1.

4. Second act—Fighting: Validating the numerical


model
As a first benchmark, we performed a comparison between the pro-
posed chemical boundary condition and the analytical solution proposed by
Lev^eque’s for a 2D (laminar) reactive flow between parallel plates (Falcucci
et al., 2016; Levęque, 1928). In Lev^eque’s framework, reactions are located
at the bottom side of the plate. As a result, the concentration vertical gradient
at the reactive wall can be determined as:
!13
∂C R 1 4Pe
¼   (10)
∂y 1 4 x
93 Γ Ly
3

in which CR is the reactant concentration, x is the streamwise direction, Γ is


the Gamma function, Ly is the dimension along the crossflow direction y,
and Pe is the Peclet number, set to 253 for the simulation at hand. This
test was performed on a 150  50 2D computational domain; as regards
the boundary conditions, a body force was implemented in the computa-
tional domain, to simulate a constant pressure gradient along the whole
channel. In the simulation, the reactant species is continuously injected at
the inlet and it reacts only with the lower plate. The reaction probability
p in the sputtering boundary condition has been set to p ¼ 1, in order to
simulate instantaneous reaction (corresponding to a Damk€ ohler number
Da ! ∞): thus, all reactants impacting on the catalytically active nodes
are instantly converted into products.
90 Giacomo Falcucci

Fig. 4 Result of the benchmark test against Lev^eque’s analytical solution for a 2D laminar
reactive flow between parallel plates. The concentration gradient along the crossflow y
direction at the reaction plate is plotted along the channel length x streamwise direction.

The comparison reported in Fig. 4 shows the remarkable agreement with


Lev^eque’s analytical solution, with the overall mean absolute error approx-
imately at 4% (Falcucci et al., 2016).
As a further validation, we considered the catalytic conversion processes
that take place inside the reactors for nonsteady temporal analysis of products
(TAP) experiments (Feres et al., 2009b). In Feres et al. (2009b), experimen-
tal and numerical results for TAP analysis are reported together with a prob-
abilistic theory for catalytic conversion, according to three different reactor
geometries, as sketched in Figs. 5 and 6. The experimental conditions in
Feres et al. (2009b) correspond to pure Knudsen diffusion (i.e., Kn > 1), with
a single reactant diffusing through the test chamber. In our numerical
simulations, the considered species flow through the domain with a
characteristic Knudsen number Kn  0.65.
In the computational reactors depicted in Fig. 5A–C, the black square
represents the active catalytic media; the surrounding porous region is
chemically inert and it is characterized by an average volumetric porosity
of ε ¼ Vvoid/Vtot  70%.
In Feres et al. (2009b), a platinum (Pt) foil was employed as the catalyst
medium for the single-particle experiments: such a foil was characterized by
a thickness of 0.1 mm and a width of 3.17 mm, and it was enclosed within an
inert quartz porous medium, characterized by an average pore dimension of
210–250 μm.
LBM for 2D and 3D chemical reactors 91

Fig. 5 Sketch of the three computational domains realized to reproduce the experimen-
tal setups studied in Feres et al. (2009b). The black square represents the single-catalyst
particle, and it is located in different positions inside the surrounding (inert) porous
media. In all the panels, black stands for solid obstacles, whereas light-gray represents
the void through which the fluid flows. The experimental apparatus is characterized by a
length of 8.27 mm (along the horizontal direction) and a height of approximately
4.93 mm (vertical direction). The porous region is reconstructed generated via the ran-
dom seeding algorithm proposed in Falcucci et al. (2016), Krastev and Falcucci (2018),
and Montessori and Falcucci (2018). The above-reported three layouts are all character-
ized by the same value of the volumetric porosity ε ¼ Vvoid/Vtot  70%; the flow direction
is left-to-right in each schematization. (A) Case A; (B) Case B; (C) Case C.

Fig. 6 Sketch of the experimental apparatus for TAP tests. The yellow dot marks the posi-
tion of the catalytic single particle, which is reproduced in Case A in Fig. 5. In the above
sketch, reactant flow is from top to down (left-to-right in Fig 5).
92 Giacomo Falcucci

To assess the grid-independency of our results, we performed the sim-


ulations in each of the four layouts in Fig. 5, according to four different space
resolutions, resulting in the following computational domain sizes:
• 124  84, i.e., “small” domain;
• 248  168, “medium” domain;
• 496  336, “large”, and
• 992  672, “extra-large”.
The above domain sizes correspond to approximately 2.5, 5, 10, and 20 grid
points per pore.
At the inlet of the computational domain, a pulsed mass flow boundary
condition was set: C and R species are injected for the first 6000 time
steps, corresponding to roughly  6  104s, since the lattice time step is
Δt  107 s. After this initial time span, the inlet condition was switched
off and the species were let diffuse throughout the domain for the rest of
the simulation, which was set to last from 106 to 107 time steps (corresponding
to 101–100 s), depending on the chosen grid resolution.
The overall reactant species conversion was measured according to
Falcucci et al. (2016) and Feres et al. (2009b):
M tot,P
η¼ , (11)
M tot,R
where Mtot,P and Mtot,R denote the total masses of products and reactants.
Details of the numerical simulations are reported in Table 1, where the
parameters were chosen based on experimental input (Falcucci et al., 2016).

Table 1 Details on the baseline parameters for the benchmark simulations.


Lattice units
ε 70%
Mean pore diameter δ (inert porosity) 4
p 0.9992
pRS 1.0
pPS 0.0
uin 0.08
Carrier concentration at inlet 0.60
Reactant concentration at inlet 0.20
Product concentration at inlet 0.0
LBM for 2D and 3D chemical reactors 93

Fig. 7 reports the time evolution of the integral of P concentration within


the computational domain. In the figure, time is normalized according to
tMax, that is the time at which the maximum value of Product mass is
obtained in each simulation.
From Fig. 7, it is apparent that the resolutions for the medium, large and
extra-large domain yield basically the same results, for Cases A and B.
Considering Case C, the differences may be due to the position of the cat-
alytic particle, which is close to the outlet of the domain. Such a layout
significantly affects the catalytic performance and it led to a partial mismatch
between theoretical and experimental values in Feres et al. (2009b), as well.
To evaluate the conversion efficiency in Eq. (11), we compute the mass
of Reactant species as follows:
Z tsim
M tot,R ¼ ΦR jin dt, (12)
0

in which tsim is the total simulation time and ΦRjin is the flux of Reactant
species through the inlet. Through Eq. (12), we consider the actual mass
of Reactants entering the computational domain; thus, this is the total
available mass to convey the chemical conversion within the TAP apparatus.
The total mass of P species, Mtot,P, is evaluated by considering the peak value
in the trends reported in Fig. 7, according to the different grid resolutions.
Fig. 8 reports the conversion efficiency as a function of the position of the
catalyst within the reactor: in the figure, theoretical and experimental values
from Feres et al. (2009b) are reported as a benchmark.
As shown in Fig. 8, LB simulations provide a fairly reasonable agreement
with the results in Feres et al. (2009b), considering the nonnegligible differ-
ences in the flow regime. In fact, it is important to stress that in Feres et al.
(2009b), pure Knudsen diffusion cases where investigated, with Kn > 1, in
the so-called ballistic regime, which is known to be critical for fluid dynam-
ics, in general (Marin et al., 2019; Succi, 2001). Our analysis was centered on
flow regimes marginally in the Knudsen diffusion regime, that is character-
ized by a Kn number 0.65. Moreover, from an operative point of view, in
Feres et al. (2009b), only the first 10 reactant pulses where analyzed as an
average: considering just few pulses, in fact, allowed to focus on the max-
imum activity of the catalytic surface, before oxygen consumption from
the surface of the catalyst would have became the bottleneck of the chemical
conversion. In our schematization, no consumption mechanisms were
included and thus, only one single pulse was simulated.
450 450 450
Integral of product concentration

A 400 B C
400 400
350 350 350
300 300 300 124 x 84
248 x 168
250 250 250 496 x 336
200 200 200 992 x 672
150 150 150
100 100 100
50 50 50
0 0 0
0.01 0.1 1 10 0.001 0.01 0.1 1 10 100 0.001 0.01 0.1 1 10 100
t/tMax t/tMax t/tMax

Fig. 7 Evolution in time of P mass integral, inside the whole computational domain, according to the three layouts reported in Fig. 5. In each
panel, the trends according to the four grid resolution are reported; tMax is the time step corresponding to the maximum value of P, and it is
used to normalize the time axis. (A) Case A; (B) Case B; (C) Case C.
LBM for 2D and 3D chemical reactors 95

1
Case A
0.9 Case B
Case C
Conversion efficiency
0.8

0.7

0.6

0.5

0.4

0.3
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Position in reactor
Fig. 8 Conversion efficiencies obtained by means of LB numerical simulations and the
experimental and theoretical predictions reported in Feres et al. (2009b). LBM results are
represented by the black dots, equipped with the error bars related to the different grid
resolutions explored for each of the reactor layouts reported in Fig. 5; the vertical
segments express the magnitude of theoretical predictions, from Feres et al. (2009b);
finally, the dashed boxes represent the experimental uncertainty, as discussed in
Feres et al. (2009b).

Given these nonnegligible differences between the two investigations,


the results shown in Fig. 8 provide a fairly accurate prediction of the hetero-
geneous catalysis in pulsed flow reactors for transitional flows, that is for 0.1 <
Kn < 1: such a regime, even though not as extreme as the ballistic one, still
represents a terra incognita between high-vacuum chemical methods (charac-
terized by Kn > 1 with pressure in the order of p  103  104 torr) and
molecular-based theories (in which p > 101 torr) (Marin et al., 2019).

5. Third act—The good win: Answering the question


Once the validation phase was accomplished, we turned back to the
initial, sparking question on the morphology changes of np-Au ingots. To
convey more details, the considered chemical conversion is the following:
2 CH3 OH + O2 ! HCOOCH3 + 2 H2 O (13)
which is carried out at 150°C in a continuous flow of reactant (6.5%
MeOH–20% O2–73.5% He) with a volumetric flow rate of 50 mL/min.
The effluent gas is analyzed by means of a gas chromatograph (Agilent
HP 7890A) coupled to a mass spectrometer (MSD 5975C), equipped with
96 Giacomo Falcucci

two columns in tandem operation (HP-PLOT/Q and CARBONPLOT).


Scanning electron microscopy (SEM) analyses are performed on a Zeiss
Supra 55VP instrument. The computational simulation is a 2D grid of
500  250 lattice units, representing a 5 μm  2.5 μm section of the reactor,
with a space resolution of 10 nm. The porous substrate (which represents
np-Au) is reconstructed by placing a solid rectangular area in the middle
of the computational domain and seeding this solid zone with a number
N of “void seeds”; the diameter of this seeds is, then, let grow up to a max-
imum value of 30 nm (corresponding the actual pores in the real ingot). The
number N of seeds is chosen in order to meet the typical void fraction of
np-Au derived from Ag70Au30 alloy, which corresponds to a porosity
ε (void volume/total volume) of ε  0.70  0.005. The detailed space res-
olution that this methodology ensures allows to match the most relevant
experimental parameters, such as the typical diameter-to-thickness aspect
ratio of np-Au (20), and it proved sufficient to study 30 nm pores. The
characteristics of the simulation domain and the details of the digital porous
medium are sketched in Fig. 9.
For the experimental visualization of the gas diffusion kinetics in np-Au,
alumina (Al2O3) atomic layer deposition (ALD) experiments were per-
formed. Instead of using long saturation exposures that lead to the formation
of homogeneous coatings (Biener et al., 2011), short trimethyl-aluminum
(TMA)/H2O ALD cycles (1 s/1 s at 0.8 torr, 125°C) were considered, as
related to diffusion-limited regime.

Fig. 9 Sketch of the computational domain, with the detail on the porous morphology
of the np-Au reconstructed as proposed in Falcucci et al. (2016).
LBM for 2D and 3D chemical reactors 97

Table 2 Main chemical and physical parameters used in the numerical


simulations; for each parameter the lattice nondimensional values and
the corresponding physical units are reported; all values are reported
for the reference case, corresponding to w ¼ 10 and p ¼ 0.45.
Lattice units Physical units
L 250 2.5  106 m
H 500 5  106 m
h 200 2  106 m
w 10 107 m
ρinlet 1 0.833 mL/s
δx 3 3  108 m
pffiffiffi
v 1= 3 1000 m/s
pffiffiffi
τtr 3 3 5  1011 s
τch 2 2.5  1011 s
pRS 0.95 NA
pPS 0.00 NA
p 0.45 NA

Table 2 summarizes the most chemical and physical parameters of our


simulations, both in lattice nondimensional units and the corresponding
physical parameters.
The main physical parameters of the simulation are the Peclet number,

2 u R
Pe ¼ , (14)
D
the diffusive Damk€
ohler number,
D τch
DaD ¼ 2 , (15)
R
and Knudsen number,
 
Δt 1
Kn ¼ c s τ  , (16)
2 δ
 is the
In the above, u is the reference value for the velocity magnitude, R
reference physical dimension, D the diffusion coefficient of the Reactant
98 Giacomo Falcucci

(taken to be the same as the Product) and τch the typical time scale for
local chemical reactions, τch ¼ Δt/p. In the above, δ is the typical pore size
(3 lattice units), τ is the collision time and Eq. (16) stems from the expres-
 
sion of the mean free path in LBM fluids, λ ¼ c s τ  Δt 2 , compared to the
reference length, which corresponds to the average pore size (Falcucci
et al., 2016).
In lattice nondimensional units, the initial values for the densities of the
Carrier (ρC), and the two reactive species (ρR and ρP) inside the computa-
tional domain, are:

ρC ¼ 0:8, ρR ¼ ρP ¼ 0 (17)

thus, neither R nor P species are present, inside the computational domain,
at the beginning of the simulation. Considering the inlet boundary condi-
tion, located at x ¼ 0, Carrier and Reactant are injected at a constant speed,
Uin ¼ 0.08, with the corresponding densities ρC ¼ 0.8 and ρR ¼ 0.2, respec-
tively. Once R species reaches the porous slab, Products start to be delivered,
according to the chemical reaction R ! P that sketches reaction (13). R and P
species are passively transported by the Carrier and their flow regime is char-
acterized by Knudsen diffusion within the pores. The sticking probabilities of
the reactant and products are fixed to pRS  0:5 and pPS  0:0, respectively: in
Falcucci et al. (2016), it was explained that these values came from lower-
scale electronic structure simulations, which clearly indicated that
Reactants are attracted and stick to the solid surface, as the absorption
energies are in the order of Eabs  0.8–1 eV, while Products perceive a
smaller desorption barrier (with a desorption barrier in the order of
Edes  0.1–0.2 eV), thus P tends to desorb more easily. At 150°C, the
choice of a zero sticking coefficient for the Product species is fully justified
(Falcucci et al., 2016). Fig. 10 sketches the absorption and desorption
energy landscape. In the layout reported in Fig. 9, the rear (lower) side of
the ingot is almost inactive, and η is expected to scale like pu ¼ h + 2w;
whereas, for full perimeter conversion (ingot parallel to the fluid stream),
the efficiency should scale with the total perimeter ptot ¼ 2h + 2w. These
scaling relationships rely on homogeneity assumptions, which are eventually
broken at the corners of the ingot, where larger gradients are found. Note
that the scaling is the same as for a 3D slab of cross section h  h and
depth w, with w ≪ h.
The results of the simulations are reported in the density contour plots of
Reactants and Products, flowing around and through the catalytic ingot.
LBM for 2D and 3D chemical reactors 99

Fig. 10 Sketch of the energetic prospect of Reactant absorption on solid sites, together
with Product desorption back to the fluid region. Once they get in proximity of the pore
walls, Reactants are attracted with an absorption energy of Eabs  0.8–1 eV. Reactants,
then, turn into Products by overcoming an activation barrier of Ereact  0.4 eV. Finally,
Products face a much smaller desorption barrier of Edes  0.1–0.2 eV.

Fig. 11 Examples of Reactant (A) and Product (B) concentrations at t ¼ 10,000 time
steps, in the presence of p ¼ 0.45 reaction parameter; the flow direction is from left
to right in all panels.

Fig. 11 shows the results in terms of concentration contours for R and P


species. Inspection of the (steady-state) density contours of the reactant
ρR(x) and product densities ρP(x) reveals that:
1. Reactants do not diffuse deeply into the porous structure, as they are effi-
ciently converted into Products on the gas-facing surface of the ingot;
100 Giacomo Falcucci

2. the majority of the products leave the porous structure through the inlet-
facing surface, while a smaller fraction diffuses through the sample and
leaves through the back face.
This finding was confirmed by the trend of P flux along the ingot width,
which is reported in Fig. 12
Fig. 12 shows that the flux of Products exhibits an abrupt increase in the first
layers followed by a much slower rise inside the bulk of the ingot. A shallow
catalyst layer on the ingot side facing the gas stream shows a remarkable con-
version efficiency, thus shadowing the rear face of the ingot, which does not
contribute substantially to the overall conversion efficiency (see Fig. 11).
We suggested that such a significantly higher conversion at the gas-facing side
(due to different local gas-surface chemistry) is the “smoking gun” for the exper-
imentally observed change in the surface pattern between the front and rear
sides, as shown in Fig. 1.
To corroborate these findings, we resorted to the experimental activities of
our colleagues from Harvard University. Fig. 13 reports the cross-sectional
SEM images collected from the surface of a np-Au sample after exposure
to one short TMA/H2O ALD cycle (1 s/1 s at 0.8 torr). Microscopy analysis
shows that the deposition of Al2O3 is observed only in a thin layer of 20 μm
thickness close to the sample surface; moreover, such a film appears to be
separated from unreacted material by a sharp interface.
For the purpose of our numerical simulations, Al2O3 ALD represents the
(ceiling) case of a surface reaction characterized by a TMA reactive sticking

w/h = 0.10
1.2
p = 0.35
1 p = 0.45
p = 0.55
p = 0.65
0.8
Φ Prod

0.6

0.4
p = 0.35
0.2 p = 0.45
p = 0.55
p = 0.65
0
100 110 120 130 140 150
y
Fig. 12 Product flux along the main flow direction (top to bottom in Fig. 11) after
t ¼ 30,000 time steps, for w/h ¼ 0.1 and for different values of the reaction probability
p. The area of the ingot is delimited by the black, vertical dashed lines.
LBM for 2D and 3D chemical reactors 101

Fig. 13 Cross-sectional SEM image of a fractured, 200 μm thick np-Au sample after
exposure to 1 TMA/H2O ALD cycle (1 s/1 s at 0.8 torr). The 1 s long sequential
TMA/H2O exposure results in the deposition of a one monolayer thick Al2O3 coating
on the ligaments of np-Au within a 20 μm thick surface layer (bright area of the cross
section) separated by a sharp interface from uncoated np-Au (Falcucci et al., 2016;
Montemore et al., 2017).

probability pRS of 103–104 and 0.00 on empty and filled sites, respectively
(Ott et al., 1997).
In contrast to our results, Products formed during TMA exposure
(a chemisorbed 0AlMe2 species) do not desorb. The SEM cross section
shown in Fig. 13 reveals that a 1 s TMA/H2O exposure (at 0.8 torr) allows
the diffusion through the first 20 μm layer of the porous structure. The rel-
atively thick reaction layer is due to the fact that Products do not desorb, as a
consequence of the low sticking coefficient. The desorption of Products
would lead to surface regeneration near empty catalytic centers and, conse-
quently, to a faster consumption of the Reactant species. Thus both the
Reactant sticking probability and Products lifetime seem to influence
how much the bulk of the sample contributes to Product formation.
Once these results were obtained, we investigated the effects of sample
thickness on conversion efficiency, η, especially in the presence of low reac-
tion probabilities. The overall conversion efficiency is computed consider-
ing the integral of reactant flux through the exposed area of the pores, as a
function of the reaction probability. We found that the number of molecules
that actually benefit of the increase in thickness is actually very small, because
102 Giacomo Falcucci

the probability of deep molecular diffusion decreases linearly with the


sample thickness, as exposed in Gordon et al. (2003): thus, for a low reaction
probability, the loss of unreacted R species via backscattering through the
surface facing the gas inlet represents the dominant contribution in reducing
the conversion efficiency, for thicker samples. Thus, we investigated how
thick a horizontal catalyst should be to still allow for efficient conversion,
by performing a systematic study on the effect of the reaction probability p
(taking, p ¼ 0.35, 0.45, 0.55, and 0.65) and the ingot dimension w/h
(w/h ¼ 0.03, 0.05, 0.1, and 0.15), at the given porosity ε  0.70  0.005.
The analysis of flux of products along the y direction (i.e., perpendicular to
the gas flow),
X
ΦP ðyÞ ¼ ρP ðx, yÞ uðx, yÞ, (18)
x

is reported in Fig. 14: from the panel, it is apparent that at higher values of p,
there is almost no appreciable variation in the depth of the reaction: the large
part of R conversion in P develops in a thin, shallow layer of the ingot on the
side facing the gas flow.
To further corroborate these findings, we studied the effects on the con-
version efficiency η of the relative time scales of diffusion and mass transport,
for different values the reaction probability. It is known that η is governed by
the Damk€ ohler number, which is given ratio between the chemical reaction

1.1

0.9

0.8
ΦProd

0.7
w/h = 0.03
0.6 w/h = 0.10
w/h = 0.15
0.5

0.4
0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7
p
Fig. 14 Product flux at y ¼ 150 for different aspect ratios w/h, as a function of the reac-
tion parameter p.
LBM for 2D and 3D chemical reactors 103

time scale τch ¼ Δt/p ¼ 1/p in lattice units, and the mass transport time scale
τtr ¼ δ=u, where u is the average molecule velocity inside the pore, which
we take of the order of the lattice speed of sound cs (Falcucci et al., 2016).
The limit Da ! 0 corresponds to infinitely fast chemical reactions, in
which the reactivity is limited by transport phenomena. The opposite limit
is the case of infinitely slow chemistry, Da !∞, in which the bottleneck is
represented by the timescale of the reactions. For Da ! 0, the thickness w of
the slab is not expected to have an impact on η, while for Da !∞, w plays a
major role, as Reactant residence time through the ingot must be enough
for the conversion to take place. In Falcucci et al. (2016), we considered
Da  0.4, marginally in the fast-chemistry limit.
In physical units, we estimated τch  25 ps and τtr  50 nm/1000 m/s
50 ps. We define the conversion efficiency as the mass of the P species
at the outlet over the mass of R at the inlet, that is:
M P jy¼L
η¼ , (19)
M R jy¼0
P
where MR,P(y) ¼ xρR,P(x, y) is the total Reactant mass at the given loca-
tion. We studied the effect of ingot thickness w (ranging from 6, 10, 20, to
30, all in lattice units with Δx ¼ 10 nm) on the conversion efficiency η while
keeping constant the ratio between the inner glass tube diameter H and the
ingot length h (i.e., 200:500), similar to the experimental value 1:2.5. In all
simulations, the porosity is fixed to ε  0.70  0.005.
Fig. 15 reports the trends of η for different values of w/h and p parameters.
More specifically, we considered the aspect ratio corresponding to
w/h ¼ 0.05, that is the actual ratio of the ingot employed in the experiments
(a np-Au disk of 5 mm diameter and 250 μm thickness). Fixing a width of
w ¼ 10 in lattice units, we were able to reproduce the experimentally mea-
sured conversion efficiency of η ¼ 0.2 by setting a reaction probability
p ¼ 0.45. Such a probability corresponds to a characteristic chemical time
of 25 ps. The results of our simulations showed that the contribution of
the ingot thickness on the conversion process is negligible compared
to the effect of the sample perimeter, as apparent by the values of η for
w/h ! 0.
The chemical efficiency in Eq. (19) can be related to the chemo-physical
and geometrical parameters, as follows:
 
h v τch
η¼ φ , (20)
H l
104 Giacomo Falcucci

35

30

25

20
h

15

10 p = 0.35
p = 0.45
5 p = 0.55
p = 0.65

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
w/h
Fig. 15 Conversion efficiencies η as functions of the ingot geometrical ratio w/h; differ-
ent values of the conversion efficiency p are reported (the pink open square represents
the experimental value), considering a ratio 1:2.5 between the ingot diameter h and the
glass tube dimension H. The trends exhibit a nonzero value of η as w/h ! 0: this con-
firms that the ingot perimeter (equivalent to surface in 3D) plays a dominant role in the
conversion process, rather than its area (volume in 3D).

where l  w is the typical path length within the ingot, u the average molec-
 
ular speed and τch the chemical time scale. In the above expression, φ u lτch
denotes a generic (decreasing) functional dependence on the dimensionless
quantity, Da ¼ u τch =l (Damk€ ohler number at the ingot level). This can also
be decomposed as Da ¼ ð u τch =δÞ  ðδ=lÞ Dap  Ap , where Dap is the
Damk€ ohler number at the pore level and Ap ¼ δ/l is the aspect ratio of
the pores.
As a further benchmark, we investigated the differences in conversion effi-
ciency as due to the sample orientation in respect to the gas flow. More spe-
cifically, we considered an ingot placed parallel to the flow stream, as shown in
Fig. 16: such a layout is frequently used in filtration devices and, for the current
application, may have led to higher efficiencies in the presence of very high
reaction probabilities (and, thus, to a very shallow depth distribution of P).
Fig. 17 highlights the different conversion efficiency of the two config-
urations, due to the fluid dynamic distribution of the reactants (as reported in
Fig. 16).
Finally, We investigated the effect of the reaction parameter p on the R
conversion, for an aspect ratio of w/h ¼ 0.05. From our simulations, we
saw that even though the parallel setup exposes both faces of the plate, thus
LBM for 2D and 3D chemical reactors 105

Fig. 16 Reactant (A) and Product (B) concentrations at t ¼ 10,000 time steps, in the pres-
ence of p ¼ 0.45 reaction parameter, for an ingot major axis placed parallel to the flow
direction (from left to right).

100 0

80 20

Conversion efficiency
60 40
Rout/Rin

40 Horizontal ingot 60
Vertical ingot

20 80

0 100
0 5000 10000 15000 20000 25000 30000
Time steps
Fig. 17 Conversion efficiency of the np-Au ingots, due to their orientation in respect to
the flow directions, as illustrated in Figs. 11 and 16.

doubling the active catalytic surface, the ingot now intercepts a consider-
ably smaller fraction of the gas flow, with the large majority bypassing the
sample without ever coming in contact with its surface. Thus, despite the
increase in exposed surface area, the overall conversion efficiency gets
dramatically reduced.
All the trends and fields reported in the previous figures provides evi-
dence that the large part of the chemical conversion process takes place in
a shallow layer of the porous catalyst, in its inlet-facing side. For all values
106 Giacomo Falcucci

of p, conversion from R to P is characterized by two trends across the ingot


thickness: a first steep trend in this shallow, inlet-facing layer, and a slight
slope through the sample bulk. This two-trend behavior supported the con-
clusion that the perimeter of the ingot (in 2D, its area in 3D) played the
major role in the conversion process, as featured in Fig. 15. The outlet-
facing side of the sample was characterized by the absence of steep gradients
in P flux, for all the values of p: this confirmed that the two sides of the ingot
were characterized by a totally different activity in the conversion process, as
also suggested by the experimental observations. Thus, the difference in
their surface pattern, as shown in Fig. 3 appeared, now, as a natural conse-
quence of such different chemical contributions.

6. “First steps in a larger world”a : np-Au hollow spheres


Once we accomplished our investigations on the small, rectangular
np-Au ingot, we considered a more general landscape for nano-catalysis: that
represented by the exploitation of micro- and nano-spheres, that allow very
efficient conversion performance, due to their high surface-to-volume ratio.
We carried out our simulations in a 380  300 grid box, considering a
single-pulse regime, as described in Section 4. The adopted grid spacing was
the same as in Falcucci et al. (2016). We schematized the small porous
spheres as thin shell with 6 lattice site thickness; their average diameter is
d ¼ 120 lattice sites, corresponding to a physical diameter of 8 μm with a
lattice spacing δx  67 nm. The corresponding value of the lattice time steps
is δt  107 s.
In Fig. 18, the layout of the np-spheres used in our simulations is
reported. We focused on the same experimental apparatus as the one
described in the previous Section; as a further step, instead of a nano-porous
ingot (thin cylinder), here we use broken or unbroken spheres (Falcucci
et al., 2016, 2017).
The catalytic nano-spheres (n-s) are surrounded an inert porous zone, as
sketched in Fig. 18: such a media is only employed to keep the catalyst with a
prescribed position and orientation during the experiment (Feres et al.,
2009b) and it is characterized by a volumetric porosity of ε  85%. The
porous n-s, instead, have ε  70%. Table 3 summarizes the main physical
characteristics of both porous media.

a
Comment by Obi-Wan Kenoby to Luke Skywalker, "Star Wars: Episode IV—A New Hope," 1977.
LBM for 2D and 3D chemical reactors 107

Fig. 18 (A) and (B): Geometrical configurations of the porous spheres in the concave
and convex layouts, as employed in our simulations; (C) layout of the entire hollow
sphere of np-Au; for all these geometries, the flow direction is left-to-right, in our sim-
ulation. (D) SEM image of a cracked np-Au hollow sphere. Panel (D): Courtesy of Dr.
Branko Zugic.

Table 3 Main characteristics of the catalytic nano-sphere and the


surrounding inert porous media.
Lattice units Physical units
Hollow spheres
Average diameter 120 8 μm
Thickness 6 400 nm
Pore diameter 2 130 nm
Porosity 0.5 0.5
Inert porous media
Pore diameter 10 500 nm
Porosity 0.82 0.82
108 Giacomo Falcucci

Due to the remarkable thickness of the porous n-s, in the actual exper-
iments it is common to find them broken, as shown in Fig. 18D. This is the
reason why we chose to simulate not only whole nano-spheres, but also
cracked layouts, and in a concave and convex configuration in respect to
the gas flow, as shown in Fig. 18A–C.
Considering the gas flow inlet in our simulations, we fixed an injection
timing corresponding to the first 6000 time steps, for the C and R species.
After this initial, pulsed flow, the inlet condition is set to zero for the rest of
the simulation, leaving all species to diffuse through the computational domain.

6.1 Results and discussion


Referring to the nondimensional numbers defined in Eqs. (14)–(16), here we
have u representing the average flow velocity, as evaluated at the middle of the
domain, the location in which the porous n-s are placed; R, as a consequence,
is the n-s radius and the pore dimension δ is fixed to 2 lattice units.
We considered very long simulation duration, up to 106 to 2  107 time
steps, depending on the values of the considered Kn numbers. For all these sim-
ulations, we evaluated the conversion efficiency η according to Eq. (11): we
considered the integral of P concentration within the entire computational
domain, evaluated at the time step corresponding to its peak value, as reported
in Fig. 19.

600
Integral of product concentration

Concave, Kn = 0.75
500 Concave, Kn = 0.15
Concave, Kn = 0.05
400 Convex, Kn = 0.75
Convex, Kn = 0.15
300 Convex, Kn = 0.05

200

100

0
1000 10,000 100,000 1x106 1x107 1x10
8
1x10
9

Time steps
Fig. 19 Time evolution of the integral of P concentration within the entire computa-
tional domain, for different values of the Kn number. From the figure, we see that
the convex layout yields a marginally higher conversion of R species, but the dominant
effect on conversion attains to the Knudsen number of the flow.
LBM for 2D and 3D chemical reactors 109

From our simulation, it is possible to quantify the effect of gas rare-


faction on the overall chemical conversion process: the reference experi-
ments conducted in Feres et al. (2009b) are characterized by a Knudsen
regime of Kn  0.1  1, according to the operating pressure, 104 torr,
and to the given reference length of inert particles within the porous
media, L  200 μm.
LBM is known to recover the hydrodynamic solution of Navier–Stokes
equations in the limit of small Kn numbers (Succi, 2001), but we have
already seen in the previous section that it shows reliable results also in
the slip-flow regimes, Kn  0.5 (Falcucci et al., 2016).
In this section, we further increase the value of Kn number, exploring
the effects of its variation according to three values: Kn  0.04,  0.15,
and  0.70; such a value gets marginally in the Knudsen diffusion regime
(Michaelides, 2014; Succi, 2001). In Fig. 19, the integral of P concentration
inside the whole computational domain is reported, for different values of
Kn number and for both the concave and convex n-s layouts. From such
a figure, it is apparent that P yield is only marginally affected by the catalyst
layout, with a slightly higher conversion from the convex n-s, while the
dominant impact is due to the Kn flow regime.
In our simulations, due to the chosen Kn, species diffusion becomes
more and more important, as Kn gets higher, compared to standard advec-
tion. As the flow gets more rarefied, in fact, the diffusion through porous
media becomes faster: such an increase in diffusivity provides a more effi-
cient conversion of R into P, as expected from the results in the literature
(Feres et al., 2009a,b).
Figs. 20 and 21 report snapshots of the evolution in time of R and P spe-
cies concentrations for the concave and convex n-s layout. In both panels,
the concentrations at 9500 and 19,500 simulation steps, are reported, respec-
tively: from the visual inspection, it is possible to see that for both the con-
sidered layouts, there is a dominant back-diffusion of P toward the inlet of
the computational domain: such a behavior has already been addressed in the
literature, in the presence of nano-scale porous materials (Falcucci et al.,
2016, 2017).
It is interesting to compare the results obtained with broken and unbroken
n-s layouts, according to the layouts reported in Figs. 18A–C. Fig. 22 reports
the contours of R and P species concentrations for the unbroken n-s, at the
same time steps as in Figs. 20 and 21, that is t ¼ 9500 and t ¼ 19,500 time steps;
Fig. 23 shows the evolution in time of the integral of P concentration inside
the whole domain is reported, for the cracked layouts. In order to effectively
110 Giacomo Falcucci

Fig. 20 Reactant (left) and Product (right) concentrations for the concave n-s at 9500
and 19,500 time steps of the simulation. The panel highlights that the inlet-facing side
of the catalytic n-s is the most active zone and that P species diffuse within the whole
porous media, with a prevalence toward the inlet of the domain (Falcucci et al., 2016;
Feres et al., 2009b). The concave layout magnify such an effect, as the Products remain
“trapped” inside the n-s. (A) t ¼ 8500 time steps; (B) t ¼ 19,500 time steps.

Fig. 21 Reactant (left) and Product (right) concentration magnitudes at 9500 and 19,500
time steps of the simulation, for the convex hollow sphere. As in Fig. 20, the inlet-facing
portion of the n-s is the most active in the chemical conversion process. (A) t ¼ 8500
time steps; (B) t ¼ 19,500 time steps.
LBM for 2D and 3D chemical reactors 111

Fig. 22 Reactant (left) and Product (right) concentration magnitudes for the unbroken
hollow sphere, at 9500 and 19,500 time steps of the simulation. As for the broken lay-
outs, the inlet-facing side of the n-s is the most active. The “trapping” effect is milder,
compared to the concave layout, and the outlet-facing side of the n-s shows to provide a
contribution in the conversion process as R flow reaches the second half of the compu-
tational domain. (A) t ¼ 8500 time steps; (B) t ¼ 19,500 time steps.

600
Integral of product concentration

Kn = 0.75
500 Kn = 0.15
Kn = 0.05
400

300

200

100

0
100 1000 10,000 100,000 1x106 1x107
Time steps
Fig. 23 Evolution in time of the integral of P concentration, evaluated inside the entire
computational domain, for the unbroken n-s layout. The trends are reported according
to three values of the Kn number. The figure provides further evidence of the impor-
tance of Kn number in P yield, and highlights that the unbroken layout provides a
slightly higher conversion, compared to the broken layouts.

compare the conversion efficiency of the broken and unbroken layouts, we


have considered three inlet velocities for the pulsed injection, for a given value
of the flow Knudsen number, Kn ¼ 0.75: uin ¼ 0.08, 0.12, and 0.16. The
corresponding values of Pe and DaD numbers are reported in Table 4.
112 Giacomo Falcucci

Table 4 Main nondimensional numbers in our numerical simulations.


DaD Pe
uin ¼ 0.08 1.07  104 6.00  105
uin ¼ 0.12 1.07  104 7.20  105
uin ¼ 0.16 1.07  104 8.44  105

1
Hollow sphere
Convex
0.8 Concave

0.6
h

ζ C = 0.678
0.4
ζ C = 0.616
ζ C = 0.594
0.2

0
0.06 0.08 0.1 0.12 0.14 0.16 0.18
u in
Fig. 24 Trends of the catalytic conversion efficiency η for different values of the inlet
velocity, at Kn ¼ 0.75. In the panel, the value of the chemical conversion gain ζC are
reported, for the various Pe numbers, as well. It is apparent that the global efficiency
of the broken layouts is definitely comparable to that of the entire sphere, as it is a func-
tion of the catalytic surface exposed to the reactant stream.

In Fig. 24, the trends of conversion efficiency for the three values of uin
are reported: from the panel, it is apparent that η is a decreasing function of
the Peclet number. This is a logical consequence of the fact that Pe  u/vth
Kn, hence reducing the ratio u/vth  Ma has the same effect as increasing Kn.
From the comparison between the conversion efficiencies of the three
different layouts, it is possible to infer the effect of sphere rupture on η.
To this purpose, we defined a chemical conversion gain ζ C, due mechanical
rupture, as follows:
ηCC + ηCV  ηHS
ζC ¼ , (21)
ηHS
in which ηCV and ηCC are the chemical conversion efficiencies for the
convex and concave n-s, respectively, while ηHS is the conversion efficiency
for the unbroken n-s. All the efficiencies in Eq. (21) are computed according
to Eq. (11).
LBM for 2D and 3D chemical reactors 113

Fig. 24 provides the values of ζ C for the three values of inlet velocity:
considering the according to Eq. (21), we find ζ C  60% in the three cases.
Thus, ζ C highlights the variation in η according to the superficial layout of
the n-s: ζC > 0 corresponds to a surplus in reactivity of the cracked layouts.
This is in line with what expected as broken n-s are characterized by a higher
surface/volume ratio, compared to the whole n-s.
Finally, by assuming that all the external surface of the hollow sphere is
active in the chemical conversion process, the geometrical analog of ζ C in
terms of exposed surfaces is:

2ðR2 + R1 Þ  2R2 R1
ζG ¼ (22)
2R2 R2
where R1 and R2 are the inner and outer radii of the n-s. In our case,
R1 ¼ 4  0.2 ¼ 3.8 μm and R1 ¼ 4 + 0.2 ¼ 4.2 μm, respectively, yielding
ζ G ¼ 0.905, above the observed value of ζ C  0.6, as reported in Fig. 24.
Thus, ζ G can be regarded as an upper bound to ζC.

7. Microbial fuel cells (MFC)


In this section, we export the concept of our active porous media to
the study of microbial fuel cells (MFC) ( Jannelli et al., 2017; Logan and
Regan, 2006; Logan et al., 2006; Puig et al., 2010). MFCs are microbial
electrochemical reactors that exploit particular strains of microbes that digest
organic waste releasing electrons e and protons H+ as byproducts of their
own metabolism (Logan, 2009; Nastro et al., 2017a; Santoro et al., 2013).
This numerical activity is grounded on the experiments conducted at the
Laboratory of Energy Systems of the University of Naples “Parthenope”
(Gambino et al., 2017; Nastro et al., 2017b). According to this experimental
background, single-chambered MFC layout has been considered, with the
anodic chamber characterized by the presence of a porous anode realized
with carbon fibers and a porous cathode exposed to ambient air. The anodic
chambered is filled with a slurry composed of water and solid organic waste
( Jannelli et al., 2017). Electroactive bacteria adhere to the anode electrode
and digest the organic material, releasing H+ and e.
The electrodes are reconstructed via a porous medium, realized by means
of the procedure discussed in Section 5, with a number N of void seeds
grown up to a diameter of 200 μm (corresponding to the average pore
dimension in the employed electrodes, as shown in Figs. 25 and 26).
114 Giacomo Falcucci

Fig. 25 Numerical and experimental layouts of the MFC reactors considered in our
investigation; (A) the Falcon test tube described in Jannelli et al. (2017); (B) the carbon
fiber anode and (C) the porous cathode. Panel (D) shows the computational domain,
with details of the porous structures of the cathode and anode electrodes; in (D), the
presence of the Nafion membrane is highlighted: it is implemented as a wall no-slip
for water and solid-waste species, while it is a transparent internal boundary for the
H+ ions.

Fig. 26 Qualitative comparison between the original and the digitalized porous
structure of the cathode: (A) cathodic porous structure, designed and realized in the
Energy System Laboratory of the University of Naples “Parthenope” (Jannelli et al.,
2017); (B) detail of the porous material at the cathode realized through the fractal
methodology described in Section 5 (Falcucci et al., 2016).
LBM for 2D and 3D chemical reactors 115

The number of disks is chosen in order to achieve a final volumetric


porosity E (void volume/total volume) within the electrode Eanode  0.72
 0.005 and Ecathode  0.76  0.005, corresponding to the void fractions
of the employed porous materials ( Jannelli et al., 2017).
According to the active (sputtering) boundary condition presented in
Section 3, the “active” zone is represented by the anode electrode, in which
H+ species is generated from the digestion (i.e., conversion) of the solid
waste; the cathode, on the other hand, is schematized as an inert porous
media: this choice does not imply any loss of generality, as the model can
account for more complex chemical reactions at the cathode side.
Finally, the presence of Nafion membrane at the cathode, as in Jannelli
et al. (2017), is taken into account by implementing a variable internal
boundary in front of the cathode: it is treated as a wall no-slip for water
and solid-waste species, while it acts as a transparent boundary for H+ species.

7.1 Results and discussion


Before presenting the results, we briefly recall the conversion between
lattice, nondimensional units to physical parameters, which represents one
of the crucial aspects when dealing with LB simulations (Benzi et al.,
1992b; Chen et al., 2015). We assume that the conversion for any physical
parameter is based on the following, simple, equation:

Γphys ¼ Υ Γlb (23)


in which Γ represents a generic physical quantity and Υ is the relative con-
version scale.
The kinematic viscosity of the slurry inside our reactors must be
evaluated considering the feedstock as a dilute suspension, thus we can
employ the following formula by Albert Einstein (Einstein, 1956):
 
5
μS ¼ μL 1 + ϕ , (24)
2
in which ϕ is the solid volume fraction of the suspension, which is ϕ ¼ 0.25 in
the case of our reactors ( Jannelli et al., 2017). Thus, according to Eq. (24), we
have μS  1.63  103 kg/ms. Assuming a density of ρS  1.0 103 kg/m3
( Jannelli et al., 2017) we find νS  1.63  106 m2 s.
Now, we known that in the LB universe the value of the kinematic
viscosity (and, thus of the relaxation parameter τ) can become critical for
the onset of numerical instabilities (Benzi et al., 1992a; Montessori and
116 Giacomo Falcucci

Table 5 Main physical parameters and their LB conversions.


Physical quantity LB quantity (lu)
Length 0.1 m 400
Time 1s 256
Viscosity 1.63  106 m2/s 1.66667
pH 3.75 6.75
Mass concentration 1.0  10 kg/m
3 3
1
4
+
H concentration 1.78  10 mol/L 1.78  107
OCV 0.5 V 0.005

Falcucci, 2018). By fixing νlb ¼ 1.66667 lu (thus ensuring numerical


stability), we find Υvisc ¼ 9.78  107 m2 s lu.
Once we have the conversion length scale for kinematic viscosity, we
found the scale for physical length: since our computational domain was fixed
to 400  100 lattice sites, Υlength ¼ 0.1 m/400 ¼ 2.50  104 m/lu.
Repeating the same procedure for time, we have Υtime ¼
Υ2length =Υvisc ¼ 256: each lattice time step, thus, corresponds to roughly
4  103 s/lu.
Table 5 summarizes the most relevant physical quantities, expressed by
means of their physical units and the converted lattice, nondimensional value.

7.2 Activity of bacteria strains


The behavior of bacteria strains at both electrodes has a major impact on the
reactor electrochemical performance (Cheng et al., 2006; Frattini et al.,
2016; Logan and Regan, 2006; Logan et al., 2015).
Without loosing generality, in our schematization we considered the
presence of electro-active bacteria strains only at the anode, and we
accounted for their activity through a typical thriving curve, as that reported
in Fig. 27 (Di Ilio et al., 2019).
Such a curve epitomizes the evolution of microbe strains during the
experiment: in the very first days, there is a growing phase, followed by a
thriving period, in which the number and activity of bacteria remains
almost constant, and a decay stage, during which the number (and
activity) of microorganism reduces, due to the loss of material to be
digested and to the decrease in pH caused by bacteria metabolism
( Jannelli et al., 2017).
LBM for 2D and 3D chemical reactors 117

1.2

pB 0.8

0.6

0.4

0.2

0
0 5 10 15 20 25 30 35 40
Time (days)
Fig. 27 Thriving curve for anode bacteria strains. The trend has been normalized along
the experiment duration (which typically lasted 35  40 days), thus providing the value
of p for our conversion model.

The time trend for our numerical experiments was devised by means of a
set of flexible functions (Falcucci, 1985) in the form of

k 
Y 
Ci
yðtÞ ¼ 1  eAi ðBi tÞ : (25)
i¼1

To reconstruct the typical shape of bacterial thriving curve (Chotibut et al.,


2017; Zwietering et al., 1990) in Eq. (25) k ¼ 2 and the constants Ai, Bi, Ci
are retrieved from fitting the experimental data in Jannelli et al. (2017) and
Nastro et al. (2017b). To be noted, the trend in Fig. 27 is normalized, thus
providing the time evolution of conversion probability pB of Species 2 into
Species 3, according to the activity of anodic microbes.

7.3 Ion motion within MFC reactor


Inside the MFC reactor, ions move due to the presence of the following
effects (Di Ilio and Falcucci, 2020):
• potential difference between the electrodes.
• diffusion;
• repulsion between positive ions;
The LB scheme accounts automatically of the species diffusion within the
reactor; on the other hand, repulsive interactions as well as the effect of elec-
trode potentials need to be modeled.
118 Giacomo Falcucci

The repulsion between H+ ions is schematized according to Gauss’s Law,


r  E ¼ r2 ϕ ¼ ½Hε , in which E is the local electrical field, ϕ is the electrical
+

potential associated with E, [H+] is the local concentration of H+ ions and ε


is the permittivity of the medium, taken to be the same as water, ε  80 F/m.
The potential difference between the electrodes is a function of the electric
field, according to the following:

r  Ecell ¼ r2 V , (26)
in which Ecell is the local magnitude of the electrical field, related to the
potential difference ΔV. The repulsive force between positive ions, then,
is modeled via a body force that accounts for the local concentration of
H+ ions. The LBM implementation of such repulsive force is obtained
through a pseudo-potential approach, a simple and versatile technique that
has been successfully employed in the LBM framework for multiphase and
multicomponent investigations (Benzi et al., 2006; Falcucci et al., 2007,
2008; Montessori and Falcucci, 2018):

r ψ2
F rep ¼ G c 2s , (27)
2
in which ψ is the pseudo-potential that accounts for the local density of the
H+ species, ψ ¼ 1  e½H  and G represents the strength of the repulsive
+

interaction, which is given by G ¼ q=c 2s , with q the electrical charge of


H+ ions. From a numerical point of view, the force in Eq. (27) is positively
defined, as it is a repulsive interaction, acting on species H+, according to the
local concentration, [H+]. Besides this repulsive interactions, ions experi-
ence the potential difference between the electrodes, which is embedded
into our model by means of the Lorentz force:

F ¼ q  Ecell (28)
where q is the charge of each particle.

7.4 Polarization and power curves


Fig. 28 shows the comparison between power and polarization curves,
according to our numerical simulations and the experimental data reported
in Jannelli et al. (2017). The trends reported in Fig. 28 are normalized against
the maximum values obtained in the simulations and in the experiments,
respectively. The Figure shows a notable accord in the trends predicted
by our LB scheme, compared to the experimental measurements reported
LBM for 2D and 3D chemical reactors 119

1.2 0.5
LBM
EXP
1 Power - LBM
Power - EXP 0.4

0.8
ΔV / ΔVOCV 0.3

P / (Vmax •Imax )
0.6
0.2
0.4

0.1
0.2

0 0
0 0.2 0.4 0.6 0.8 1 1.2
I / I max

Fig. 28 Trends of polarization and power curves obtained by means of the proposed LB
model, compared to the experimental data presented in Jannelli et al. (2017) and
Krastev and Falcucci (2019).

in Jannelli et al. (2017). More specifically, Fig. 28 shows that the proposed
LB scheme is able to correctly capture the typical trends of MFC operation:
• for small I current values, the model does not show any appreciable
activation losses, as typically happens for MFC operation (Logan and
Regan, 2006);
• the second phase is that of ohmic losses, which represents the large part of
our reactor operation capability (Krastev and Falcucci, 2019): for this
region, the agreement between LBM and experiments is remarkable;
• finally, no appreciable concentration losses are found, both with our
numerical method and in the experimental measurements, which again
is a typical operating condition for solid-waste MFCs ( Jannelli et al.,
2017; Nastro et al., 2017a,b), and again shows a good agreement
between the two approaches.
The comparison shown in Fig. 28 highlights the reliability and accuracy of
our methodology, despite the adopted simplified reaction mechanism.
Fig. 29 reports the evolution of H+ concentration in time, according to
our numerical simulations. The visual inspection of the species concentra-
tion and current fields, related to the last days of operation, shows H+ ions
accumulating in front of the cathode, in the region of the Nafion membrane:
this provides an interesting explanation for the experimental data reported in
Jannelli et al. (2017): in the region close to the cathode, in fact, pH values
lower than those in proximity of the anode electrode were found. The
numerical simulations highlight that such a phenomenon is due to the
limited motility of H+ species through the cathodic porous media, rather
Fig. 29 Evolution in time of H+ relative concentration within the computational domain:
0.0 represents the initial concentration, corresponding to pH¼ 3.75 (6.75 in lattice units,
according to Table 5). The panel highlights the migration of H+ concentration peak from
the anode toward the cathode electrode. In the final days of MFC operation, after day
20 according to Fig. 27, snapshots (D) and (E) highlight the higher concentration of H+
close to the cathode, as experimentally measured in Jannelli et al. (2017). (A) Day 5;
(B) Day 15; (C) Day 25; (D) Day 30.
LBM for 2D and 3D chemical reactors 121

than to MFC inversion (Hartline and Call, 2016; Kim et al., 2007; Logan
et al., 2015; Oh and Logan, 2007).
In Fig. 30, the insight of the H+ current field within MFC reactor is
provided. Form the snapshots, the peak in proton current in the thriving

Fig. 30 Time evolution of H+ current within the computational reactor. The panel high-
lights the peak in H+ current in the first days, corresponding to the initial part of the
thriving phase of bacteria activity, see Fig. 27. The panel shows an interesting detail
of the current peaks both inside anode and cathode electrodes, providing interesting
insights and new criteria for the design of innovative material for electrode realization.
(A) Day 5; (B) Day 15; (C) Day 25; (D) Day 30.
122 Giacomo Falcucci

phase of the bacteria (approximately, days 5–20) located in the part of the
anode electrode in front of the cathode: this aspect must be taken into
account for the design of innovative types of MFC reactor, in order to avoid
the fouling of the cathode membrane during the experiment (Rabaey and
Verstraete, 2005).
Finally, the insight in the current field provides interesting details in the
local current peaks in the cathode electrode: this aspect, as well, must be care-
fully accounted for in the adoption of innovative materials for industrial appli-
cations of MFC’s ( Jannelli et al., 2017; Liu et al., 2008; Nastro et al., 2017a, b).

8. 3D Microbial fuel cells


In this section, the extension of the previous model to a 3D microbial
reactor will be presented. For these simulations, Eq. (1) is discretized in a
D3Q19 lattice in three dimensions, with 19 characteristic velocities for
particle streaming, as depicted in Fig. 31. The layout of the MFC reactors
considered in this Section is based on that described in Logan et al.
(2007). More specifically, experiments have been designed, realized and
carried out at University of Naples “Parthenope”, as reported in Jannelli
et al. (2018) and Minutillo et al. (2018). In these reactors, the carbon fiber
brush at the anode is replaced by a porous compound made by biochar.

Fig. 31 Sketch of the lattice stencil employed for 3D simulations of MFCs (Di Ilio
et al., 2019).
LBM for 2D and 3D chemical reactors 123

Fig. 32 (A) Microbial reactor realized from Logan et al. (2007); (B) 3D computational
domain.

Fig. 32A and B shows the experimental reactor and the digitalization of such
a setup, with the porous anode and cathode reconstructed as explained in
Section 5 (Falcucci et al., 2016, 2017; Krastev and Falcucci, 2018, 2019;
Montessori and Falcucci, 2018). The anode and the cathode electrodes
are characterized by a porosity Eanode  56% and Ecathode  77%, respectively,
where the porosity is defined as the ratio between the void volume and the
overall electrode volume.
To investigate the performance of the 3D layout, the same law for
bacteria growth, thriving and decay periods as represented by Eq. (25)
was adopted (Di Ilio et al., 2019). Figs. 33 and 34 show the evolution in time
of H+ and OFMSW within the electrochemical reactors, for different values
of the electric potential.
124 Giacomo Falcucci

Fig. 33 Time evolution of H+ concentration within the computational electrochemical


reactor: the initial value of H+ concentration corresponds to pH ¼ 4. The panel shows the
migration in time (from A to C) of H+ ions from the anode toward the cathode electrode,
according to the different values of the potential (and thus, the presence of an external
load). Figures in the right panel correspond to a high external resistance and low current
(ideally, OCV conditions), while the left side corresponds to the case of very low external
resistance (ΔV ! 0). The results are in line with the experimental measurements con-
ducted in Jannelli et al. (2017).
Fig. 34 Evolution in time (from A to C) of the OFMSW concentration within the reactor.
The right side of the panel corresponds to a high external resistance and low current (ide-
ally, OCV conditions), while the left side corresponds to the case of very low external resis-
tance (ΔV ! 0).
126 Giacomo Falcucci

Fig. 35 Polarization and power trends obtained via the proposed 3D LB model, com-
pared to the experimental data presented in Jannelli et al. (2017). The comparison
shows a slightly worse agreement between the experimental and numerical results,
which may be ascribed to the differences in the considered layouts.

With the 3D model it is possible to evaluate the electrochemical perfor-


mance of the reactors in terms of polarization and power trends, for different
values of the external resistance. Fig. 35 reports these trends: from the figure
it is apparent that the 3D model is fairly in agreement with the experimental
data, but with a slightly less accuracy as compared to the 2D predictions
reported in Fig. 28. This might be ascribed to the differences between
the layouts of the two experiments. However, the 3D model is capable of
reproducing the activation and concentration losses, as well as the ohmic loss
region (Di Ilio et al., 2019).
LBM for 2D and 3D chemical reactors 127

9. Selective catalytic reactors (SCR) for diesel engine


applications
For this particular application, the numerical model is chosen such
that, the probability p of the conversion reaction to take place is connected
to the chemical kinetics of standard NO conversion to N2 in presence of
NH3, given by

4NH3 + 3O2 ! 2N2 + 6H2 O, (29)

which is known to be governed by an Arrhenius-type law in the form of

k ¼ A eRT
Ea
(30)

In the above, A is the Arrhenius pre-exponential factor, Ea the activation


energy, R the universal gas constant and T the absolute temperature. For the
case represented by Eq. (29), we have adopted the following values (Krastev
et al., 2016a; Madia, 2002): A ¼ 8.35  1010 cm3/gs, Ea ¼ 8.5 103 J/mol,
R ¼ 8:314 J=mol K and T ¼ 350 K. With the above values, we find k ¼
6.233 m3/kg s.
Once the kinetic parameter of the chemical reaction is set, it is linked to
the residence time θ of reactant species on the catalyzed surface of the porous
media, which c an be computed according to the following relation, which
governs the evolution of species concentration in time:

C ¼ C 0 ekθ : (31)

In Eq. (31), C0 and C are the NO concentration inside the exhaust gas at the
inlet and outlet section of the SCR reactor, respectively.
As a first application case (Krastev et al., 2016a), we have considered a
heavy-duty diesel engine certified according emission standard adopted
by California ARB on October 21, 2014, the Optional low NOx standards
for heavy-duty engines (DieselNet, n.d.). Under the program, manufacturers
may choose to certify their engines to three optional NOx emission stan-
dards: 0.10, 0.05, or 0.02 g/bhph. If we consider an engine characterized
by 0.10 g/bhph, the SCR system to be designed to reach the 0.02 g/bhphr
standard class must convert the 75% of NOx content: for a clean, new
reactor, we can fix a ratio C/C0  0.15, in order to assure the desired
NOx conversion in time.
128 Giacomo Falcucci

With the adopted conversion rate and considering Eq. (31), we find
 
τ ¼ 1k ln CC0  1:1 s. For the conversion parameter p, we imposed a value
of p ¼ 0.9 corresponding to a local 90% of reactant conversion into
products.

9.1 Results and discussion


Table 6 reports the main parameters for our simulation.
Fig. 36 shows the field of concentrations of the two species NO and N2
after t ¼ 95,000 time steps. The figure highlights that the large part of NO
conversion to N2 takes place close to the inlet section of the SCR reactor, as
reported by other works in literature (Krastev et al., 2016b). Fig. 36 shows
the trends of species concentration along the reactor axis: the figure confirms
that a considerable conversion of NO to N2 takes place in the first part of the
SCR reactor. The concentration of N2, then, decreases as the products
diffuse in the channel toward the exit of the domain.
Fig. 37 reports the current magnitude for NO and N2. The current J is
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
defined as Jðx, tÞ ¼ ρðx, tÞ juðx, tÞj2 + jvðx, tÞj2, with u(x, t) and v(x, t) the
components of flow velocity along the horizontal and vertical directions,
respectively.

Table 6 Main thermo-physical and chemical parameters for our simulation.

nx 4000
ny 200
Chemical species 3
Viscosity 1/6
Δx 30 (μm)
ε 70%
Re 100
Reaction parameter p 0.05
Inlet NO concentration 0.18
Inlet N2 concentration 0.0
Exhaust gas temperature 623 (K)
LBM for 2D and 3D chemical reactors 129

Fig. 36 Density contours of NO (A) and N2 (B) species after t ¼ 95,000 time steps.

6 0.02
[NO]
[N2]
5
0.016

4
0.012
[NO]

[N2]

3
0.008
2

0.004
1

0 0
0 500 1000 1500 2000 2500 3000 3500 4000
x
Fig. 37 Species concentration along the SCR axis, after t ¼ 95,000 time steps.
130 Giacomo Falcucci

100
Numerical simulations
Expected

80

60
h (%)

40

20

0
0 50,000 100,000 150,000 200,000
Time steps
Fig. 38 Species concentration along the SCR axis, after t ¼ 95,000 time steps.

We define the conversion efficiency as


½NOinlet  ½NOoutlet
η¼ (32)
½NOinlet
where [NO] is the total concentration of NO at the inlet and outlet section.
Fig. 38 shows the trend of conversion efficiency in time, compared to the
expected value computed in the previous sections. The figure shows the
agreement between our numerical results and the expected theoretical effi-
ciency, based on kinetic predictions.

10. 3D SCR systems


In this section, we extend to three dimensions the approach illustrated
in the previous Section, for the analysis of Selective Catalytic Reactor
Systems for automotive applications. Here we will leverage on the approach
already discussed in the previous sections to reconstruct a complex 3D
porous reactor. Through our approach, the fluid dynamic fields together
with the chemical reactions for NOx reduction can be efficiently dissected
across the various involved physical scales. In fact, through the LBM it is
possible to achieve a remarkably higher detail, compared to traditional
LBM for 2D and 3D chemical reactors 131

Navier–Stokes-based solvers, with a light computational effort. Such an


accuracy can be employed, not only to characterized SCR components a
posteriori, but also as an efficient support tool for the reactor design.
Also for our 3D analysis, we will refer to the coast-line sputtering boundary
condition, depicted in Fig. 3, presented in Section 3.
Fig. 39 highlights the computational domain. It includes four channels of
a Corning Celcor SCR substrate, whose characteristics are reported in
Table 7. At the inlet of the computational domain (left in Fig. 39 (top)) a

Fig. 39 Representation of the simulated SCR channel: (top) computational domain: it


represents the entire length of a Corning Celcor SCR; (bottom-left) detail of the inlet sec-
tion; (bottom-right) detail of the microstructure of the channel inner surface. The texture
of the cordierite substrate is apparent from the figures. As wall free-slip, in order to
account for the symmetry properties of the cordierite support.

Table 7 Main characteristics of Corning Celcor SCR substrate (Corning


Environmental Technologies, n.d.).

Product cpsi/web 400/6


Bulk density (g/L) 395
Open frontal area 0.757
2 3
Geometric surface area (cm /cm ) 27.4
Heat capacity 200° (J/KL) 352
Hydraulic diameter (mm) 1.11
Wall thickness (mm) 0.18
Porosity ε 55%
132 Giacomo Falcucci

velocity inlet condition is imposed; at the outlet (right in the same panel), an
outflow condition is implemented. The upper, bottom, front and rear
boundaries are set as wall free-slip, in order to account for the symmetry
properties of the cordierite support.
The LB solver is kept isothermal (Krastev et al., 2019); Table 8 reports
the main parameters for our simulation.
Figs. 40 and 41 report the contours of species concentration and velocity
magnitude, respectively, at the inlet section of the SCR reactor. From the

Table 8 Main thermo-physical and chemical parameters for our simulation.

Domain length L (mm) 50


Domain height and width (mm) 2.74
# chemical species 3
2
Kinematic viscosity (m /s) 5.50  105
Δx (μm) 50
Re 100
Reaction parameter p 0.9
Exhaust gas temperature (K) 623

Fig. 40 Detail of (A) Reactant and (B) Product concentration fields at the inlet of the SCR
reactor.
LBM for 2D and 3D chemical reactors 133

Fig. 41 Detail of (A) Reactant and (B) Product velocity magnitude at the inlet of the SCR
reactor.

Figure, it is apparent the conversion of reactant into product (NO to N2),


which takes place mainly at the inlet section of the reactor (ANSYS, 2020).
To quantify the effective conversion performed by our SCR model, we
define the conversion efficiency as:
½N2 outlet
η¼ (33)
½NOinlet
where [NO] is the total concentration of NO at domain inlet and [N2] the
concentration of N2 at domain outlet.
Fig. 42 shows the trend of conversion efficiency in time, compared
to the expected value computed in the previous Sections. The figures
shows three values of η according to three values of the species residence
time τ ¼ L/uin. The Figures highlights the influence of τ on the overall
conversion efficiency, which appears to be lower than the expected
values of η  90%. Besides the influence of τ, such a value can also
be ascribed to the overall length of the SCR reactor reported in
Table 8 which is the minimum value among those commercialized by
Corning (Corning Environmental Technologies, n.d.).
The chosen grid spacing reported allows to accurately capture the
chemo-physical phenomena at pore scale. Fig. 43 shows the detail of
134 Giacomo Falcucci

0.8

0.7

0.6
Conversion efficiency h

0.5

0.4 t1 = 1075
t2 = 1000
0.3 t3 = 909

0.2

0.1

0
0 2000 4000 6000 8000 10,000 12,000
Time (l.u.)
Fig. 42 Trends of conversion efficiency η in time, according to different values of the
species residence time within the reactor.

Fig. 43 Detail of Product concentration and velocity magnitude at the inlet of the SCR
reactor; the vectors are colored according to the respective species concentration. The
figures highlights the detail of chemical conversion at pore scale provided by the LBM
solver.

the Product concentration at the inlet section of the reactor: the release of the
new species from the reactor active sites is apparent. The random and discrete
nature of the sputtering boundary condition is reflected in the orientation of
the released P populations, which reproduces the mesoscopic effect of the wall
orientation on the gas dynamics: such a randomness in the released P field is
clearly visible in the velocity vectors reported in Fig. 43.
LBM for 2D and 3D chemical reactors 135

11. Conclusions
In this chapter, the capabilities of LBM employment for the simulation
of phenomena characterized by moderately high values of the characteristic
Knudsen numbers have been assessed. More specifically, starting from a
sparking question related to the reshaping of surface pattern of a nano-
porous gold (np-Au) catalytic ingot, the Method has been extended toward
the simulation of heterogeneous catalytic reactions within nano- and micro
pores. First of all, we conceived a methodology to accurately reconstruct the
nano-porous structure of the catalyst, without having to resort to scanning
electron microscopy (SEM) and digitalization. The random nano-porous
structure recreated by means of our methodology proved accurate, from
a fluid dynamic point of view, being equivalent to the original one in terms
of porosity, permeability, and tortuosity. Then, we developed a chemically
active boundary condition, for the conversion of Reactant into Products,
with a coast-line sputtering of the released species, that proved particularly
suited to render the behavior of species desorption from active nano-pores.
After a preliminary validation phase, in which our model proved reliable and
versatile, with a light computational effort, our simulations attacked the ingot
surface problem, highlighting that the large part of conversion reactions took
place on a thin, shallow layer on the inlet-facing side. The outlet-facing side,
on the other hand, was characterized by a much lower activity: this behavior,
was the “smoking gun” for the ingot grain reconfiguration, due to the energy
release related to the involved chemical reactions. Once the sparking prob-
lem of ingot reconfiguration was solved, other phenomena of scientific and
technical concern have been investigated, providing interesting and reliable
results with light computational burden. More specifically, heterogeneous
catalysis within pulsed-flow reactors was investigated: in these experiments,
the flow conditions are even more extreme, leaning toward the ballistic regime,
which represents the forbidden planet for non-ad hoc solvers. Even if Kn ¼ 1
remains unbearable for a standard LBM implementation, we obtained reliable
results for Kn numbers up to Kn  0.75, a regime that is marginally in the terra
incognita of rarefied fluid dynamics, confirming the accuracy, reliability and
versatility of our numerical approach.
Once the investigations on heterogeneous catalysis were accomplished,
we turned out attention to a completely different field, that of electrochemical
reactors, with particular focus on microbial fuel cells (MFCs). In these devices,
complex chemical reactions take place, entangled with challenging transport
136 Giacomo Falcucci

phenomena that involve the slurry within the cell. Even for this analysis, and
for both 2D and 3D numerical schemes, the developed model was able accu-
rately capture and predict the electrochemical and power performance of the
reactor in terms of polarization and power curves, and to properly identify
the species density and velocity fields, helping to providing reliable answers
to the open questions rising from the experimental results.
As a final application, the focus was set on SCR reactors, which are fun-
damental devices for emission control of internal combustion engines. In
such reactors, chemical conversions take place on the surface of a catalyzed
substrate, which is exposed to fully unsteady flow conditions, also in the
presence of variable (high) mass flow rates. The proposed methodology
was capable to deliver accurate and reliable predictions, both for 2D and
3D implementations, with a light computational effort, for such a challeng-
ing framework, as well.
Summarizing, the novel proposed approach for heterogeneous catalysis
proved efficient and reliable in many different applications; it aptly merged
with LBM transport properties, delivering a thorough and stable numerical
solver capable of disentangling complex phenomena of reactive flows within
nontrivial geometries and for finite Kn numbers up to Kn  0.75.
The obtained results provide a computational tool for the design of
new-generation catalytic material and substrates for many applications across
physical scales.

Acknowledgments
This work was partially supported by the Italian Government Research Project “FCLab - Sistemi
innovativi e tecnologie ad alta efficienza per la poligenerazione,” PON03PE_00109_1, with
Prof. Elio Jannelli as the Scientific Responsible, partially by the Italian Ministry Program
PRIN, grant no. 20154EHYW9 “Combined numerical and experimental methodology for
fluid structure interaction in free surface flows under impulsive loading,” with Prof. Chiara
Biscarini as the scientific responsible, and by the University of Rome "Tor Vergata" Grant
Mission: Sustainability (D.R. 2817/2016), CUP: E86C18000400005, with G. Falcucci as the
Principal Investigator. The simulations were performed on Zeus, the HPC facility of the
University of Naples “Parthenope.” Zeus was realized through the Italian Government
Grant PAC01_00119 MITO—Informazioni Multimediali per Oggetti Territoriali, with Prof. Elio
Jannelli as the scientific responsible.
This chapter was concluded during the terrible COVID-19 epidemic that so many lives
cut all over the world. In those very long and difficult days segregated at home, I received the
comfort of many colleagues and friends in Rome (Prof. S. Succi, Prof. G. Bella, Prof. V.
Krastev, Prof. G. Di Ilio, Dr. Andrea Montessori, and Dr. Marco Lauricella), which
I wish to thank for their closeness and their precious contributions to the realization of
this chapter. Along with them, I would like to thank Prof. E. Kaxiras, from Harvard
University, for his friendship, his very kind support and for the many valuable discussions
we had to deliver the novel LBM methodology here presented.
LBM for 2D and 3D chemical reactors 137

References
Ansumali S, Karlin IV: Kinetic boundary conditions in the lattice Boltzmann method, Phys
Rev E 66(2):26311, 2002.
ANSYS: ANSYS release 20.0 documentation, 2020, ANSYS Inc.
Bell AT: The impact of nanoscience on heterogeneous catalysis, Science 299(5613):
1688–1691, 2003. https://doi.org/10.1126/science.1083671.
Benzi R, Succi S, Vergassola M: The lattice Boltzmann equation: theory and applications,
Phys Rep 222(3):145–197, 1992a.
Benzi R, Succi S, Vergassola M: The lattice Boltzmann equation: theory and applications,
Phys Rep 222(3):145–197, 1992b.
Benzi R, Biferale L, Sbragaglia M, Succi S, Toschi F: Mesoscopic modeling of a two-phase
flow in the presence of boundaries: the contact angle, Phys Rev E 74(2):21509–21514,
2006. https://doi.org/10.1103/PhysRevE.74.021509.
Bernaschi M, Fatica M, Melchionna S, Succi S, Kaxiras E: A flexible high-performance
Lattice Boltzmann GPU code for the simulations of fluid flows in complex geometries,
Concurrency Comput Pract Exper 22(1):1–14, 2010.
Biener J, Hodge AM, Hayes JR, et al: Size effects on the mechanical behavior of nanoporous
Au, Nano Lett 6(10):2379–2382, 2006.
Biener MM, Biener J, Wichmann A, et al: ALD functionalized nanoporous gold: thermal sta-
bility, mechanical properties, and catalytic activity, Nano Lett 11(8):3085–3090, 2011.
Bird GA, Brady JM: Molecular gas dynamics and the direct simulation of gas flows, (vol 5), Oxford,
1994, Clarendon Press.
Cali A, Succi S, Cancelliere A, Benzi R, Gramignani M: Diffusion and hydrodynamic
dispersion with the lattice Boltzmann method, Phys Rev A 45(8):5771, 1992.
Chen S, Chen H, Martnez D, Matthaeus W: Lattice Boltzmann model for simulation of
magnetohydrodynamics, Phys Rev Lett 67(27):3776, 1991.
Chen L, Wu G, Holby EF, Zelenay P, Tao W-Q, Kang Q: Lattice Boltzmann pore-scale
investigation of coupled physical-electrochemical processes in C/Pt and non-precious
metal cathode catalyst layers in proton exchange membrane fuel cells, Electrochim Acta
158:175–186, 2015. https://doi.org/10.1016/j.electacta.2015.01.121.
Cheng S, Liu H, Logan BE: Increased performance of single-chamber microbial fuel cells
using an improved cathode structure, Electrochem Commun 8(3):489–494, 2006.
https://doi.org/10.1016/j.elecom.2006.01.010.
Chotibut T, Nelson DR, Succi S: Striated populations in disordered environments with
advection, Phys A Stat Mech Appl 465:500–514, 2017.
Corning Environmental Technologies, n.d., https://www.corning.com/cala/en/products/
environmental-technologies/products/ceramic-substrates/corning-celcor-substrates.html.
Di Ilio G, Falcucci G: Multiscale methodology for microbial fuel cell performance analysis,
Int J Hydrogen Energy, 2020. https://doi.org/10.1016/j.ijhydene.2020.04.057, ISSN:
0360-3199.
Di Ilio G, Dorschner B, Bella G, Succi S, Karlin IV: Simulation of turbulent flows with the
entropic multirelaxation time lattice Boltzmann method on body-fitted meshes, J Fluid
Mech 849:35–56, 2018.
Di Ilio G, Chiappini D, Ubertini S, Bella G, Succi S: A moving-grid approach for fluid-
structure interaction problems with hybrid lattice Boltzmann method, Comput Phys
Commun 234:137–145, 2019.
Di Ilio G, Nastro RA, Di Trolio P, Falcucci G: A comprehensive multiscale methodology to
investigate microbial fuel cell performance, 2019: In 2019, pp 3795–3811. https://www.
scopus.com/inward/record.uri?eid¼2-s2.0-85081534903&partnerID¼40&
md5¼712e67a5a963cffc5a73868e8b63cbd3.
Di Staso G, Clercx H, Succi S, Toschi F: DSMC-LBM mapping scheme for rarefied and
non-rarefied gas flows, J Comput Sci 17:357–369, 2016.
DieselNet, n.d., http://www.dieselnet.com
138 Giacomo Falcucci

Duan H, Wang D, Li Y: Green chemistry for nanoparticle synthesis, Chem Soc Rev 44(16):
5778–5792, 2015.
Einstein A: Investigations on the theory of the Brownian movement, 1956, Courier Corporation.
Falcucci LF: Funzioni flessibili, 1985, Ed. Accademia Aeronautica.
Falcucci G, Bella G, Chiatti G, Chibbaro S, Sbragaglia M, Succi S: Lattice Boltzmann models
with mid-range interactions, Commun Comput Phys 2(6):1071–1084, 2007.
Falcucci G, Chibbaro S, Succi S, Shan X, Chen H: Lattice Boltzmann spray-like fluids,
Europhys Lett 82:24005, 2008.
Falcucci G, Chiatti G, Succi S, Mohamad AA, Kuzmin A: Rupture of a ferrofluid droplet in
external magnetic fields using a single-component lattice Boltzmann model for nonideal
fluids, Phys Rev E Stat Nonlinear Soft Matter Phys 79(5 Pt 2):056706, 2009. https://doi.
org/10.1103/PhysRevE.79.056706.
Falcucci G, Ubertini S, Biscarini C, et al: Lattice Boltzmann methods for multiphase flow
simulations across scales, Commun Comput Phys 9(2):269–296, 2011a.
Falcucci G, Ubertini S, Chiappini D, Succi S: Modern lattice Boltzmann methods for
multiphase microflows, IMA J Appl Math 76(5):712–725, 2011b. https://doi.org/
10.1093/imamat/hxr014.
Falcucci G, Jannelli E, Ubertini S, Succi S: Direct numerical evidence of stress-induced
cavitation, J Fluid Mech 728:362–375, 2013. https://doi.org/10.1017/jfm.2013.271.
Falcucci G, Succi S, Montessori A, et al: Mapping reactive flow patterns in monolithic
nanoporous catalysts, Microfluidics Nanofluidics 20(7):105, 2016. https://doi.org/10.1007/
s10404-016-1767-5.
Falcucci G, Amati G, Krastev VK, Montessori A, Yablonsky GS, Succi S: Heterogeneous
catalysis in pulsed-flow reactors with nanoporous gold hollow spheres, Chem Eng Sci
166:274–282, 2017.
Feres R, Cloninger A, Yablonsky G, Gleaves JT: A general formula for reactant conversion
over a single catalyst particle in TAP pulse experiments, Chem Eng Sci 64(21):4358–4364,
2009a.
Feres R, Yablonsky GS, Mueller A, Baernstein A, Zheng X, Gleaves JT: Probabilistic analysis
of transport-reaction processes over catalytic particles: theory and experimental testing,
Chem Eng Sci 64(3):568–581, 2009b. https://doi.org/10.1016/j.ces.2008.09.033.
Frattini D, Falcucci G, Minutillo M, Ferone C, Cioffi R, Jannelli E: On the effect of different
configurations in air-cathode MFCs fed by composite food waste for energy harvesting,
Chem Eng Trans 49:85–90, 2016. https://doi.org/10.3303/CET1649015.
Gambino E, Toscanesi M, Del Prete F, et al: Polycyclic aromatic hydrocarbons (PAHs)
degradation and detoxification of water environment in single-chamber air-cathode
microbial fuel cells (MFCs), Fuel Cells 17(5):618–626, 2017.
Gordon RG, Hausmann D, Kim E, Shepard J: A kinetic model for step coverage by atomic
layer deposition in narrow holes or trenches, Chem Vap Depos 9(2):73–78, 2003.
Hartline RM, Call DF: Substrate and electrode potential affect electrotrophic activity of
inverted bioanodes, Bioelectrochemistry 110:13–18, 2016.
Jannelli N, Nastro RA, Cigolotti V, Minutillo M, Falcucci G: Low pH, high salinity: too
much for microbial fuel cells? Appl Energy 192:543–550, 2017.
Jannelli E, Di Trolio P, Flagiello F, Minutillo M: Development and performance analysis of
biowaste based microbial fuel cells fabricated employing additive manufacturing
technologies, Energy Procedia 148:1135–1142, 2018.
Keil FJ: Multiscale modelling in computational heterogeneous catalysis. In Kirchner B,
Vrabec J, editors: Multiscale molecular methods in applied chemistry, Berlin, Heidelberg,
2012, Springer Berlin Heidelberg, pp 69–107. ISBN: 978-3-642-24968-6. https://
doi.org/10.1007/128_2011_128.
Kickelbick G: Hybrid materials: synthesis, characterization, and applications, 2007, John Wiley &
Sons.
LBM for 2D and 3D chemical reactors 139

Kim JR, Jung SH, Regan JM, Logan BE: Electricity generation and microbial community
analysis of alcohol powered microbial fuel cells, Bioresource Technol 98(13):2568–2577,
2007.
Krastev V, Falcucci G: Simulating engineering flows through complex porous media via the
lattice Boltzmann method, Energies 11(4):715, 2018.
Krastev VK, Falcucci G: Evaluating the electrochemical and power performances of microbial
fuel cells across physical scales: a novel numerical approach, Int J Hydrogen Energy
44(9):4468–4475, 2019. https://doi.org/10.1016/j.ijhydene.2018.11.226. European Fuel
Cell Conference & Exhibition 2017.
Krastev VK, Amati G, Jannelli E, Falcucci G: Direct numerical simulation of SCR reactors through
kinetic approach, 2016a, SAE Technical Paper. Technical Report 2016-01-0963.
Krastev VK, Russo S, Verdemare D, Recine G, Biferale L, Falcucci G: CFD aided optimi-
zation of an innovative SCR catalyst design for heavy-duty marine diesel engines. In AIP
conference proceedings, (vol 1738), 2016b, p 270017.
Krastev V, Di Ilio G, Bella G, Ubertini S, Falcucci G: Multidimensional modeling of SCR systems
via the lattice Boltzmann method, 2019, SAE Technical Paper. Technical Report 2019-24-
0048.
Kr€uger T, Kusumaatmaja H, Kuzmin A, Shardt O, Silva G, Viggen EM: The lattice Boltzmann
method, (vol 10), 2017, Springer. pp 978–973.
Levęque MA: Les lois de la transmission de chaleur par convection, Ann Mines 13:381–412,
1928.
Liu H, Cheng S, Huang L, Logan BE: Scale-up of membrane-free single-chamber microbial
fuel cells, J Power Sources 179(1):274–279, 2008.
Logan BE: Exoelectrogenic bacteria that power microbial fuel cells, Nat Rev Microbiol
7(5):375, 2009.
Logan BE, Regan JM: Electricity-producing bacterial communities in microbial fuel
cells, Trends Microbiol 14(12):512–518, 2006. https://doi.org/10.1016/j.tim.2006.
10.003.
Logan BE, Hamelers B, Rozendal R, et al: Microbial fuel cells: methodology and technol-
ogy, Environ Sci Technol 40:5181–5192, 2006.
Logan BE, Cheng S, Watson V, Estadt G: Graphite fiber brush anodes for increased
power production in air-cathode microbial fuel cells, Environ Sci Technol 41:
3341–3346, 2007.
Logan BE, Wallack MJ, Kim KY, He W, Feng Y, Saikaly PE: Assessment of microbial
fuel cell configurations and power densities, Environ Sci Technol Lett 2(8):
206–214, 2015.
Ma M, Sokolov IM, Wang W, Filippov AE, Zheng Q, Urbakh M: Diffusion through
bifurcations in oscillating nano-and microscale contacts: fundamentals and applications,
Phys Rev X 5(3):031020, 2015.
Machado R: Numerical simulations of surface reaction in porous media with lattice
Boltzmann, Chem Eng Sci 69(1):628–643, 2012.
MacMinn CW, Dufresne ER, Wettlaufer JS: Fluid-driven deformation of a soft granular
material, Phys Rev X 5(1):011020, 2015.
Madia GS: Measures to enhance the NOx conversion in urea-SCR systems for automotive
applications (Ph.D. thesis), 2002, ETH-Z€ urich.
Marin GB, Yablonsky GS, Constales D: Kinetics of chemical reactions: decoding complexity, 2019,
Wiley-VCH.
Michaelides EES: Nanofluidics: thermodynamic and transport properties, 2014, Springer. https://
doi.org/10.1007/978-3-319-05621-0. 9783319056203 (print); 9783319056210 (online).
Minutillo M, Flagiello F, Nastro RA, Di Trolio P, Jannelli E, Perna A: Performance of two
different types of cathodes in microbial fuel cells for power generation from renewable
sources, Energy Procedia 148:1129–1134, 2018.
140 Giacomo Falcucci

Montemore MM, Montessori A, Succi S, et al: Effect of nanoscale flows on the surface struc-
ture of nanoporous catalysts, J Chem Phys 146(21):214703, 2017.
Montessori A, Falcucci G: Lattice Boltzmann modeling of complex flows for engineering applications,
2018, Morgan & Claypool Publishers. ISBN: 978-1-6817-4672-22053–2571. https://
doi.org/10.1088/978-1-6817-4672-2.
Montessori A, Prestininzi P, La Rocca M, Succi S: Lattice Boltzmann approach for complex
nonequilibrium flows, Phys Rev E 92(4):043308, 2015.
Montessori A, Prestininzi P, La Rocca M, Falcucci G, Succi S, Kaxiras E: Effects of Knudsen
diffusivity on the effective reactivity of nanoporous catalyst media, J Comput Sci
17:377–383, 2016.
Montessori A, Amadei CA, Falcucci G, Sega M, Vecitis CD, Succi S: Extended friction
elucidates the breakdown of fast water transport in graphene oxide membranes,
Europhys Lett) 116(5):54002, 2017.
Nastro RA, Falcucci G, Minutillo M, Jannelli E: Microbial fuel cells in solid waste valoriza-
tion: trends and applications. In Sengupta D, Agrahari S, editors: Modelling trends in solid
and hazardous waste management, Singapore, 2017a, Springer Nature Singapore,
pp 159–171.
Nastro RA, Jannelli N, Minutillo M, et al: Performance evaluation of microbial fuel cells fed
by solid organic waste: parametric comparison between three generations, Energy Procedia
105:1102–1108, 2017b.
Niu XD, Hyodo SA, Munekata T, Suga K: Kinetic lattice Boltzmann method for microscale
gas flows: issues on boundary condition, relaxation time, and regularization, Phys Rev E
76(3):036711, 2007.
Oh SE, Logan BE: Voltage reversal during microbial fuel cell stack operation, J Power Sources
167(1):11–17, 2007.
Ott AW, Klaus JW, Johnson JM, George SM: Al3O3 thin film growth on Si(100) using binary
reaction sequence chemistry, Thin Solid Films 292(1–2):135–144, 1997.
Puig S, Serra M, Coma M, Cabre M, Balaguer MD, Colprim J: Effect of pH on nutrient
dynamics and electricity production using microbial fuel cells, Bioresource Technol
101(24):9594–9599, 2010. https://doi.org/10.1016/j.biortech.2010.07.082.
Qian Y, D’Humières D, Lallemand P: Lattice BGK models for Navier-Stokes equation,
Europhys Lett 17(6):479–484, 1992.
Rabaey K, Verstraete W: Microbial fuel cells: novel biotechnology for energy generation,
Trends Biotechnol 23(6):291–298, 2005.
R€ohe S, Frank K, Schaefer A, et al: CO oxidation on nanoporous gold: a combined TPD and
XPS study of active catalysts, Surf Sci 609:106–112, 2013.
Salciccioli M, Stamatakis M, Caratzoulas S, Vlachos DG: A review of multiscale modeling of
metal-catalyzed reactions: mechanism development for complexity and emergent
behavior, Chem Eng Sci 66(19):4319–4355, 2011.
Santoro C, Ieropoulos I, Greenman J, et al: Current generation in membraneless single
chamber microbial fuel cells (MFCs) treating urine, J Power Sources 238:190–196, 2013.
Shekhtman SO, Yablonsky GS, Chen S, Gleaves JT: Thin-zone tap–reactor–theory and
application, Chem Eng Sci 54(20):4371–4378, 1999.
Shekhtman SO, Yablonsky GS, Gleaves JT, Fushimi R: “State defining experiment” in
chemical kinetics–primary characterization of catalyst activity in a TAP experiment,
Chem Eng Sci 58(21):4843–4859, 2003.
Succi S: The lattice Boltzmann equation for fluid dynamics and beyond, Oxford, 2001, Clarendon.
Succi S: Mesoscopic modeling of slip motion at fluid-solid interfaces with heterogeneous
catalysis, Phys Rev Lett 89(6):064502, 2002.
Succi S: Lattice Boltzmann across scales: from turbulence to DNA translocation, Eur Phys J B
64(3–4):471–479, 2008.
LBM for 2D and 3D chemical reactors 141

Succi S, Smith G, Kaxiras E: Lattice Boltzmann simulation of reactive microflows over


catalytic surfaces, J Stat Phys 107(1–2):343–366, 2002.
Succi S, Amati G, Bernaschi M, Falcucci G, Lauricella M, Montessori A: Towards exascale
lattice Boltzmann computing, Comput Fluids 181:107–115, 2019.
Weaver P, Jansen L, Van Grootveld G, Van Spiegel E, Vergragt P: Sustainable technology
development, 2017, Routledge.
Wittstock A, Zielasek V, Biener J, Friend CM, B€aumer M: Nanoporous gold catalysts for
selective gas-phase oxidative coupling of methanol at low temperature, Science
327(5963):319–322, 2010.
Xu B, Haubrich J, Baker TA, Kaxiras E, Friend CM: Theoretical study of O-assisted selective
coupling of methanol on Au (111), J Phys Chem C 115(9):3703–3708, 2011.
Zhou L, Qu ZG, Chen L, Tao WQ: Lattice Boltzmann simulation of gas-solid adsorption
processes at pore scale level, J Comput Phys 300:800–813, 2015.
Zwietering MH, Jongenburger I, Rombouts FM, Van’t Riet K: Modeling of the bacterial
growth curve, Appl Environ Microbiol 56(6):1875–1881, 1990.

You might also like