You are on page 1of 18

Microfluid Nanofluid (2016) 20:162

DOI 10.1007/s10404-016-1829-8

RESEARCH PAPER

A study on pressure‑driven gas transport in porous media:


from nanoscale to microscale
Yoshiaki Kawagoe1 · Tomoya Oshima1 · Ko Tomarikawa1 · Takashi Tokumasu2 ·
Tetsuya Koido3 · Shigeru Yonemura2   

Received: 25 April 2016 / Accepted: 8 November 2016


© The Author(s) 2016. This article is published with open access at Springerlink.com

Abstract  Gas flow in porous media can be seen in various with micro-/nanoscale pores. Using the obtained results,
engineering devices such as catalytic converters and fuel we constructed expressions to estimate the pressure-driven
cells. It is important to understand transport phenomena in gas transport in porous media with micro-/nanoscale pores
porous media for improvement of the performance of such and porosity ranging from 0.3 to 0.5. The flow velocities
devices. Porous media with pores as small as the mean free estimated by using the constructed expressions agreed well
path of gas molecules are used in such devices as proton with those obtained in the DSMC simulations.
exchange membrane fuel cells. It is difficult to measure
molecular transport through such small pores in the experi- Keywords  Porous media · Pressure-driven gas transport ·
mental approach. In addition, even when using theoretical Direct simulation Monte Carlo method · High Knudsen
or numerical approaches, gas flow through nanoscale pores number flow · Rarefied gas dynamics
must be treated by the Boltzmann equation rather than the
Navier–Stokes equations because it cannot be considered
as a continuum. Thus, conventional analyses based on the 1 Introduction
continuum hypothesis are inadequate and the transport phe-
nomena in porous media with nanoscale pores are not yet Gas flow in porous media can be observed in various engi-
clearly understood. In this study, we represented porous neering devices such as catalytic converters and fuel cells.
media by randomly arranged solid spherical particles and Because of their large effective surface area, porous media
simulated pressure-driven gas flow through the porous are used to facilitate surface reactions such as dissociation
media by using the direct simulation Monte Carlo (DSMC) of H2 in electrodes of proton exchange membrane fuel cells
method based on the Boltzmann equation. DSMC simula- (PEMFCs). In order to improve the performance of such
tions were performed for different porosities and different devices, it is important to understand transport phenomena
sizes of solid particles of porous media. It was confirmed in porous media.
that Darcy’s law holds even in the case of porous media Gas flow through a porous medium is usually ana-
lyzed by using Darcy’s law (1856), which states that the
discharge rate through a porous medium is proportional
to the pressure gradient ∇p and the permeability K and is
* Shigeru Yonemura inversely proportional to the coefficient of viscosity µ of
yonemura@ifs.tohoku.ac.jp gas. Darcy’s law can be written as
1
Department of Nanomechanics, Graduate School K
of Engineering, Tohoku University, 6‑6‑01 Aoba, Aramaki, Us = − ∇p, (1)
µ
Aoba‑ku, Sendai 980‑8579, Japan
2
Institute of Fluid Science, Tohoku University, 2‑1‑1 Katahira, where p is the gas pressure and Us is the superficial veloc-
Aoba‑ku, Sendai 980‑8577, Japan ity, which is defined as the volume flow rate through a
3
Honda R&D Co., Ltd., Automobile R&D Center, Advanced unit cross-sectional area of the solid plus fluid of a porous
Research Division, 1‑4‑1 Chuo‑ku, Wako 351‑0193, Japan medium (Bird et al. 2007). The superficial velocity Us is

13
162   Page 2 of 18 Microfluid Nanofluid (2016) 20:162

not the velocity of the gas traveling through the pores. The frequently than with other gas molecules, and thus, the effect
average velocity U of the gas traveling through the pores of molecule–molecule collisions is negligible. Due to the
has the following relation with the superficial velocity Us: diffuse reflections at the pore wall, gas molecules diffuse
gradually. This phenomenon is called “Knudsen diffusion”
Us
U= , (2) and the pressure-driven gas transport in such a high Knudsen
ε number system is governed by Knudsen diffusion. Knudsen
where ε is the porosity of the porous medium. In order to (1909) obtained the expression for Knudsen diffusivity in a
estimate gas flow rate through a porous medium, evaluating cylindrical tube. Evans et al. (1980) investigated effective
its permeability correctly is important. In previous studies, Knudsen diffusivity in porous media, which are represented
the following empirical models for permeability have been by packing solid spherical particles, by performing Monte
proposed: Carlo simulations of molecular motion in porous media.
Tomadakis and Sotirchos (1991) also investigated the effec-
dp2 ε3 tive Knudsen diffusivity by performing Monte Carlo simula-
KB−K = , (3)
150 (1 − ε)2 tions, porous media being represented by randomly arranged
solid fibers, which can interpenetrate one another.
dp2 ε3 In the case where the pore size is comparable with the
KC−K = , (4)
180 (1 − ε)2 mean free path of gas molecules, gas flow in the porous
medium is in the slip or transition flow regime and both
dp2 ε5.5 the contributions of Knudsen diffusion and viscous flow
KR−G = , (5) appear. Knudsen (1909) found that as the averaged pressure
5.6 1.05
is reduced, the gaseous mass flux through capillaries for a
where dp is the effective average diameter of packed par- given pressure drop decreases at first due to the decrease of
ticles of the porous medium, KB−K, KC−K, and KR−G are gas density, but reaches a minimum value when the Knud-
the Blake–Kozeny model (Blake 1922; Kozeny 1927), sen number is about unity, and then starts to increase. This
the Carman–Kozeny model (Carman 1937, 1938, 1956; is caused by the appearance of velocity slip at the capil-
Kozeny 1927), and the Rumpf–Gupte model (Rumpf and lary wall and Knudsen diffusion. A similar phenomenon
Gupte 1971), respectively. The Blake–Kozeny and the Car- will appear in porous media. As the average pressure p̄
man–Kozeny models were derived by regarding a porous decreases, the gas permeability K of a porous medium will
medium as a bundle of capillary tubes and by considering increase and deviate from its liquid permeability K0, which
laminar flow in tubes (Dullien 1979; Bird et al. 2007). The is equal to the gas permeability K in the continuum limit.
Rumpf–Gupte model was derived by finding the relation- Klinkenberg (1941) derived the following expression for
ship between the friction factor of the porous medium and the effective gas permeability.
the porosity by dimensional analysis (Dullien 1979). The  
b
constants in these models were determined to fit experi- K = K0 1 + , (6)

mental data on packed beds. In the derivation of these mod-
els, the effect of the mean free path of molecules in gase- where b is called the Klinkenberg coefficient. For a large
ous flow was neglected. The pores of porous media used pressure, the gas permeability K is equal to the liquid per-
in the above-mentioned devices may be in the micro-/ meability K0. This phenomenon is called the “Klinkenberg
nanoscale range, e.g., those of porous media used for elec- effect” and is induced by the appearance of velocity slip
trodes of PEMFCs may be as small as the mean free path at the pore walls and Knudsen diffusion. For low perme-
of gas molecules. The Knudsen number of gas flow in such ability porous media, Jones (1972) found that the Klinken-
porous media is on the order of unity, and hence, the gas is berg coefficient b and the liquid permeability K0 have the
in nonequilibrium because of a lack of intermolecular colli- relation, b ∝ K0−0.36. Bravo (2007) developed the general-
sions. Such kind of gas flow is called “high Knudsen num- ized model of the Klinkenberg coefficient b for the whole
ber flow” and cannot be treated as a continuum. Since the range of the Knudsen number. In his model, the Klinken-
effect of the mean free path was neglected in the derivation berg coefficient b is a function of the Knudsen number. It
of the above-mentioned permeability models, these models explains the pressure dependence of the Klinkenberg coef-
are not applicable to gas flow in such porous media with ficient b as was seen in experiments, whereas in earlier
micro-/nanoscale pores. models the Klinkenberg coefficient b was considered inde-
In the case where the pore size of porous media is much pendent of pressure.
smaller than the mean free path of gas molecules, i.e., in For transition flow in porous media, Mason et al. (1967)
the case where the Knudsen number is much larger than proposed the dusty gas model, in which the solid body
unity, gas molecules collide with the pore walls much more of a porous medium is regarded as large molecules fixed

13
Microfluid Nanofluid (2016) 20:162 Page 3 of 18  162

in space and these imaginary molecules are considered


to constitute a dusty gas component of the gas mixture.
In the dusty gas model, the mass flux of one gas compo-
nent is given by the sum of diffusive flux and viscous flux,
where the diffusive flux is the combined flux of molecular
diffusion and Knudsen diffusion. Beskok and Karniadakis
(1999) proposed the expression
1
µe = µ0 (7)
1 + a Kn
Fig. 1  Porous medium represented by randomly arranged solid
for the effective viscosity µe in the transition flow regime
spherical particles
and validated it by comparing with the linearized Boltz-
mann solution and with the results obtained by the direct
simulation Monte Carlo (DSMC) method (Bird 1994) (see among neighboring particles, it was allowed in the pre-
also Karniadakis et al. 2005). Here, µ0 is the viscosity in sent study. Steady gas flows in the porous media driven by
the continuum limit and Kn is the Knudsen number given a steady pressure gradient in the y direction were studied.
by Kn = mol /L, mol is the mean free path of gas mole- The surface reaction was neglected here. The lengths of the
cules, and L is the characteristic length of the flow. Micha- computational domain in the x and z directions were set to
lis et al. (2010) performed DSMC simulations of hard be 3 times as long as the diameter dp of the solid particles
sphere gas flow between parallel plates and found that the and that in the y direction was set to be 30 times as long as
numerical factor a has the value of 2 for length-to-width dp. The centers of spherical particles with the diameter dp
ratio of 20 over the entire transition flow regime. Kalara- were placed randomly in the region of 1.5dp < y < 28.5dp.
kis et al. (2012) incorporated the effect of rarefaction on In the present simulation, the diameter dp of the solid parti-
the viscosity, Eq. (7), into the dusty gas model and derived cles is varied from 10 to 2000 nm.
the expression of the effective permeability with the form The gas treated here was H2. The hard sphere model
of the Klinkenberg equation, Eq. (6). They showed that the (Bird 1994) for intermolecular collisions was used in the
permeability of a porous medium obtained in the gas flow present simulation. H2 molecules were treated as hard
simulations by the lattice Boltzmann method incorporat- spheres with diameters of dm = 0.274 nm and masses
ing the effective gas viscosity in the transition flow regime of m = 3.3476 × 10−27 kg. The molecular diameter dm
agrees well with that obtained by the DSMC simulations. was determined from the value of the viscosity µ of H2 ,
Dreyer et al. (2014) investigated the gas diffusion pro- µ = 8.8 × 10−6 Pa s, in the case of the temperature of
cess in high-porosity nanoscale porous media by using the T = 293 K and the pressure of p = 101, 325√ Pa by using the
DSMC method and compared the DSMC results with those relations µ = (1/2)ρ C̄mol and mol = 1/ 2nπdm 2 , where

of the dusty gas model. Christou and Dadzie (2016) inves- ρ is the gas density and given by ρ = mn, n is the number
tigated the effect of the Knudsen number for the velocity density of molecules, C̄ is the mean molecular speed and

profile in a Berea porous structure, which was obtained given by C̄ = 8kB T /πm, and kB is the Boltzmann con-
with micro-CT, by performing DSMC simulations. stant. The mean free path mol for intermolecular collisions
The objectives of the present study are to investigate is given by mol = 119.3 nm in the case of H2 with the tem-
governing factors of molecular transport due to the pressure perature of T = 293 K and the pressure of p = 101, 325 Pa.
gradient through porous media with pores ranging from In the case where the diameter dp of solid spherical
nanometer to micrometer by performing numerical simu- particles is as small as the mean free path mol, the pore
lations of gas flow through porous media and to construct size will also be as small as the mean free path mol. Since
an expression to estimate pressure-driven gas transport the Knudsen number of gas flow through such porous
through such porous media by using the present simulation media will be on the order of unity, such high Knudsen
results and by considering the effects of Knudsen diffusion number flow cannot be treated as a continuum, and the
and viscous flow with a velocity slip. governing equation of the gas flow is the Boltzmann equa-
tion rather than the Navier–Stokes equations. In order to
analyze high Knudsen number gas flows such as rarefied
2 Numerical method gas flows or micro-/nanoscale gas flows, several methods
have been developed, such as the DSMC method (Bird
In this study, porous media were represented by randomly 1994), the lattice Boltzmann method with the Bhatnagar–
arranging solid spherical particles as shown in Fig. 1. Gross–Krook (BGK) collision operator (Nie et al. 2002;
Although random arrangement causes interpenetration Zhang et al. 2005; Guo and Zheng 2008), the linearized

13
162   Page 4 of 18 Microfluid Nanofluid (2016) 20:162

Boltzmann equation (Ohwada et al. 1989; Doi 2010), and (a)


the information preservation (IP) scheme (Fan and Shen 40
Re=0.12,

Flow velocity U obtained in DSMC [m/s]


ε=0.5, dp=10 nm
1999, 2001). In the present work, the DSMC method, Kn=0.40
35 ε=0.3, dp=100 nm
which is the stochastic solution of the Boltzmann equa-
ε=0.5, dp=100 nm
tion (Nanbu 1980), was used for numerical simulations of 30
ε=0.8, dp=100 nm
gas flows in porous media. The DSMC method has been
25 ε=0.5, dp=1000 nm
used to simulate gas flow in micro-/nanoscale porous
media (Saito et al. 1995; Tomarikawa et al. 2011; Kala- Re=0.16,
20
rakis et al. 2012; Oshima et al. 2012; Dreyer et al. 2014; Kn=0.13
Christou and Dadzie 2016). 15
In the present DSMC simulation, the computational Re=0.0086, Kn=1.3
domain was divided into cells, the length of which was 10
about the same or smaller than the mean free path mol of 5
gas molecules. The motions and collisions of molecules
were traced, and hence, time evolution of the flow field 0
0 5 10 15 20 25 30 35 40
was simulated. Intermolecular collisions in the same cell
were calculated stochastically using the maximum colli- Pressure gradient -dp/dy [Pa/nm]
sion number method (Nanbu 1992, 2000). Except for the (b)
cases treated in Fig. 2, the gas pressure at the inlet was 4

Flow velocity U obtained in DSMC [m/s]


ε=0.5, dp=10 nm Re=0.0019, Kn=2.2
set at 1 atm and the mean free path at the inlet, mol,inlet, 3.5 ε=0.3, dp=100 nm
was 119.3 nm. In the cases of dp = 10–500 nm, the com-
ε=0.5, dp=100 nm
putational domain with the size 3dp × 30dp × 3dp was 3
ε=0.8, dp=100 nm
divided into 10 × 100 × 10 cubic cells. In those cases, the
2.5 ε=0.5, dp=1000 nm
cell size lcell was given by 0.3dp (=3–150 nm), which was
about the same or smaller than the mean free path. In the 2
cases of dp = 1000 nm and dp = 2000 nm, the computa-
tional domain was divided into 15 × 150 × 15 cubic cells 1.5
Re=0.00085, Kn=13
and 30 × 300 × 30 cubic cells, respectively. In those cases,
1
the cell size lcell was 200 nm. It is slightly larger than the
mean free path but much smaller than the particle diameter 0.5
dp. Therefore, this choice of the cell size is also reasonable
to describe gas flow through the porous medium, which is 0
0 5 10 15 20 25 30 35 40
represented by the arrangement of particles. Since gas mol- Pressure gradient -dp/dy [Pa/nm]
ecules can flow only through a void region in the porous
medium, the volume of a void region in each cell needs to
Fig. 2  Relation between the flow velocity U and the pressure
be calculated for intermolecular collision judgment in the gradient. U is plotted in the ranges of a 0 ≤ U ≤ 40 m/s and b
DSMC simulations. In order to calculate it, we considered 0 ≤ U ≤ 4 m/s. ( p̄ = (pin + pout )/2 = 0.95 atm)
grid points, which were equally spaced in the x, y, and z
directions in each cell. The spacing between grid points
in x, y, and z directions was 10 times smaller than the cell and the surface temperature of solid spherical particles
size. The effective volume Vcell of a cell, i.e., the volume of was also set at 293 K. The gas pressures at the inlet and
the void region in the cell, is given by the outlet were set in accordance with the condition of
  each case considered in the present simulations. How-
Nsolid
Vcell = (lcell )3 1 − , (8) ever, the pressure at the outlet was always set to be lower
Ngrid
than that at the inlet in order to generate a negative pres-
where Nsolid is the number of grid points in the solid region sure gradient of dp/dy. The periodic boundary condition
in the cell and Ngrid is the total number of grid points in the was applied for the boundaries in both the x and z direc-
cell and was equal to 1000 in the present simulations. tions. In the present simulation, the arrangement of solid
The post-collisional velocities of molecules impinging spheres was also considered to be repeated periodically
on the solid wall were determined by using the diffuse in both the x and z directions. Therefore, if a molecule
reflection model with wall temperature. The accommoda- goes out of the computational domain through a point
tion coefficient on the wall was set at unity. The gas tem- in a void region on one periodic boundary, it will nec-
peratures at the inlet and at the outlet were set at 293 K, essarily re-enter the computational domain through the

13
Microfluid Nanofluid (2016) 20:162 Page 5 of 18  162

corresponding point in the void region on the opposite the values (J+ − J− )/n of the two preceding time steps as
boundary. follows:
The time step was set at ∆t = αt τmol,inlet, where τmol,inlet i−1 i−2 i−1 i−2
is the mean free time for intermolecular collisions at J+ + J+ − J− − J−
i
UB.C. = , (10)
the inlet. In all cases considered here, the pressure at the 2n
inlet is the highest in the computational domain. There- where the superscripts i, i − 1, and i − 2 indicate the values
fore, the mean free time at the inlet will be the small- at the ith, (i − 1)th, and (i − 2)th time steps, respectively.
est in the domain. This is the reason why we adopted the Saito et al. (1995) improved this method and proposed
mean free time at the inlet, τmol,inlet, as the reference time the following technique to adjust the tentative boundary
length. The coefficient αt was set so that the time step condition:
∆t was much smaller than the mean free time for inter-
i−1
molecular collisions, i.e., αt ≪ 1, and the moving dis- i i−1 α   k k

tance of molecules would be much smaller than the cell
UB.C. = (1 − α)UB.C. + J+ − J− , (11)
3n
k=i−3
size lcell, i.e., C̄∆t ≪ lcell. Since ∆t = αt τmol,inlet and
τmol,inlet = mol,inlet /C̄, the latter condition for αt results in where α is the relaxation coefficient. In the present simula-
αt ≪ lcell /mol,inlet. Since mol,inlet = 119.3 nm and lcell was tions, we adopted Eq. (11) to adjust the boundary condition
set at 0.3dp for dp = 10–500 nm and was set at 200 nm for of the flow velocity at both the inlet and the outlet bounda-
dp = 1000 and 2000 nm, αt was set at 0.01, 0.015, 0.02, ries. The relaxation coefficient α was set at 0.01.
0.05, 0.05, 0.1, 0.1, and 0.1 for dp = 10, 20, 50, 100, 200, After the gas flow through the porous medium had
500, 1000, and 2000 nm, respectively. In the case where reached the steady state, the flow properties such as gas den-
the gas pressure at the inlet was set at 1 atm, the mean free sity, gas velocity, gas pressure, and gas temperature were
time at the inlet was τmol,inlet = 6.8 × 10−11 s. sampled in each cell. Since the flow in the vicinity of the
In the DSMC method, the velocities of incoming mole- edge of the assembly of solid particles has properties differ-
cules are determined stochastically from the velocity distri- ent from those in the bulk of porous medium which we want
bution at the corresponding boundary. Even if the velocity to know, the flow properties in the region 3dp ≤ y ≤ 27dp
distribution is assumed to be Maxwellian, the flow velocity were adopted as those in the porous medium. The porosity
at that boundary is required to determine the velocity dis- of the porous medium was also calculated from the volume
tribution. However, in the present case of pressure-driven fraction of gas in the region 3dp ≤ y ≤ 27dp.
flow, the flow velocity at the inlet or the outlet boundaries
cannot be known before the numerical simulation is per-
formed because the flow velocity is an outcome of applied 3 Results and discussion
pressure gradient and selected parameters of porous media.
Therefore, in the present study, the tentative conditions of 3.1 The validity of Darcy’s law in the case of a
flow velocities at these boundaries were given first, and the nanoscale porous medium
velocities of incoming molecules from these boundaries
were determined by using these conditions. The result- In the present study, we defined the flow velocity U as
ant flow velocity UB at the inlet or the outlet boundaries is the average velocity of gas molecules traveling through a
given by porous medium as follows:
J+ − J− Nsmp N (k) (k) 
UB = , (9) c
n k=1 j=1 y,j
U= Nsmp , (12)
where J+ and J− are the molecular number fluxes through N (k)
k=1
that boundary in the +y direction and in the −y direction,
(k)
respectively, and n is the number density of molecules at where cy,j is the y component of the velocity of the
that boundary. After commencement of the numerical jth simulated molecule within the sampling region
simulation, the boundary condition of the flow velocity at (3dp ≤ y ≤ 27dp ) at the kth sampled time step, N (k) is the
that boundary has to be frequently adjusted to correspond number of simulated molecules in the sampling region at
to UB . However, if UB obtained at the (i − 1)th step is used the kth sampled time step, and Nsmp is the number of sam-
as the boundary condition for the ith step, the solution will pling operations. The pressure gradient dp/dy is calculated
not easily become stable. Ikegawa and Kobayashi (1988) by dividing the pressure difference between the inlet and
proposed a technique to adjust the tentative boundary con- the outlet by the length of the region, dp ≤ y ≤ 29dp, where
dition UB.C. of the flow velocity by using the average of solid particles are arranged, as follows:

13
162   Page 6 of 18 Microfluid Nanofluid (2016) 20:162

dp pout − pin collisions as the characteristic length. This is because the


= , (13)
dy 28dp pore size in a porous medium is difficult to be determined
and because the mean free path wall for molecule–wall col-
where pin and pout are the gas pressures at the inlet and the lisions is expected to represent the characteristic length of
outlet, respectively. The reason why the total length of the the inner channel of the porous medium such as the pore
computational domain, 30dp, is not used here is because size or the channel diameter. The validity of this expectation
the pressure p does not change in the regions of 0 < y < dp will be confirmed in Sect. 3.4. The mean free paths mol and
and 29dp < y < 30dp due to nonexistence of solid parti- wall are given by
cles. Figure 2 shows the effect of the pressure gradient on

the flow velocity U for various cases of particle diameter mol = , (17)
dp and porosity ε. The average pressure p̄ = (pin + pout )/2 νmol
was set at 0.95 atm for all the cases considered in Fig. 2. and
The pressures at the inlet and the outlet were set in accord-

ance with the set value of dp/dy. In Fig. 2, the flow velocity wall = , (18)
U is proportional to the pressure gradient dp/dy for all the νwall
cases considered here. From Eqs. (1) and (2), Darcy’s law respectively. Here, νmol and νwall are the molecule–molecule
with respect to the flow velocity U is given by and the molecule–wall collision frequencies, respectively.
Therefore, Eq. (16) results in
K
U=− ∇p. (14) νwall
µε Kn = .
νmol (19)
Thus, Darcy’s law states that the flow velocity U of gas
traveling through a porous medium is proportional to per- Yonemura (2008) pointed out that the Knudsen number Kn,
meability K and pressure gradient ∇p and is inversely pro- which is defined by Kn = mol /L using the characteristic
portional to the coefficient of viscosity µ and porosity ε. In length L of the flow, represents essentially the ratio of the
a set of cases with the same porosity ε and the same parti- molecule–wall collision frequency to the molecule–mol-
cle diameter dp in Fig. 2, the same arrangement of particles ecule collision frequency, νwall /νmol, although the value of
was used, and hence, the Knudsen number and the permea- νwall /νmol may differ from Kn to some extent in most cases.
bility are also the same among those cases. Since the gas is For instance, in the case of gas flow through the cylindri-
common to all the cases considered here, µ is also common cal tube of diameter L, the value of νwall /νmol is equal to
to those. Consequently, Fig. 2 shows that the flow velocity Kn, while in the case of gas flow between two parallel plates
U is proportional to pressure gradient dp/dy under the con- which are separated by a distance L, that is equal to 0.5 Kn,
ditions of constant porosity ε, constant permeability K, and and furthermore in the case of gas flow in the spherical
constant viscosity µ. Namely, Darcy’s law holds for all the region of diameter L, that is equal to 1.5 Kn. Thus, the value
cases considered here. of νwall /νmol is always proportional to Kn and is not largely
The Reynolds numbers and the Knudsen numbers for the different from Kn. Therefore, we can say that the physical
cases of the highest flow velocity among the cases with the meaning of νwall /νmol is essentially the same with that of
same porosity ε and the same particle diameter dp are also Kn. The definition of the Knudsen number in the present
shown in Fig. 2. In the present study, the Reynolds number study, Eq. (16), comes from such a consideration. Here,
Re and the Knudsen number Kn are defined as wall was obtained by using νwall, which was counted in the
present DSMC simulations. The mean free path mol for
ρUwall
Re = , (15) molecule–molecule collisions for p̄ = 0.95 atm is 125.5 nm.
µ
Thus, from Fig. 2, it is confirmed that even in the case of
and a nanoscale porous medium, i.e., even in the case of high
Knudsen number, Darcy’s law holds at least for Re < 0.1.
mol
Kn = , (16) In the following sections, we investigate the flow velocity U
wall for various porous media with various porosities and parti-
respectively. Here, mol is the mean free path for molecule– cle sizes under the condition of the fixed pressure gradient.
molecule collisions, i.e., the so-called the mean free path, Nevertheless, applying same pressure gradient for various
and wall is the mean free path for molecule–wall collisions, porous media in the present DSMC simulations is not easy.
which is defined as the average distance that a molecule For example, in the case of dp = 1000 nm, the pressure
travels between successive molecule–wall collisions. gradient dp/dy becomes −0.3619 Pa/nm under the condi-
For the definitions of the Reynolds and the Knudsen num- tions of pin = 1.0 atm and pout = 0.9 atm. In order to apply
bers, we chose the mean free path wall for molecule–wall the same pressure gradient in the case of dp = 10 nm, the

13
Microfluid Nanofluid (2016) 20:162 Page 7 of 18  162

pressure difference has to be set at 0.001 atm, i.e., the inlet l


and the outlet pressures have to be set at pin = 0.9505 atm
ε
and pout = 0.9495 atm, respectively, because the length
of the computational domain is changed according to the pin pout
diameter dp. In the DSMC simulation, such a small pres- dp
sure difference might be buried in statistical fluctuations
resulting from its stochastic approach. Consequently, the dt
flow velocity cannot be evaluated accurately in such a case. Ut
lt
Therefore, we use Darcy’s law to evaluate the flow velocity
in the case of such a small pressure difference.
From this point forward, in all cases, the inlet and the U pin > pout
outlet pressures are set at pin = 1.0 atm and pout = 0.9 atm,
respectively. Since the pressure gradient dp/dy becomes
Fig. 3  Concept of the tube bundle model for a porous medium
−0.3619 Pa/nm in the case of dp = 1000 nm, dp/dy for
other cases of different particle size dp [nm] can be given
by −0.3619(1000/dp [nm]) Pa/nm. The flow velocity U for molecular concentration. The chaotic molecular motions
dp/dy = −0.3619 Pa/nm was estimated by multiplying the after molecule–wall collisions induce a net molecular flux
obtained flow velocity by dp [nm]/1000. Such conversion from the more concentrated region to the less concentrated
of the flow velocity U was performed under the condition one. This phenomenon is called Knudsen diffusion. In the
of a very low Reynolds number, Re < 0.1. case of a cylindrical capillary tube with radius R, the Knud-
The smallest value of flow velocity before the conver- sen diffusion coefficient Dcylinder
K is given by (Knudsen
sion was U = 0.37 m/s, which was obtained in the case of 1909; Cunningham and Williams 1980)
ε = 0.3 and dp = 10 nm. In the case of such a small flow
2
velocity, sampling was repeated a lot so that the total num- K
Dcylinder = RC̄. (20)
ber of sampled molecules 3
The molar flux Ncylinder
K due to the Knudsen diffusion under
Nsmp
 the pressure gradient is given by (Cunningham and Wil-
Ntotal = N (k)
liams 1980)
k=1
K K 1
becomes large enough. In the case of ε = 0.3 and Ncylinder = −Dcylinder ∇p, (21)
Rg T
dp = 10 nm, sampling was repeated one million times and
1,609,095,123 simulated molecules were sampled. The where Rg is the universal gas constant. Dividing Eq. (21)
magnitude of fluctuation√of calculated flow velocity U can by the molar concentration c, the flow velocity Ucylinder
K in a

be estimated to be (1/ Ntotal ) kB T /m, and hence, the capillary tube is given by
estimated magnitude of fluctuation of U is 0.0274 m/s, and
K K ∇p
consequently, the error is 7.4% in this case. This is the case Ucylinder = −Dcylinder , (22)
p
of the maximum error. In the case of larger flow velocity,
the error is smaller. where the relation Ncylinder
K K
= cUcylinder and the equation of
state of an ideal gas, p = cRg T , are used.
3.2 Knudsen diffusion in porous media Let us regard a porous medium as a bundle of tortuous
capillary tubes, as shown in Fig. 3. Here, dt is the diam-
In order to construct an expression to estimate molecular eter of the capillary tubes, lt is their tortuous length, l is the
transport through porous media, we would like to start by straight-line length of the porous medium, ε is the poros-
considering the case in which the size of pores is much ity of the porous medium, and U is the flow velocity in the
smaller than the mean free path mol for molecule–mol- porous medium in the direction of the macroscopic pres-
ecule collisions. Since the mean free path wall for mole- sure gradient, not in the direction of the local pressure gra-
cule–wall collisions is as large as the pore size, the number dient, and Ut is the flow velocity along the center line of
of molecule–molecule collisions is negligible compared a tortuous tube. From Eqs. (20) and (22), the velocity Ut
with the number of molecule–wall collisions. Molecules induced by Knudsen diffusion and the pressure gradient is
colliding with the wall rebound in random directions. Since given by
the molecule–molecule collisions are negligible, the pres-
sure gradient does not induce molecular transport directly. 1 1 pout − pin
Ut = − dt C̄ . (23)
However, the pressure gradient induces the gradient of the 3 p̄ lt

13
162   Page 8 of 18 Microfluid Nanofluid (2016) 20:162

Considering the time required for gas to pass through the 1


ε=0.3
porous media, we have the following relation between l, lt , ε=0.4

Flow velocity U obtained in DSMC


U, and Ut: 0.8 ε=0.5
ε=0.6
lt l ε=0.7

−(1/3)λwallC( p/p)
= . (24) ε=0.8
Ut U 0.6

By substituting Eq. (24) in Eq. (23), the flow velocity U is 0.9ε (ε=0.5: )


given by 0.4

1 pout − pin l 2
 
1
U = − dt C̄ 0.2 0.9ε (ε=0.4: )
3 p̄ l lt 0.9ε (ε=0.3: )
1 1 pout − pin
= − dt C̄ τ, (25) 0
3 p̄ l 10 100 1000

where τ is “tortuosity” defined as follows: Diameter of particle dp [nm]

 2
l Fig. 4  Ratio of the flow velocity U obtained in the DSMC simula-
τ= . (26) tions to the value of Eq. (28)
lt

However, it is unclear how to determine the diameter dt of


tortuous capillary tubes because porous media have a very Eq. (28). If the tortuosity τ is equal to these ratios, the flow
complicated structure. In the case of a straight cylindrical velocity estimated by using Eq. (27) will agree with the
tube, the mean free path wall for molecule–wall collisions flow velocity obtained in the DSMC simulations. As men-
in equilibrium gas is exactly equal to the tube diameter. tioned above, the mean free path mol for molecule–mol-
Therefore, in the present work, we use wall as the tube ecule collisions for the case of p̄ = 0.95 atm is 125.5 nm.
diameter. Considering the importance of molecule–wall On the other hand, the mean free path wall for molecule–
collisions in the Knudsen diffusion, this replacement may wall collisions is 0.96dp for ε = 0.5, 0.73dp for ε = 0.4, and
be reasonable. The validity of this replacement will be dis- 0.55dp for ε = 0.3. Since Eq. (27) is for estimation of the
cussed in Sect. 3.4. Consequently, we have flow velocity for very high Knudsen numbers much larger
than 10, the tortuosity must agree with these ratios over the
1 ∇p
U = − τ wall C̄ . (27) ranges of dp < 13 nm for ε = 0.5, dp < 17 nm for ε = 0.4 ,
3 p
and dp < 23 nm for ε = 0.3. In Fig. 4, the three dashed
lines, which represent the values of 0.9ε for ε = 0.3 , 0.4,
The tortuosity τ = (l/lt )2 still remains unknown. Many and 0.5 from the bottom up, agree well with the ratios
tortuosity models for electric, hydraulic, or diffusive con- obtained for ε = 0.3(), ε = 0.4(◦), ε = 0.5(♦), respec-
ductivity have been proposed by empirical, analytical, and tively, for very small particle diameter dp. Therefore, in the
numerical approaches (Ghanbarian et al. 2013). In porous present study we adopt the tortuosity,
media with micro-/nanoscale pores, however, both contri-
butions of viscous flow and Knudsen diffusion appear, and
τ = 0.9ε, (29)
hence, the tortuosity of such porous media is unclear. Zalc for the case of low porosity, 0.3 < ε < 0.5. On the other
et al. (2004) calculated the diffusive tortuosity by using hand, the ratios obtained for the higher porosities, ε > 0.6 ,
the mean-square displacement method and test-particle are much less than 0.9ε for very small particle diame-
method. Asinari et al. (2007) also calculated the diffusive ter dp. This may be because the concept of the tube bun-
tortuosity by using the lattice Boltzmann method. In order dle model is not appropriate in the case of high porosity
to estimate the tortuosity, let us compare the flow velocities since the arrangement of solid particles is coarse. In the
obtained in the DSMC simulations with the values of present study, we consider only the cases of low porosity,
0.3 < ε < 0.5.
1 ∇p
− wall C̄ (28) In Fig. 5a, b, the flow velocities U estimated by using
3 p
Eq. (27) and τ = 0.9ε are compared with those obtained in
obtained under the same conditions. Figure 4 shows the the DSMC simulations on a log-log scale and on a linear
ratios of the flow velocities obtained in the DSMC simu- scale, respectively. For reference, the flow velocities esti-
lations to the values of Eq. (28). Here, wall’s obtained in mated by using the three empirical models, i.e., the Blake–
the DSMC simulations were used to give the values of Kozeny model given by Eq. (3), the Carman–Kozeny model

13
Microfluid Nanofluid (2016) 20:162 Page 9 of 18  162

(a) 10
DSMC simulation. This is because in the case of smaller dp,
Our expression Eq. (27)
2000
the Knudsen number becomes the same or more than unity
Blake-Kozeny Eq. (3)
Flow velocity U estimated by conventional

1 Carman-Kozeny Eq. (4) 1000 and the effect of rarefaction becomes stronger, whereas the
Rumpf-Gupte Eq. (5) 500 effect of the Knudsen number is not considered in these
200
conventional models. Therefore, these conventional mod-
models and Eq. (27) [m/s]

0.1 100
50 els are not appropriate for nanoscale porous media. On
20
dp=10 nm the other hand, the velocities estimated by using Eq. (27)
0.01
agree well with those obtained in the DSMC simulations
0.001 for dp < 200 nm, i.e., for Kn > 0.6. Thus, it is confirmed
that a good estimation of the flow velocity in porous media
0.0001 can be made by using Eq. (27) in the high Knudsen num-
ber regime of Kn > 0.6. However, for dp > 200 nm, i.e.,
10-5 for Kn < 0.6, the larger the particle diameter dp becomes,
10-5 0.0001 0.001 0.01 0.1 1 10
the more greatly the velocity estimated by using Eq. (27)
Flow velocity U obtained in DSMC [m/s]
deviates from that obtained in the DSMC simulation. This
is because as the Knudsen number becomes smaller, the
(b)
effects of molecule–molecule collisions, i.e., the character-
2.5
Our expression Eq. (27) istics of gas as a continuum, becomes stronger, whereas the
Flow velocity U estimated by Eq. (27) [m/s]

effects of molecule–molecule collisions were not consid-


dp=2000 nm,
2 dp=50 nm, Kn=2.6 ered when deriving Eq. (27). In the next section, we will
Kn=0.067
dp=20 nm, Kn=6.6
dp=10 nm, Kn=13
modify Eq. (27) by considering the characteristics of gas as
1.5 a continuum.

3.3 Construction of an expression to estimate gas flow


1 dp=1000 nm, Kn=0.13 velocity in porous media considering Knudsen
diffusion and viscous flow
0.5 dp=500 nm, Kn=0.26
dp=200 nm, Kn=0.66 As discussed in the previous section, although Eq. (27) is
dp=100 nm, Kn=1.3 valid in the case of a Knudsen number higher than unity,
0
0 0.5 1 1.5 2 2.5 3 3.5 it is not applicable in the case of a Knudsen number much
Flow velocity U obtained in DSMC [m/s] smaller than unity. Here, in order to remedy the short-
coming of our expression, i.e., Eq. (27), we will modify
Fig. 5  Comparison of flow velocities estimated by using conven- it by considering the contribution of molecule–molecule
tional models and Eq. (27) with those obtained in the DSMC simula-
collisions.
tions: a on a log–log scale; b on a linear scale (ε = 0.49, τ = 0.9ε,
dp = 10−2000 nm, p̄ = 0.95 atm, dp/dy = −0.3619 Pa/nm) In the limit of a small Knudsen number, molecules fre-
quently collide with each other and gas behaves as a contin-
uum. Under a pressure gradient of ∇p, gas is pushed by the
given by Eq. (4), and the Rumpf–Gupte model given by difference between the gas pressures on both its sides and
Eq. (5), are also shown in Fig. 5a. In Fig. 5b, the Knudsen is set in motion. If the gas moves with respect to the porous
numbers estimated by using Eq. (16) are also shown. Here, medium, momentum will be transferred from the gas to the
the porosity is set at ε = 0.49, the pressure gradient is set at porous medium due to the normal and the viscous tangen-
dp/dy = −0.3619 Pa/nm, and the diameter of solid parti- tial stresses and gas will be subjected to its reaction force.
cles dp is varied from 10 to 2000 nm. Note that, in Fig. 5a, The gas flow becomes steady when the reaction force from
the marks with the same values on the horizontal axis rep- the porous medium balances out the force exerted on the
resent the data for the same particle diameter dp. The solid gas by the pressure gradient. In the case of laminar flow in
line in each figure represents a set of points such that the a long cylindrical tube under a steady pressure gradient, the
estimated velocity is equal to the velocity obtained in the flow maintaining such a balance is called “Poiseuille flow.”
DSMC simulations. Therefore, if a mark for some models As can be inferred from “Poiseuille flow,” in the case of a
is on this solid line, it shows that the velocity estimated by Knudsen number much smaller than unity and a low Reyn-
using that model agrees with that obtained in the DSMC olds number, molecular transport under a steady pressure
simulation. In cases of conventional empirical models, the gradient is governed by the viscosity of the gas. Therefore,
smaller the particle diameter dp becomes, the more greatly such a regime is called the “viscous regime” (Cunningham
the estimated velocity deviates from that obtained in the and Williams 1980). On the other hand, in the case of a

13
162   Page 10 of 18 Microfluid Nanofluid (2016) 20:162

Knudsen number much higher than unity, molecule–mol-  1 


8 2 R

ecule collisions are negligible and the Knudsen diffusion is c1K = 2.00 ,
π µC̄
dominant. Such a regime is called the “Knudsen regime”  1   (35)
8 2 R
(Cunningham and Williams 1980). c2K = 2.47 .
In the case of gas flow in a cylindrical capillary, there π µC̄
are several expressions of molecular transport which are Wakao et al. (1965) also proposed an expression to esti-
suggested as valid for the viscous and Knudsen regimes mate the molar flux in a cylindrical capillary of radius R as
and for the entire transition regime between these two follows:
extremes. For example, Knudsen (1909) and Wakao et al.   2 
(1965) constructed an expression to estimate the molar flux K R π K 1
Ncylinder = − φDcylinder + (1 − φ) p + Dcylinder ∇p,
of gas in a cylindrical capillary by superposing the molar 8µ 4 Rg T
fluxes due to Knudsen diffusion and due to viscous flow. (36)
Knudsen (1909) proposed a semiempirical expression to where
describe the molar flux Ncylinder of gas in a cylindrical cap-
wall collision frequency
illary as follows: φ= . (37)
total collision frequency
K  1
Dividing Eq. (36) by the molar concentration c, we have

K K 1 + c1 p
Ncylinder = − a p+b ∇p, (30)
1 + c2K p Rg T   
K R2 π K ∇p
Ucylinder = − φDcylinder + (1 − φ) p + Dcylinder .
where aK, bK, c1K, and c2K are coefficients that depend on 8µ 4 p
the nature of the gas and the capillary (Cunningham and (38)
Williams 1980). The first term of Eq. (30) represents the
viscous flow, and the second term represents the Knudsen Using our definition of Knudsen number, i.e., Eq. (16), the
diffusion. Dividing Eq. (30) by the molar concentration c, ratio φ is given by
we have
νwall
φ=

1 + c1K p ∇p
 νwall + νmol
Ucylinder = − aK p + bK , (31) νwall /νmol
1 + c2K p p =
νwall /νmol + 1
where Ncylinder = cUcylinder and p = cRg T are used. (C̄/wall )/(C̄/mol )
By considering the case of large p, from comparison of
= (39)
(C̄/wall )/(C̄/mol ) + 1
Eq. (31) with the Hagen–Poiseuille law mol /wall
=
πR4 mol /wall + 1
2
π R Ucylinder =− ∇p, (32) Kn
8µ = .
Kn + 1
where R is the radius of the cylindrical capillary, it is found
that aK is given by The first term of Eq. (38) represents the Knudsen diffu-
sion, the second term represents the viscous flow, i.e., the
R2
aK = (33) Poiseuille flow, and the third term represents the contribu-
8µ tion of the velocity slip on the capillary wall. In the case
and represents the viscous flux coefficient. By consider- of Kn ≫ 1, i.e., in the case of φ ≈ 1, Eq. (38) results in
ing the case of p → 0, it is found from the comparison of Eq. (22) and represents the flow velocity only due to Knud-
Eq. (31) with Eq. (22) that the coefficient bK is the Knud- sen diffusion, whereas in the case of Kn ≪ 1, i.e., in the
sen diffusivity Dcylinder
K for a cylindrical tube as follows case of φ ≈ 0, Eq. (38) represents Poiseuille flow with a
(Cunningham and Williams 1980): velocity slip. However, since the viscous flux coefficient
R2 /8µ is proportional to R2 and Dcylinder
K is proportional to
2
bK = Dcylinder
K
= RC̄. (34) R, the contribution of the velocity slip will become negligi-
3 ble in the case of Kn ≪ 1.
The coefficients c1K and c2K are to describe the slip flow in Let us apply the above expressions to gas flow in porous
the transition regime. Knudsen obtained the values of c1K media. In Sect. 3.2, a porous medium was regarded as a
and c2K to fit his experimental data as follows:

13
Microfluid Nanofluid (2016) 20:162 Page 11 of 18  162


bundle of tortuous capillary tubes with a diameter of wall. Using p = nkB T , C̄ = 8kB T /πm, ρ = mn, µ = (1/2)
The Knudsen diffusivity for such a tube is given by ρ C̄mol, and Kn = mol /wall, Eq. (45) can be rewritten as
K 1
Dtube = wall C̄. (40)
3 2wall 128
    
32
U = −τ φ Kn + (1 − φ) 1 + a + Kn ∇p
32µ 3π 3
Therefore, Eq. (27) is given by (46)
∇p Here, let us call Eq. (42) the Knudsen-type expression for
K
U = −τ Dtube . (41) porous media and call Eq. (46) the Wakao-type expression
p
for porous media.
The coefficient Dtube K corresponds to bK in Knudsen’s In Fig. 6a, b, the flow velocities U estimated by using
expression, Eq. (31), and Dcylinder
K in Wakao’s expression, expressions proposed here and those obtained in the DSMC
Eq. (38). Supposing that the tortuosity for the viscous flow simulations are compared on a log–log scale and on a lin-
with a velocity slip is the same as the tortuosity τ for Knud- ear scale, respectively. In the present study, the numerical
sen diffusion since both flows go through the same channel factor a in the Wakao-type expression for porous media
in the porous medium, we can modify Eq. (41) by adding was set at 2 because the factor a in Eq. (44) was found by
the contribution of viscous flow with a velocity slip as in Michalis et al. (2010) to have the value of 2 in the case of a
Knudsen’s expression and Wakao’s expression. Referring gas flow of hard sphere molecules between parallel plates.
to Knudsen’s expression, Eq. (41) is modified as The cases treated in Fig. 6 are the same as those treated in
K Fig. 5. The porosity was set at ε = 0.49, the pressure gra-
(wall /2)2
 
K 1 + c1 p ∇p
U = −τ Dtube + p dient was set at dp/dy = −0.3619 Pa/nm, and the diam-
1 + c2K p 8µ p
eter of solid particles dp was varied from 10 to 2000 nm.
 
1 1 + c1K p 2wall ∇p Here, the gas viscosity µ in Eqs. (42) and (46) was given
= −τ wall C̄ + p ,
3 1 + c2K p 32µ p by µ = (1/2)ρ C̄mol. Although the flow velocity estimated
 1   (42) by using the expression Eq. (27) considering only Knud-
K 8 2 wall sen diffusion deviates from that obtained in the DSMC
c1 = 2.00 ,
π 2µC̄ simulation for a Knudsen number lower than 0.6, the flow
 1   velocities estimated by using the Knudsen-type expression,
8 2 wall
c2K = 2.47 . Eq. (42), and by using the Wakao-type expression, Eq. (46),
π 2µC̄
agree well with those obtained in the DSMC simulations
Note that the mean free path wall for molecule–wall col- over the whole range of Knudsen numbers considered here
lisions is used as the diameter 2R of the tortuous capillary due to the consideration of both contributions of Knudsen
tube. Similarly, referring to Wakao’s expression, Eq. (41) is diffusion and viscous flow with a velocity slip.
modified as In Fig. 7a, b, the flow velocities U estimated by using
   the Knudsen-type expression, Eq. (42), and by using the
1 2wall π ∇p Wakao-type expression, Eq. (46), are compared with those
U = −τ φ wall C̄ + (1 − φ) p + wall C̄ .
3 32µ 12 p obtained in the DSMC simulations for the cases of various
(43) porosities ε and various particle diameters dp on a log–log
scale and on a linear scale, respectively. The flow velocities
Since the viscosity varies depending on the Knudsen num- estimated by using both expressions proposed here agree
ber under conditions of rarefaction, let us replace the vis- well with those obtained in the DSMC simulations. This
cosity µ in Eq. (43) by the effective viscosity µe which result indicates that both expressions proposed here well
reflects the dependence on the Knudsen number. The effec- describe gas transport in porous media from nanometer to
tive viscosity µe is given by micrometer scale.
The superficial velocity is given by the Knudsen-type
1
µe = µ (44) expression for porous media as follows:
1 + a Kn
(Michalis et al. 2010; Kalarakis et al. 2012), where a is a Us = εU
 
numerical factor. Equation (43) results in 1 1 + c1K p 2wall ∇p (47)
= −ετ wall C̄ K
+ p .

1

2 π

∇p 3 1 + c2 p 32µ p
U = −τ φ wall C̄ + (1 − φ) wall p +  C̄
3 32µe 12 wall p √
   Using p = nkB T , C̄ = 8kB T /πm, ρ = mn, µ = (1/2)
1 2 π ∇p
= −τ φ wall C̄ + (1 − φ) wall (1 + a Kn)p + wall C̄ . ρ C̄mol, and Kn = mol /wall, this results in
3 32µ 12 p
(45)

13
162   Page 12 of 18 Microfluid Nanofluid (2016) 20:162

(a) dp=2000 nm, (a)


B
Knudsen-type expression B B (42)
for porous media, Eq.
10 Expression considering only Kn=0.067 ε=0.3 ε=0.4 ε=0.5
Knudsen diffusion, Eq. (27) 10 Wakao-type
Knudsen-type expression B expression forBporous media, Eq.B(46)

Flow velocity U estimated by expressions


ε=0.3 ε=0.4 ε=0.5
Flow velocity U estimated by expressions

for porous media, Eq. (42)

proposed in the present study [m/s]


Wakao-type expression
proposed in the present study [m/s]

for porous media, Eq. (46)


1
1 dp=1000 nm
dp=500 nm, Kn=0.26 dp=1000 nm,
dp=200 nm, Kn=0.66 Kn=0.13
0.1 dp=100 nm dp=500 nm
0.1
dp=100 nm, Kn=1.3
dp=50 nm, Kn=2.6
0.01 dp=20 nm, Kn=6.6
0.01 dp=50 nm
dp=10 nm, Kn=13

dp=10 nm
0.001
0.001 0.01 0.1 1 10 0.001
0.001 0.01 0.1 1 10
Flow velocity U obtained in DSMC [m/s]
Flow velocity U obtained in DSMC [m/s]
(b)
4 Expression considering only (b)
Knudsen diffusion, Eq. (27) B
Knudsen-type B porous media, Eq.
expression for B (42)
Knudsen-type expression
ε=0.3 ε=0.4 ε=0.5
Flow velocity U estimated by expressions

3.5 for porous media, Eq. (42) 1.5 Wakao-type


Wakao-type expression B expression forBporous media, Eq.B(46)
proposed in the present study [m/s]

for porous media, Eq. (46) ε=0.3 ε=0.4 ε=0.5


3 Flow velocity U estimated by expressions
proposed in the present study [m/s]
2.5
dp=1000 nm, Kn=0.13
1
2

1.5
dp=2000 nm, Kn=0.067 dp=500 nm dp=1000 nm
1
0.5
0.5 dp=500 nm, Kn=0.26
dp=200 nm, Kn=0.66
0
0 0.5 1 1.5 2 2.5 3 3.5 4
Flow velocity U obtained in DSMC [m/s] 0
0 0.5 1 1.5

Fig. 6  Comparison of flow velocities estimated by using the three Flow velocity U obtained in DSMC [m/s]
expressions proposed in the present study with those obtained
in the DSMC simulations: a on a log–log scale; b on a linear Fig. 7  Comparison of flow velocities estimated by using the Knud-
scale. (ε = 0.49, τ = 0.9ε, dp = 10−2000 nm, p̄ = 0.95 atm, sen-type expression and by using the Wakao-type expression with
dp/dy = −0.3619 Pa/nm) those obtained in the DSMC simulations for the cases of various
porosities ε: a on a log–log scale; b on a linear scale. (τ = 0.9ε,
dp = 10−1000 nm, p̄ = 0.95 atm, dp/dy = −0.3619 Pa/nm)
2wall 128 1 + c1K p ∇p
 
Us = −ετ 1+ Kn . (48)
32 3π 1 + c2K p µ
2wall
K0 = ετ . (50)
From this equation and Darcy’s law, the permeability of 32
porous medium is given by
Since the permeability can be rewritten as
128 1 + c1K p
 
128 1 + c1K p
 
K = K0 1 + Kn , (49) · Kn · p 
3π 1 + c2K p  3π 1 + c2K p
K = K0 1 + , (51)
 
p
where K0 is the permeability in the continuum limit,
 
Kn → 0 , and given by

13
Microfluid Nanofluid (2016) 20:162 Page 13 of 18  162

this has the form of the Klinkenberg equation (1941). The Thus, the Klinkenberg coefficient b for the Knudsen-type
Klinkenberg coefficient b for the Knudsen-type expression expression for porous media monotonically increases from
for porous media is given by 11.00 Kn · p to 13.58 Kn · p as Kn increases from 0 to infinity.
On the other hand, the superficial velocity is given by
128 1 + c1K p the Wakao-type expression for porous media as follows:
b= · Kn · p. (52)
3π 1 + c2K p
2
    
128 32 ∇p
Us = −ετ wall φ Kn + (1 − φ) 1 + a + Kn .
The coefficients c1K and c2K depend on the nature of the 32 3π 3 µ
gas and the material of porous medium (Cunningham (59)
and Williams 1980). In the case where the coefficients c1K From this equation and Darcy’s law, the permeability of
and c2K are given as shown in Eq. (42), using p = nkB T , porous medium is given by

C̄ = 8kB T /πm, ρ = mn, µ = (1/2)ρ C̄mol, and     
128 32
Kn = mol /wall, we have K = K0 φ Kn + (1 − φ) 1 + a + Kn . (60)
 3π 3
π 1
K
c1 p = , (53) From Eq. (39), this permeability of porous medium can
2 Kn be rewritten in the form of the Klinkenberg equation as
 follows:
π 1
c2K p = 1.235 . (54)  �
128 Kn

32

1
� 
2 Kn · + a+ −1 · Kn · p 
 3π Kn + 1 3 Kn + 1
1 +
K = K0  .
The superficial velocity Us, the permeability of porous p 
medium K, and Klinkenberg coefficient b are given by the
Knudsen-type expression for porous media as follows: (61)
Consequently, the Klinkenberg coefficient b for the Wakao-
type expression for porous media is given by

π 1
 
1+
2  128 2 Kn   ∇p ,
Us = −ετ wall 
   
1+ Kn � (55) 128 Kn 32 1
32  3π π 1  µ b= · + a+ −1 · Kn · p.
1 + 1.235 3π Kn + 1 3 Kn + 1
2 Kn
(62)
Thus, the Knudsen number also contributes to the Klinken-

π 1
 
1+ berg coefficient b for the Wakao-type expression for porous
 128 2 Kn 
1 + 3π Kn
K = K0  � , (56) media. In the case of a = 2, the coefficient b monotonically
π 1 
1 + 1.235 increases from 11.66 Kn · p to 13.58 Kn · p as Kn increases
2 Kn
from 0 to infinity.

π 1
1+ 3.4 Estimation of mean free path wall
128 2 Kn
b= ·  Kn · p, (57) for molecule–wall collisions
3π π 1
1 + 1.235
2 Kn In the previous section, the validity of the Knudsen-type
where Kn · p is given by expression, Eq. (42), and the Wakao-type expression,
Eq. (46), both of which are proposed in the present study,
mol were confirmed for gas transport in porous media with
Kn · p = · nkB T
wall micro-/nanoscale pores. However, the mean free paths wall
1 1 for molecule–wall collisions obtained from the results of the
= ·√ · nkB T (58)
wall 2nπdm 2 DSMC simulations were used in both expressions. In order
1 1 to use both expressions to estimate gas transport in porous
= ·√ · kB T , media, it is necessary to know wall without the aid of DSMC
wall 2πdm2
simulation. In this section, we consider how to estimate wall.
and hence, it depends only on the temperature in the case of Let us consider the void region of a porous medium,
hard sphere molecules since wall is constant for the same whose volume available for flow and total wetted surface
porous medium as is shown in Sect. 3.4. This suggests that area are V and S, respectively. This region is filled with a
the Knudsen number contributes to the Klinkenberg coef- gas of molecular number density n. Since the mean free
ficient b for the Knudsen-type expression for porous media. path wall for molecule–wall collisions is defined as the

13
162   Page 14 of 18 Microfluid Nanofluid (2016) 20:162

average distance that a molecule travels between successive 4ε


wall =
molecule–wall collisions, wall can be given by Sv

total distance that all molecules travel =
wall = . (63) 6
total number of molecule − wall collisions (1 − ε) (69)
dp
In the case where the gas flow velocity U is much smaller 2 ε
than the mean molecular speed C̄ of gas molecules, the num- = dp ,
31−ε
ber of molecule–wall collisions per unit time is given by
(1/4)nC̄S. On the other hand, the total distance that all mol- where N is the number of particles per unit volume of
ecules travel per unit time is given by nV C̄. Therefore, the porous medium, Vp is the volume of a particle and given by
mean free path wall for molecule–wall collisions is given by Vp = (1/6)π dp3, and Sp is the surface area of a particle and
given by Sp = πdp2. However, in the porous medium treated
nV C̄
wall = in the present DSMC simulations, particles were arranged
1 randomly and were interpenetrating one another. For such a
nC̄S
4 (64) porous medium, ε, Sv, and N are given by (Weissberg 1963;
4V
= . Weissberg and Prager 1970)
S
Dividing the numerator and the denominator of Eq. (64) by ε = e−Vp N , (70)
the total volume of the porous medium including both solid
Sv = εSp N, (71)
and void regions, wall is four times as large as the ratio of
the porosity ε to the wetted surface area Sv per unit volume 1
of porous medium as follows: N =− log ε, (72)
Vp
4V /(volume of porous medium)
wall = , respectively. Consequently, wall for the porous medium used
S/(volume of porous medium)
(65) in the present study is given by

= .
Sv 4ε
wall =
Interestingly, the mean free path wall for molecule–wall Sv
collisions, which is defined at the microscopic level, results 4
=
in the macroscopic hydraulic diameter DH for the case Sp N
when the porous medium is regarded as a conduit. Here, 4 (73)
=−
DH is given by (Dullien 1979) Sp
log ε
4 × void volume of porous medium Vp
DH = . (66) 2 dp
surface area of channels in porous medium =− .
3 log ε
Thus, if we know the porosity ε and the wetted surface area
Sv per unit volume of porous medium, we can estimate the Equations (69) and (73) show that the mean free path wall for
value of wall. molecule–wall collisions depends on the porosity ε and the
In the case where a porous medium is a packed column of diameter dp of solid particles, i.e., the parameters determining
spherical particles of diameter dp without interpenetration of the channel configuration of the porous medium, but does not
particles, ε, Sv, and wall are easily obtained by depend on the molecular number density, the gas pressure,
the gas temperature, and the nature of gas such as viscosity.
ε = 1 − Vp N
In Fig. 8, wall’s estimated by using Eq. (73) are com-
1 (67) pared with wall’s obtained in the DSMC simulations,
= 1 − πdp3 N,
6 which were used in Eq. (27), the Knudsen-type expres-
Sp N sion Eq. (42), and the Wakao-type expression Eq. (46) in
Sv = Sects. 3.2 and 3.3, for the cases of various porosities ε and
Vp N/(1 − ε)
various diameters dp of solid particles. The estimated wall
πdp2
= (1 − ε) ’s are in good agreement with those obtained in the DSMC
1 3 simulations. Since Eq. (73) is derived from Eq. (65) for the
πd (68)
6 p porous media treated in the present study, this agreement
6 shows that the estimation of wall by using Eq. (65) is valid.
= (1 − ε),
dp

13
Microfluid Nanofluid (2016) 20:162 Page 15 of 18  162

10000 (a)
ε=0.3 B
Knudsen-type B porous media, Eq.
expression for B (42)
ε=0.4 ε=0.3 ε=0.3 ε=0.3
ε=0.5 10 Wakao-type
B expression forBporous media, Eq.B(46)
ε=0.3 ε=0.3 ε=0.3
λwall obtained in DSMC [nm]

1000 ε=0.6

Flow velocity U estimated by expressions


ε=0.7

proposed in the present study [m/s]


ε=0.8
1 dp=1000 nm

100

dp=100 nm dp=500 nm
0.1
10

0.01 dp=50 nm

1
1 10 100 1000 10000 dp=10 nm
λwall estimated by using Eq. (73) [nm]
0.001
0.001 0.01 0.1 1 10

Fig. 8  Comparison of wall’s estimated by using Eq. (73) with those Flow velocity U obtained in DSMC [m/s]
obtained in the DSMC simulations for the cases of various porosi-
ties and various diameters of spherical particles (dp = 10−1000 nm, (b)
p̄ = 0.95 atm) B
Knudsen-type B porous media, Eq.
expression for B (42)
ε=0.3 ε=0.4 ε=0.5
1.5 Wakao-type
B expression forBporous media, Eq.B(46)
ε=0.3 ε=0.4 ε=0.5
As mentioned above, the mean free path wall for mol-
Flow velocity U estimated by expressions
proposed in the present study [m/s]
ecule–wall collisions depends only on the channel configu-
ration of a porous medium and is estimated to be equiva-
lent to the hydraulic diameter DH for the porous medium. 1

Therefore, our choice of wall as the characteristic length


for both Knudsen and Reynolds numbers is valid, and fur- dp=1000 nm
dp=500 nm
thermore, using wall as the tube diameter in our tube bun-
dle model for porous media is reasonable. 0.5
Lastly, let us check the validity of the flow velocities
estimated by using the Knudsen-type expression Eq. (42)
and by using the Wakao-type expression Eq. (46) together
with the mean free path wall for molecule–wall collisions 0
0 0.5 1 1.5
estimated theoretically by using Eq. (65). Actually, wall
Flow velocity U obtained in DSMC [m/s]
was estimated by using Eq. (73) here because Eq. (73) is
derived from Eq. (65) so as to be specialized for the porous
Fig. 9  Comparison of flow velocities estimated by using the Knud-
medium treated in the present study. In Fig. 9a, b, the flow sen-type expression and by using the Wakao-type expression together
velocities U estimated by using the Knudsen-type expres- with theoretically estimated wall with those obtained in the DSMC
sion Eq. (42) and by using the Wakao-type expression simulations for the cases of various porosities ε: a on a log–log scale;
Eq. (46) together with wall estimated theoretically by using b on a linear scale (τ = 0.9ε, dp = 10−1000 nm, p̄ = 0.95 atm,
dp/dy = −0.3619 Pa/nm)
Eq. (65) are compared with those obtained in the DSMC
simulations for the cases of various porosities ε and vari-
ous particle diameters dp on a log–log scale and on a linear expressions, Eqs. (46) and (59), together with Eq. (65)
scale, respectively. The flow velocities estimated by using for porous media with pores ranging from nanometer to
both expressions with theoretically estimated wall agree micrometer and porosity of 0.3 < ε < 0.5.
well with those obtained in the DSMC simulations.
Thus, if we know the porosity ε and the wetted surface
area Sv per unit volume of a porous medium, we can make 4 Conclusion
a good estimate of the gas flow velocity U and the superfi-
cial velocity Us, i.e., the volume flux, by using the Knud- In this study, a porous medium was represented by arrang-
sen-type expressions, Eqs. (42) and (48), or the Wakao-type ing solid spherical particles randomly. Using that, gas flow

13
162   Page 16 of 18 Microfluid Nanofluid (2016) 20:162

through a porous medium with micro-/nanoscale pores


 
1 1 + c1K p 2wall ∇p
was reproduced by performing DSMC simulations. Using U = −τ wall C̄ + p ,
3 1 + c2K p 32µ p
the results obtained in the DSMC simulations, the two
2 128 1 + c1K p ∇p
 
expressions to estimate gas flow velocity through a porous Us = −ετ wall 1 + Kn ,
medium were proposed. The results are summarized as 32 3π 1 + c2K p µ
follows. 
128 1 + c1K p

K = K0 1 + Kn ,
3π 1 + c2K p
1. From the results of the DSMC simulations, it was
found that the gas flow velocity through a porous 128 1 + c1K p
b= · Kn · p,
medium increases in proportion to increasing pressure 3π 1 + c2K p
gradient at least for Re < 0.1. This result shows that  1 
8 2 wall

K
Darcy’s law holds even in the case of a porous medium c1 = 2.00 ,
π 2µC̄
with micro-/nanoscale pores.  1 
2. The gas flow velocity estimated by using conventional

K 8 2 wall
c2 = 2.47 .
models for permeability, i.e., the Blake–Kozeny, the π 2µC̄
Carman–Kozeny, and the Rumpf–Gupte models, more
greatly deviated from that obtained in the DSMC simu- The Wakao-type expressions for porous media:
lations as the particle size of a porous medium became
smaller from micrometer to nanometer scale. 2wall

128
 
32
 
3. We regarded the porous medium as a bundle of tortu- U = −τ φ Kn + (1 − φ) 1 + a + Kn ∇p,
32µ 3π 3
ous capillary tubes with a diameter of wall, where wall 2

128
 
32
 
∇p
is the mean free path for molecule–wall collisions. At Us = −ετ wall φ Kn + (1 − φ) 1 + a + Kn ,
32 3π 3 µ
first, considering only Knudsen diffusion, the expres- 
128
 
32
 
sion to estimate gas flow velocity U through a porous K = K0 φ Kn + (1 − φ) 1 + a + Kn ,
3π 3
medium was constructed as follows: 
128

32

b= φ + (1 − φ) a + − 1 Kn · p,
3π 3
1 ∇p
U = − τ wall C̄ . Kn
3 p φ= .
Kn + 1
In the case where τ is set at τ = 0.9ε, the flow veloc-
ity estimated by using this expression agreed well with Here, the Knudsen number Kn is given by
that obtained in the DSMC simulations for particle Kn = mol /wall and K0 is the permeability in the con-
diameter dp < 200 nm, i.e., for Kn > 0.6. However, tinuum limit and given by K0 = ετ (2wall /32).
for dp > 200 nm, i.e., for Kn < 0.6, the larger the par- 5. We derived the expression to estimate the mean free
ticle diameter became, the more greatly the velocity path wall for molecule–wall collisions as follows:
estimated by using this expression deviated from that

obtained in the DSMC simulations. wall = .
Sv
4. Our expression shown in the previous item is not appli-
cable to the case of a Knudsen number much smaller The mean free paths wall for molecule–wall colli-
than unity. This is because viscous flow is not consid- sions estimated by using this expression agreed well
ered in that expression. Referring to Knudsen’s expres- with those obtained in the DSMC simulations, and
sion and Wakao’s expression proposed to estimate hence, the validity of this expression was confirmed.
gas flux through a straight cylindrical capillary, our This expression suggests that the mean free path
expression was modified by adding the contribution wall for molecule–wall collisions depends only on
of viscous flow with a velocity slip. The obtained sets the channel configuration of a porous medium and is
of expressions for the flow velocity U, the superficial equivalent to the hydraulic diameter for the porous
velocity Us, the permeability K, and the Klinkenberg medium. Therefore, our choice of wall as the charac-
coefficient b are as follows: The Knudsen-type expres- teristic length for both Knudsen and Reynolds num-
sions for porous media: bers is valid, and furthermore, using wall as the tube

13
Microfluid Nanofluid (2016) 20:162 Page 17 of 18  162

diameter in our tube bundle model for porous media Doi T (2010) Numerical analysis of the Poiseuille flow and the
is reasonable. thermal transpiration of a rarefied gas through a pipe with a
rectangular cross section based on the linearized Boltzmann
6. The flow velocities estimated by using the Knudsen- equation for a hard sphere molecular gas. J Vac Sci Technol A
type expression and by using the Wakao-type expres- 28(4):603–612
sion together with wall estimated in such a way as Dreyer JAH, Riefler N, Pesch GR, Karamehmedović M, Fritsching U,
mentioned in the previous item agreed well with those Teoh WY, Mädler L (2014) Simulation of gas diffusion in highly
porous nanostructures by direct simulation Monte Carlo. Chem
obtained in the DSMC simulations for the range of Eng Sci 105:69–76
the porosity of 0.3 < ε < 0.5 and the pore size from Dullien FAL (1979) Porous media: fluid transport and pore structure,
nanometer to micrometer scale. Thus, if we know the chap 4. Academic Press, New York
porosity ε and the wetted surface area Sv per unit vol- Evans JW, Abbasi MH, Sarin A (1980) A Monte Carlo simula-
tion of the diffusion of gases in porous solids. J Chem Phys
ume of a porous medium, we can make a good estimate 72(5):2967–2973
of the gas flow velocity and the volume flux by using Fan J, Shen C (1999) Statistical simulation of low-speed unidirec-
these expressions proposed here together with the esti- tional flows in transitional regime. In: Brun R, Campargue R,
mated mean free path wall for molecule–wall collisions Gatignol R, Lengrand J (eds) Rarefied gas dynamics, vol 2,
Cepadues editions, pp 245–252
for porous media with pores ranging from nanometer Fan J, Shen C (2001) Statistical simulation of low-speed rarefied gas
to micrometer and porosity of 0.3 < ε < 0.5 . flows. J Comput Phys 167:393–412
Ghanbarian B, Hunt AG, Ewing RP, Sahimi M (2013) Tortuos-
Acknowledgments The DSMC simulations were carried out using ity in porous media: a critical review. Soil Sci Soc Am J
the scalar parallel supercomputer SGI UV2000 of the Advanced Fluid 77(5):1461–1477
Information Research Center, Institute of Fluid Science, Tohoku Uni- Guo Z, Zheng C (2008) Analysis of lattice Boltzmann equation for
versity. One of the authors (Y.K.) is supported by JSPS Research Fel- microscale gas flows: relaxation times, boundary conditions and
lowship for Young Scientists. the Knudsen layer. Int J Comput Fluid D 22(7):465–473
Ikegawa M, Kobayashi J (1988) Development of rarefied gas flow
Open Access  This article is distributed under the terms of the Crea- simulator using the direct simulation Monte Carlo method (1st
tive Commons Attribution 4.0 International License (http://crea- report, 2-D flow analysis with the pressure conditions given at
tivecommons.org/licenses/by/4.0/), which permits unrestricted use, the upstream and downstream boundaries). Trans Jpn Soc Mech
distribution, and reproduction in any medium, provided you give Eng Ser B 54(507):3057–3060
appropriate credit to the original author(s) and the source, provide a Jones SC (1972) Rapid accurate unsteady-state Klinkenberg per-
link to the Creative Commons license, and indicate if changes were meameter. Soc Pet Eng J 12:383–397
made. Kalarakis AN, Michalis VK, Skouras ED, Burganos VN (2012) Meso-
scopic simulation of rarefied flow in narrow channels and porous
media. Transp Porous Med 94(1):385–398
Karniadakis GE, Beskok A, Aluru N (2005) Microflows and nano-
References flows: fundamentals and simulation. Springer, New York
Klinkenberg L (1941) The permeability of porous media to liquids
Asinari P, Quaglia MC, Spakovsky MR, Kasula BV (2007) Direct and gases. API Dril Prod Prac, pp 200–213
numerical calculation of the kinematic tortuosity of reactive mix- Knudsen M (1909) Die gesetze der molekularströmung und der
ture flow in the anode layer of solid oxide fuel cells by the lattice inneren reibungsströmung der gase durch röhren. Ann Phys
Boltzmann method. J Power Sources 170:359–375 333(1):75–130
Beskok A, Karniadakis GE (1999) A model for flows in channels, Kozeny J (1927) Über kapillare leitung des wassers im boden (auf-
pipes, and ducts at micro and nano scales. Microscale Therm stieg, versickerung und anwendung auf die bewässerung). Sitz
Eng 3(1):43–77 Ber Akad Wiss Wien, Math Nat (Abt IIa) 136(a):271–306
Bird GA (1994) Molecular gas dynamics and the direct simulation of Mason EA, Malinauskas AP, Evans RB III (1967) Flow and diffusion
gas flows. Clarendon Press, Oxford of gases in porous media. J Chem Phys 46(8):3199–3216
Bird RB, Stewart WE, Lightfoot EN (2007) Transport phenomena. Michalis VK, Kalarakis AN, Skouras ED, Burganos VN (2010) Rare-
Wiley, New York faction effects on gas viscosity in the Knudsen transition regime.
Blake FC (1922) The resistance of packing to fluid flow. Trans Am Microfluid Nanofluid 9(4):847–853
Inst Chem Eng 14:415–421 Nanbu K (1980) Direct simulation scheme derived from the Boltz-
Bravo MC (2007) Effect of transition from slip to free molecular flow mann equation. I. Monocomponent gases. J Phys Soc Jpn
on gas transport in porous media. J Appl Phys 102(074):905 49(5):2042–2049
Carman PC (1937) Fluid flow through granular beds. Trans Inst Chem Nanbu K (1992) Stochastic solution method of the Boltzmann equa-
Eng 15:150–166 tion-I. Mem Inst Fluid Sci 3:47–93
Carman PC (1938) The determination of the specific surface of pow- Nanbu K (2000) Probability theory of electron–molecule, ion–mol-
ders. J Soc Chem Ind (Trans) 57:225–234 ecule, molecule–molecule, and Coulomb collisions for particle
Carman PC (1956) Flow of gases through porous media. Butter- modeling of materials processing plasmas and cases. IEEE T
worths, London Plasma Sci 28(3):971–990
Christou C, Dadzie SK (2016) Direct-simulation Monte Carlo investi- Nie X, Doolen GD, Chen S (2002) Lattice–Boltzmann simulations of
gation of a Berea porous structure. SPE J 21(3):69–76 fluid flows in MEMS. J Stat Phys 107(1):279–289
Cunningham RE, Williams RJJ (1980) Diffusion in gases and porous Ohwada T, Sone Y, Aoki K (1989) Numerical analysis of the Poi-
media. Plenum Press, New York seuille and thermal transpiration flows between two parallel
Darcy HPG (1856) Les Fontaines Publiques de la Ville de Dijon. Vic- plates on the basis of the Boltzmann equation for hard-sphere
tor Dalmont, Paris molecules. Phys Fluids A 1(12):2042–2049

13
162   Page 18 of 18 Microfluid Nanofluid (2016) 20:162

Oshima T, Yonemura S, Tokumasu T (2012) A numerical study for Weissberg HL (1963) Effective diffusion coefficient in porous media.
transport phenomena of nanoscale gas flow in porous media. AIP J Appl Phys 34(9):2636–2639
Conf Proc 1501:809–815 Weissberg HL, Prager S (1970) Viscous flow through porous media.
Rumpf H, Gupte AR (1971) Einflüsse der porosität und korngrößen- III. Upper bounds on the permeability for a simple random
verteilung im widerstandsgesetz der porenströmung. Chem Ing geometry. Phys Fluids 13(12):2958–2965
Tech 43(6):367–375 Yonemura S (2008) Particle modeling of high Knudsen number
Saito A, Okawa S, Maeda H, Suzuki T (1995) Simulation of rarefied flows and plasmas. In: Maruyama S, Ohara T (eds) Nano-mega
gas flow through porous media using direct simulation Monte scale flow dynamics in highly coupled systems, chap 3, vol 10.
Carlo method. Trans Jpn Soc Mech Eng Ser B 61(582):606–613 Tohoku University Press, Sendai, pp 57–99
Tomadakis MM, Sotirchos SV (1991) Effective Knudsen diffu- Zalc JM, Reyes SC, Iglesia E (2004) The effects of diffusion mecha-
sivities in structures of randomly overlapping fibers. AIChE J nism and void structure on transport rates and tortuosity factors
37(1):74–86 in complex porous structures. Chem Eng Sci 59(14):2947–2960
Tomarikawa K, Yonemura S, Tokumasu T, Koido T (2011) Numerical Zhang YH, Qin RS, Sun YH, Barber RW, Emerson DR (2005) Gas
analysis of gas flow in porous media with surface reaction. AIP flow in microchannels—a lattice Boltzmann method approach. J
Conf Proc 1333:796–801 Stat Phys 121(1):257–267
Wakao N, Otani S, Smith JM (1965) Significance of pressure gradi-
ents in porous materials: part I. diffusion and flow in fine capil-
laries. AIChE J 11(3):435–439

13

You might also like