Sei sulla pagina 1di 21

Transp Porous Med (2015) 109:433453

DOI 10.1007/s11242-015-0527-4

Scaling Invariant Effects on the Permeability of Fractal


Porous Media
Y. Jin1,2 Y. B. Zhu1,2 X. Li1,2 J. L. Zheng1,2
J. B. Dong1,2

Received: 25 November 2014 / Accepted: 3 June 2015 / Published online: 17 June 2015
Springer Science+Business Media Dordrecht 2015

Abstract Porous media are interconnected systems, in which the distribution of pore sizes
might follow scaling invariant property and will affect the fluid flow through it significantly.
Thus, except for a detailed understanding of the fundamental mechanism at pore scale,
hard is it to determine the appropriate relationship between the permeability and the basic
properties. In this study, in terms of the size distribution and spatial arrangement of the
pores, we analytically derived a permeability model using seriesparallel flow resistance
mode firstly. And then, together with the scaling invariant characteristics of the porosity,
specific area and hydraulic tortuosity, the analytical permeability model is reformulated into
a fractal permeabilitypore structure relationship. The results indicate that: (1) the square of
the porosity () is proportional to the permeability in a fractal porous media, not the cubic
law described in KozenyCarman (KC) equation; (2) the hydraulic tortuosity is a power
law model of the minimum particle size with the exponent (Df d), where Df and d are
the pore size fractal dimension and Euclidean space dimension, respectively, while  is a
parameter characterizing the spatial arrangement of pores; (3) the KC numerical prefactor
is not a constant in fractal porous media. Its value, however, increases linearly with the size
ratio of the minimum to the maximum pores but decreases exponentially with Df . More
importantly, it is found to be a parameter characterizing the difference of fluid flow in porous
media from that in a straight tube described by the Poiseuille law. The performance of the
new fractal permeabilitypore model is verified by lattice Boltzmann simulations, and the
numerical prefactor universality is examined as well.

This work was supported by the National Natural Science Foundation of China (Grant Nos. 41102093,
41472128), and CBM Union Foundation of Shanxi Province of China (Grant No. 2012012002).

Y. Jin
jinyi2005@hpu.edu.cn

School of Resource and Environment, Henan Polytechnic University, Jiaozuo 454003, China

Collaborative Innovation Center of Coalbed Methane and Shale Gas for Central Plains Economic
Region, Henan Province, Jiaozuo 454003, China

123

434

Y. Jin et al.

Keywords Fractal porous media Tortuosityporosity model Lattice Boltzmann method


KozenyCarman constant Permeabilitypore model

1 Introduction
Investigation of fluid flow in porous media is of paramount importance in different areas.
As one of the basic transport properties, permeability describes how easily the flow passes
through the porous media (Jin et al. 2013; Matyka et al. 2008; Costa 2006), and is a dependent
quantity on some fundamental and well-defined parameters determined solely by the geometry of porous media, such as the porosity, the specific surface area, the hydraulic tortuosity
and the shape factor. In the past, a lot of semiempirical relationships between the permeability
and the structure parameters were proposed experimentally or theoretically (Collins 1961;
Duda et al. 2011; Bear 1972; Dullien 1991), among them the modified KC equation (Carman
1937, 1939; Kozeny 1927) should be the most well-known one for solid particles packing
bed.
Natural porous media microstructure might be disordered and complicated, with
pores/particles distributed scaling invariantly (Adler and Thovert 1998; Smidt and Monro
1998; Thovert et al. 1990; Young and Crawford 1991; Krohn and Thompson 1986). Substantial difference has been observed in the measured permeability even when porous media share
the same statistical quantity of fundamental properties (such as porosity), and the so-called
KC constant in the modified KC equation is actually an empirical parameter, which will alter
with pore structures (Costa 2006; Ahmadi et al. 2011; Cai et al. 2010; Panda and Lake 1994;
Rahli et al. 1997; Xu and Yu 2008). Thus, one is hard to determine the appropriate relationship between the permeability and the basic properties because of the large number of related
parameters, except for a detailed understanding of size distribution and spatial arrangement
of pores and particles in porous media (Costa 2006).
To improve the estimation accuracy, various approaches were employed to investigate the
permeabilitypore structure (short by permeabilitypore later) relationship, approximately
catalogued into three types: experiment-based analysis, analytical derivations and numerical
simulations. The experimental results are usually influenced by testing techniques, scales,
experimental environments, etc. Meanwhile, owing to the potential continuous assumptions
underlying in experiments, the microstructural effects on fluid flows may be thus ignored
(Costa 2006; Yu et al. 2003). That will lead to some unexplained items, always called semiempirical parameters.
Following the analytical approaches, a mathematical framework can be established with
clear physical meanings for a problem at hand by some simplifications, guiding us to adapt the
inherent coefficients to suit experiments (Ahmadi et al. 2011). Pitchumani and Ramakrishnan
(1999) proposed a permeability model for fractal porous media by assuming them as fractal
tube bundles other than the idealized arrangements of tube bundles with same length (Kozeny
1927; Happel 1959; Sangani and Acrivos 1982). However, the model presented in Pitchumani
and Ramakrishnan (1999) presents unreasonable results, interested readers may consult the
comments by Yu (2001) for detail. And then, Yu et al. (2003, 2005) analytically derived a
permeabilityporosity relationship, assuming that the size distribution of pores follows fractal
statistics as well as that the relationship between the diameter and the length of capillary tubes
admits a fractal scaling law (Wheatcraft and Tyler 1988). Later, these authors developed a
new form of permeability and KC constant, and demonstrated how KC constant alters with
the range of pore size (Xu and Yu 2008). Other than the accumulation method based on the

123

Scaling Invariant Effects on the Permeability of Fractal Porous

435

nonintersect assumption of the capillary tubes, another idea is to reformulate the classical
KC equation directly, by inserting the scaling invariant relationships between fundamental
parameters and the microstructure of porous media with special settings (Sahimi 1993; Rawls
et al. 1993; Guarracino 2007; Nasta et al. 2013). Accordingly, Costa (2006) proposed a twoparameter permeabilityporosity equation, and then Henderson et al. (2010) analytically
derived a three-parameter permeability model. Jin et al. (2013) found that the permeability
is a function of the size of the largest pore, the porosity, the pore size fractal dimension (Df )
and a semiempirical parameter. Based on fractal geometry and Poiseuilles law, Wang et al.
(2014) analytically derived a semiempirical fractal permeability model, indicating that the
porosity to the power of (4 Df )/(2 Df ) is proportional to the permeability of a twodimensional porous medium. Obviously, direct investigations are more reliable than those
obtained on the assumption of the capillary tubes apart from each other (Jin et al. 2013; Costa
2006; Wang et al. 2014).
Except for the solutions mentioned above, more and more efforts are now devoted to
direct numerical simulations because of their advantages in understanding the basic physics
of a certain problem (Croce et al. 2007). In numerical experiments, one can easily select
or neglect any relevant effects, and reduce the uncertainty from coupling effects as far as
possible. Moreover, the computational fluid dynamics (CFD) models are not subjected to
any experimental techniques, scales or environments. Among these CFD models, the newly
developed lattice Boltzmann method (LBM) has drawn broad and special attention (Chen
and Doolen 1998; Kandhai et al. 1999; Koponen et al. 1997; Ladd 1994; Succi 2001), and
has been proven to be a powerful tool to explore the controlling mechanism of complex flow
problems (Nithiarasu and Ravindran 1998; Degruyter et al. 2010; Vita et al. 2012). However,
different relationships can be derived from the same numerical results if the background
guiding models are different. Thus, to understand the complex fluid flow process in porous
media mechanistically, we need to establish a mathematical framework in advance, and then
investigate the basic physics by numerical simulations to reduce the uncertainty in it.
Based on the analyticalnumerical coupling solution, here we first review some classical permeability equations. Under the hypothesis of fractal pore-space geometry and using
seriesparallel flow resistance mode, we then analytically derive a permeability estimation
model with clear physical background. With the help of LBM, we investigate the basic
physics of the geometry effects on the fluid flow. Finally, we establish a permeabilitypore
relationship for fractal porous media, and demonstrate its validity via comparisons between
the existing correlations and the numerical results.

2 Theories and Methodologies


2.1 The Fluid Flow Resistance Model of Porous Media
Since the pioneer work by Kozeny (1927), great efforts have been devoted to establishing
the estimation models of the permeability K of porous media. Carman (1937, 1939, 1956)
modified the original Kozenys equation as follows:
K =

3
,
k(1 )2 S 2

(1)

where is the porosity, S is the specific surface and k = Cf 2 is the KC constant, in which
is the hydraulic tortuosity, and Cf is a constant shape factor dependent on the capillary pore

123

436

Y. Jin et al.

shape. Assuming the porous media is deposited by monosized solid particles, substituting
for S in Eq. (1) leads to
K =

3
2 ,
C0 2 (1 )2

(2)

where is the particle size, e.g., the side length of a cubic block or the diameter of spherical
particles. C0 has been considered to be a constant relative to the shape factor of the solid
particles which are monosized filled in the porous media.
As pointed out by Bear and Verruijt (1987), the flow rate of a porous media can be
expressed as a function of its flow resistance:
Q=

P
,
R

(3)

where Q is the total flow rate, R is the flow resistance and P denotes the pressure difference
between the inlet and outlet boundary. Similar to Ohms law, R follows the parallel and series
calculation models.
Combining Darcys law with the definition of flow resistance [see Eq. (3)], the relationship
R = L/(AK ) is obtained, where is the fluid dynamic viscosity, A and L is the crosssectional area and sample length of the porous medium, respectively. By borrowing the
concept of electrical resistivity, the fluid flow resistivity Rf can be defined as the quantity
of flow resistance of certain porous media with unit sample length and cross-sectional area,
which yields Rf = /K . Together with Eq. (2), the flow resistivity of porous media filled
with monosized solid particles reads:
Rf =

C0 (1 )2 2
,
3
2

(4)

which implies that Rf is the function of and if porosity and the shape of solid particles
are same.

2.2 Lattice Boltzmann Methods


LBM is actually a mesoscopic description of microscopic physics, and has been used widely
to study the microstructural effect on the fluid flow at pore scale. For simplicity and convenience, the model description and fluids simulations are presented using the classical lattice
BhatnagarGrossKrook (BGK) model over a D2Q9 lattice structure.
In LBM framework, based on the principle of effective conversion between physical and
lattice systems, the real pore space Dp is discretized in terms of a regular lattice with spacing
x , time t in terms of a time step t and velocity space in terms of a small set of velocities ci
to ensure that ci t is a vector connecting two adjacent lattice sites (Succi 2001; Dnweg et al.
2007; Sukop and Thorne 2007). For the employed D2Q9 lattice structure (Qian et al. 1992),
a simple-cubic lattice has a set of probability functions f i , representing the mass density of
fluid particles going through one of the 9 discrete velocities ci . Thus, the local mass density
and velocity u at each lattice position x and time t are given by
(x, t) =

8

i=0

8
f i (x, t),

u(x, t) =

f i (x, t)
.
(x, t)

i=0

(5)

By ChapmanEnskog expansion of the Boltzmann equation, the time evolution of fluid


flow is described by the lattice Boltzmann equation (LBE) (Sukop and Thorne 2007), as:

123

Scaling Invariant Effects on the Permeability of Fractal Porous

437

f i (x, t) + i = f i (x + ci t , t + t ),

(6)

where i is the collision operator modifying


8 the populations
8 at x according to the mass and
momentum conservative requirements of i=0
i = i=0
i ci .
According to BGK model, the collision operator takes the single-relaxation-time approximation (Chen and Doolen 1998; Succi 2001; Qian et al. 1992),
i =

t 
lbm


eq
f i (x, t) f i (x, t) ,

(7)

eq

where lbm is a dimensionless relaxation time, and f i (x, t) is a quasi-equilibrium distribution function. And then, its discrete velocities ci are defined as

i =0
(0, 0),
(cos , sin ),
i = 1 : 4, = i1
(8)
ci = c
2 ,

2(cos , sin ) i = 5 : 8, = 2i9


4
eq

to recover the NavierStokes (NS) equation for the fluid flow, f i is constructed by Eq. (9),
and the kinetic viscosity of the fluid () is given by Eq. (10), respectively.


u ci
9(u ci )2
3u 2
eq
,
(9)

f i (x, t) = i (x, t) 1 + 3 2 +
c
2c4
2c2
= (lbm 0.5)

2x
.
3 t

(10)

with c = x / t and lattice sound speed c2s = c2 /3 due to the constrains of conservation and
isotropy. Equation (10) imposes a constraint on the choice of lbm being greater than 0.5 for
a physically correct viscosity (Sukop and Thorne 2007), here we choose lbm = 1.
And the
weight assignments follow 0 = 4/9, 14 = 1/9, and 58 = 1/36, to satisfy i = 1
for symmetry reasons.
In the complex fractal porous media, there is almost no net fluid motion that exists (Succi
2001). The boundary condition at solidfluid interfaces can, therefore, physically approximate the no-slip boundary condition (Jin et al. 2013; Wang et al. 2014; Chen et al. 2013).
For simplicity and without loss of generality, the complete bounce-back scheme is adopted
in our flow simulations.

3 Characteristics of Fractal Porous Media and Their Flow Resistance


Porous media in nature always consist of numerous irregular pores of different sizes spanning
several orders of magnitude in length scales, such as soil, sandstones in oil reservoir, matrix
pores in coal, packed beds in chemical engineering, fabrics used in liquid composite molding
and wicks in heat pipes. These media possess pore microstructure, including both pore sizes
and pore interfaces, and exhibit fractal characteristics, thus called fractal porous media (Adler
and Thovert 1998; Smidt and Monro 1998; Thovert et al. 1990; Krohn and Thompson 1986).

3.1 Fractal Characteristics of Porous Media


In a fractal pore space, the cumulative distribution of pore size has been proven to follow
fractal scaling law. Obviously, a natural porous medium is composed of one or more representative units of linear size L at which the fractal behavior starts. These representative

123

438

Y. Jin et al.

units share the same physical properties, such as porosity, pore structure, pore size range and
transport property as the porous medium in a statistical mean, but contain only one pore/solid
of the largest size. And, the fractal dimension of the pore size distribution, Df , is a function
of the porosity and the ratio of the lower limit to upper limit of self-similar regions (Yu and
Li 2001):
r dDf
,
(11)
( r ) =
L
where r is the scale, d is the Euclidean space dimension, the fractal dimension Df is in the
range of 0 < Df < 2 and 0 < Df < 3 in two and three dimensions in natural porous media
(Yu and Li 2001), respectively. Recently, some authors indicated that Df could possess a
much wide range, for more details one can refer to Ghanbarian-Alavijeh and Hunt (2012).
In our previous study (Wang et al. 2014), two definitions have been proposed to describe the
microscopic physical properties following the fractal theory (Mandelbrot 1983), denominated
as Frequency of pore growth and Lacunarity of pore size, denoted by F and P , respectively.
F is the ratio of the pore number of size i to that of size i1 , while P is the ratio of
i1 /i , where i and i1 are two successive pore sizes in a fractal porous media and
i1 > i . Consequently, the fractal dimension Df can be calculated by log(F )/ log(P )
because it is another form of the number-size relation.
Meanwhile, in a fractal porous medium (for example the Sierpinski carpet or Menger
sponge fractals), if the smallest pore size is r , the smallest particle size will be r in a pore
fractal (Ghanbarian et al. 2013a), where is a constant and 1 (in standard Sierpinski
carpet, = 1). Thus, the number-size distributions of the pores and particles are power laws
with identical exponent Df in a fractal porous media (Perrier and Bird 2002).
Consequently, after n times iterations, the accumulative poreparticle surface area in a
representative unit of a fractal porous medium is then expressed by:
As =

n


(i1)

s id1 N (i ) , where N (i ) = Fi1 , i = max P

(12)

i=1

where i and N (i ) represent the size and number of the minimum solid particles after i-time
iteration, respectively. max is the largest size of solid particle in a representative unit, s is
the shape factor, for square solid particles s = 2d (Mortensen et al. 2005).
Obviously, the cumulative poreparticle surface As is the sum of a geometric series.
Following the definitions of P and F , together with their description of Df , we obtain the
relation max = min Pn1 . Instantaneously, Eq. (12) is rearranged into:

d1Df


1 P max
d1
As () = s max
1 PDf +1d

d1Df

d1
s max
,
(13)
max
where is a function of P , a parameter relevant to pore structure (Wang et al. 2014).
Actually, by substituting the relationships of = r and P max / = L into Eq. (13), the
poreparticle surface area yields:

d1Df
1 Lr
.
(14)
As () = s d1
max
1 PDf +1d

123

Scaling Invariant Effects on the Permeability of Fractal Porous

439

By comparison, Eq. (14) is consistent with what proposed is in Yu et al. (2009) because
these authors assumed that the size of the maximum pore max is equal to the linear size of
the representative region/unit of the porous media, and the size of the minimum pores min
is equal to the measuring scale r .
Consequently, the specific area of fractal porous media is expressed by:
As ()
Ld

(d1Df )

1 max d1
= s
,
L
L
max

S() =

(15)

Equation (15) indicates that the poreparticle surface area admits a fractal scaling law
with . Meanwhile, the number-size distributions of the poreparticle surface area and the
solid particles are similar power laws with identical exponent because of the same scaling
behaviors. Instantaneously, the fractal dimensions of poreparticle surface area, the pores
and solid particles are the same in a fractal porous medium, being Df .
According to the independent measuring results, some authors considered that the fractal behaviors of the pores, solid particles and poreparticle surface area in a fractal porous
media are different (Perrier and Bird 2002; Perrier et al. 1999; Dathe and Thullner 2005).
It must be noted that the fractal dimension is just a number by which to quantify the scaling invariant degree of a pattern in geometrical or statistical scenes, meaning that just by
the fractal dimension, the geometries of fractal porous media could not be uniquely determined. Thus, to accurately describe the self-similar property of a porous media, except for
the fractal dimension, the spatial or statistical pattern which scales must be accompanied,
such as the generators of Menger sponge and the PSF model (Perrier et al. 1999). Otherwise, by measuring approaches ignoring the spatial arrangement of pores and particles
completely, such as the box-counting method, one will obtain different fractal dimensions
for the physical properties even if they possess the same scaling behaviors actually, as the
results in Perrier and Bird (2002), Perrier et al. (1999), Dathe and Thullner (2005) and Zhou
et al. (2010).
When the porous medium is filled with monosized solid particles, there will be no fractal
behavior of pore/particle sizes. In such a condition, F 0+ , thus Df will tend to be
theoretically because of Df = log(F )/ log(P ). Obviously, this is consistent with what
pointed out is in Ghanbarian-Alavijeh and Hunt (2012), in which the authors indicated that
Df = represents a uniformly grain/pore size distributed porous medium. Denoting by
S0 the specific area of such a medium, from Eq. (15) we get Eq. (16) instantaneously because
of = max .
S0 = s

d1
max
.
Ld

(16)

For a two-dimensional nonfractal porous medium, the boundary length of a solid particle
d1 , thus the specific area takes the form of Eq. (16).
is equal to 4max , also s max
As aforementioned, at a measuring scale of pore size r , the minimum measurable size of
solid particles takes the value r . Thus, together with Eq. (11) and with respect to the pore
size, Eq. (15) is then rearranged into:
S( r ) = S() = s 1Df d1Df

= Cs .
r
r

(17)

where 1 = max /L, being a constant in a fractal porous media. For example, 1 = 1/3 in
standard Sierpinski carpet.

123

440

Y. Jin et al.

Fig. 1 Basic construction of the VmSqLnRl-type fractal porous media. The solid and void phases are denoted
by black and white, respectively. In a space of Euclidean dimension d, the initiator (a) of linear size l defines
the representative region/unit of a porous media, divided into m 2 equal parts. At the first iteration step, the
generator (b) divides the m 2 parts into two sets of q 2 (solid) and m 2 q 2 (void) subregions. At the next step,
each of the void subregions is replaced by a reduced replicate of the generator

3.2 Construction of Porous Media with Arbitrary Fractal Dimension


For simplicity and without loss of generality, our attention is focused on the permeabilitypore
relationship of fractal porous media in two-dimensional context. In our previous studies (Jin
et al. 2013; Wang et al. 2014), an algorithm was proposed to construct a self-similar fractal
object with arbitrary fractal behavior and without blind pores, denominated as VmSqLnRltype fractals. The modeling algorithm is briefly as following: (1) define the representative
(R) region of linear size l as an initiator in a space of Euclidean dimension d (Fig. 1a). This
region is then divided into m m small subregions of linear size l/m and set to be void phase
(V ); (2) solidify (S) the very central region with size q l/m (Fig. 1b); (3) repeat step (2)
in the reset void subregions for n times until the predefined porosity is achieved. The basic
construction of a V5S3L2Rl-type porous media is demonstrated in Fig. 1, where Fig. 1a is an
initiator or the representative region of a fractal porous media, Fig. 1b is the generator and
Fig. 1c is a V5S3L2Rl-type fractal porous medium with Df = log 14/ log 5.
It is noticeable that, with successive iterations, pores and solid particles keep their sizes
reduced, while their number increased (Adler and Thovert 1993; Tarafdar et al. 2001). In
a representative region of linear size l, max = l and min = l are satisfied at the very
beginning. After one subdivision, min (also r ) turns into l/m, min = max = lq/m; after n
times of iterations, min = lm n , min = lq/m n . Hence, for a VmSqLnRl-type fractal porous
medium, max and min are l and lm n , respectively, while max = lq/m and min = qmin .
In these media, and 1 are expressed by
= q, 1 =

q
,
m

(18)

and the Frequency of pore growth (F ) and the Lacunarity of pore size (P ) are expressed
by
P = m d ,

F = m d q d .

(19)

Thus, the fractal dimension of pore area and porosity are given by Eqs. (20) and (21),
respectively.

123

Scaling Invariant Effects on the Permeability of Fractal Porous

441

Fig. 2 Relationship between real specific area and that by Eq. (22) of different VmSqLnRl-type porous media,
where the solid line is y = x for reference
Table 1 Pore structure
parameter of some
VmSqLn-type porous media by
the best-fitted linear model
between s (q/m)d1 (/min )
and the real specific area

Porous media type

d/Df

V3S1Ln

1.0566

1.6000

V4S2Ln

1.1158

1.5000

V5S3Ln

1.1611

1.4545

V7S5Ln

1.2246

1.4118

V9S3Ln

1.0275

1.1429

V11S3Ln

1.0164

1.1089

V11S5Ln

1.0507

1.1294

V11S7Ln

1.1214

1.1803

V25S3Ln

1.0023

1.0423

V45S7Ln

1.0032

1.0233

ln F
ln (m d q d )
,
=
ln P
ln m

n

n
dmax dmax
Df d
= m
=
.
dmax

Df =

(20)
(21)

Actually, following the PSF model (Perrier et al. 1999), F is ranged in (0, m d q d ]. When
F 0+ , Df theoretically (Ghanbarian-Alavijeh and Hunt 2012). According to
Eq. (17), the specific area of VmSqLnRl-type porous media is then expressed by Eq. (22),
which is validated from different VmSqLnRl-type porous media as shown in Fig. 2.
S ( min ) = s

q d1
.
m Df min

(22)

Some pore structure parameter are listed in Table 1. It is noted that will approach to
d/Df as Df increases.

123

442

Y. Jin et al.

3.3 Flow Resistance of Fractal Porous Media


According to Eq. (21), the porosity of a VmSqL1Rl-type porous medium is determined only
by the parameters m and q, so is the hydraulic tortuosity due to the spatial arrangement of
void and solid phase determined by the construction process in such type porous medium.
And in a VmSqLnRl-type fractal porous medium, the space is filled with VmSqL1R(lm n )
units and solid phase, as shown in Fig. 1. For clarity, the porosity and hydraulic tortuosity of
VmSqL1Rl-type porous media are denoted by 1 and 1 , respectively. Taking them together,
we can rewrite Eq. (4) into the form of Eq. (23) for certain fluid flow through VmSqL1Rl-type
porous media with determined m and q.
Rf = C 1

1
1
2,
2

(23)

where stands for proportional to, = 1 l for porous media of VmSqL1Rl-type, and
constant C 1 reads
C1 =

C0 (1 1 )2 21
13

(24)

Equation (23) implies that the spatial distribution of 2 satisfies the parallel and series
mode.
(d/i)
For convenience, denote by R f
the flow resistivity of the VmSqLiRl-type porous
medium. Thus, when a VmSqL1Rl-type porous medium turns into the VmSqL2Rl-type after
(d/1)
(d/2)
one iteration, its flow resistivity will change from R f
to R f . According to Eq. (23),
(d/1)

(d/2)

takes the value of C1 (qd/m)2 , while R f


can be calculated from Eq. (25) by the
Rf
parallelseries spatial integration in a two-dimensional space.



q
mq
(dm 1 )/1
(d/2)
Rf
+
.
(25)
= Rf
m
mq
Replacing the expression (m q)/m + q/(m q) by f (m, q) for short, we can express
the average flow resistivity of VmSqLnRl-type porous media by

n1 (d/1)
(d/n)
Rf
= m 2 f (m, q)
Rf
n1 1
2
= C 1 m f (m, q)
.
(26)
2max
Substituting max = q L/m into Eq. (26) yields
2
n
m f (m, q) 1
(d/n)
Rf
= C1 2
.
q f (m, q) L 2

(27)

where L is the linear size of the representative unit of a fractal porous medium, also the size
of the largest pore.
Although Eq. (27) is derived from the regular model, but it is applied to porous media with
solid particles distributed randomly according to the parallelseries model. To explain intuitively, we demonstrate a VmSqL2Rl-type model with the first-level solid particle distributed
randomly on porous media composed of VmSqL1Rlm1 -type models, see Fig. 3.
Denoting by R1 , R2 and R3 the flow resistivity of VmSqL1Rl-, VmSqL1Rl/m- and
VmSqL2Rl-type model, respectively, by parallelseries mode, we can get the relationship
R3 = R2 ((m q)/m + q/(m q)) for arbitrary a and b, same as Eq. (25) derived from

123

Scaling Invariant Effects on the Permeability of Fractal Porous

443

Fig. 3 Demonstration of a
VmSqL2Rl-type model with solid
particle distributed randomly

regular model. So, the flow resistivity model of Eq. (27) is not only applied to regular fractal
porous media, but also to the random ones on average. However, Ghanbarian-Alavijeh and
Hunt (2012) have pointed out that the parallelseries approach does not model connectivity
among pores, which are more descriptive rather than predictive. But we think that the randomness effect on the transport property is in the charge of the semiempirical parameter of C0 .
Thus, according to the relationship between the permeability (K ) and flow resistivity
(d/n)
Rf
and taking Eqs. (17), (21) and (27) all together, we obtain the permeability estimation
model of VmSqLnRl-type porous media, reads
K =

2 s2 1 2
C0 S()2 T 2

(28)



where T 2 = 21 f (m, q)n1 is introduced as a geometrical parameter accounting for the
hydraulic tortuosity of a fractal porous medium.

4 Results and Discussions


As a general function, Eq. (28) should be able to describe the permeabilitypore relationship
for porous media without fractal behavior. In such a condition, n = 1, 1 = , and S takes
the value of S0 . Then, Eq. (28) is reduced to
K =

13
C0 12
13

2
max
d
2

max
L

2
max

C0 12 (1 1 )2

(29)

which is equal to the modified KC function defined in Eq. (2) for porous media deposited by
monosized particles.
To make Eq. (28) practical, the parameter T must be described by the fundamental properties contained in Eq. (2). As aforementioned, the hydraulic tortuosity depends only on
the spatial arrangement of solid particles, not on their sizes. That is, there is no effect on
the hydraulic tortuosity for a determined spatial pattern when it is zoomed in or out. For

123

444

Y. Jin et al.

Fig. 4 Relationships between 1 and 1 , 1 and 2 . The circles represent the experimental data, and the solid
line its best power law-fitted result

example, in a VmSqLnRl-type porous medium, will alter with parameters m, q and n, but
not with the linear size of the representative unit L. Meanwhile, a certain spatial pattern
leads to a determined porosity; however, for a porous medium with a determined porosity,
its hydraulic tortuosity will alter with the spatial arrangement of solid phase. As pointed out
in Valds-Parada et al. (2011), the hydraulic tortuosity cannot be a function of porosity only.
To explore the tortuosityporosity relationship, we calculate 1 of different VmSqL1-type
porous media by the method in Jin et al. (2015) based on the velocity fields simulated by the
LBM. Meanwhile, inspired by the construction process of fractal porous media, we rewrite
T 2 as 22 f (m, q)n , with 1 / f (m, q)0.5 denoted by 2 , and plot the relationship between 1
and 1 , 2 , as well as that between and f (m, q)n/2 (see Figs. 4, 5).
In Fig. 4, one is hard to find a clear relation between 1 and 1 . But the best-fitted
result indicates that 2 and 1 follows the relation 2 = 1 1 after combining the spatial
arrangement. Meanwhile, the relationship between f (m, q)n/2 and follows f (m, q)n/2 =
2  , as shown in Fig. 5. 1 and 2 are the fitted coefficients of power law models and both
approximate to 1.0; and the fitted exponents are almost the same, denoted by  and satisfying
 = 0.71.
As we all know, the tortuosity was first proposed as a fundamental parameter to predict
the permeability (Carman 1937), as shown in Eq. (1). But, due to the ambiguous concept, the
tortuosity has several definitions, some are based on geometrical approaches, while others
prefer to the hydraulic ones, for details one can refer to the literatures published recently
(Ghanbarian et al. 2013a, b).
Vast investigations have shown that the tortuosity is related to the porosity, and one of
the most invoked tortuosityporosity models is given by:
= 1 pln ()

(30)

where p is a coefficient and will be semiempirically modified by the microstructure of


porous media, as the estimations have been reported in various numerical or experimental
results (Matyka et al. 2008; Comiti and Renaud 1989; Mauret and Renaud 1997).

123

Scaling Invariant Effects on the Permeability of Fractal Porous

445

Fig. 5 Relationship between and f (m, q)n/2 . The circles represent the raw data, and the solid line the best
power law-fitted result

Other investigations showed that the tortuosityporosity relationship admits a power law
model, which reads:
=

(31)

where is an empirical exponent. Mota et al. (2001) found = 0.4 for binary mixtures
of spherical particles when measuring the conductivity of porous media; Liu and Masliyah
(1996a) reported that = 0.5 for random packs of grains with porosity > 0.2. Recently,
Ghanbarian et al. (2013b) found = 0.378 in their geometrical tortuosity model for porous
media whose pore size distribution is narrow.
Actually, the hydraulic tortuosity cannot be a function of porosity only (also shown in
Fig. 4 and pointed out in Valds-Parada et al. 2011), and cannot be universal (Ghanbarian
et al. 2013a). More generally, Henderson et al. (2010) assumed a fractal scale law as follows:
= C D ,

(32)

where C and D are the fractal coefficient and fractal exponent of , respectively.
For the consolidated porous media, Liu and Masliyah (1996b) found C = 1.61 and
D = 1.15. By means of fractal geometry, together with the scalemeasurement relation
proposed by Wheatcraft and Tyler (1988) for a fractal capillary tube, Feng and Yu (2007)
derived a geometrical tortuosity estimation model, approximately follows:

Df
Df + DT 1

L
min

DT 1
(33)

where DT was defined as the tortuosity fractal dimension of the curve in Wheatcraft and Tyler
(1988), and min is the diameter of the minimum capillary tube. Actually, Eq. (33) admits a
fractal scaling law between tortuosity and porosity in terms of Eq. (11).

123

446

Fig. 6 Relationship between 1 / and T , with T calculated by T =

Y. Jin et al.


21 f (m, q)n1 and 1 , by

Eq. (21). The solid line is the best fitted power law model to the relationship represented by circles

Recently, based on the geometry of standard Sierpinski carpet named by V3S1-type


fractal here, Li and Yu (2011) proposed a tortuosityporosity model which yields =
(19/18)ln / ln (8/9) (Ghanbarian et al. 2013b). For a V 3S1-type porous medium, 1 = 8/9
according to Eq. (21), thus the tortuosity model of Li and Yu (2011) can be rewritten into
= (19/18)ln / ln 1 , indicating that the tortuosity is a function of , 1 and the prefractal
distribution of solid phase. Even though these models are either based on the highly idealized geometry or from special porous media, they do provide a mathematical framework to
establish the tortuosityporosity relationship. Of course, all these tortuosityporosity models
are validated in a certain range of porosity due to the percolation threshold, because at very
low porosity, the system cannot even percolate that makes defining a tortuosity meaningless
(Ghanbarian et al. 2013a). Obviously, the validity of porosity range is beyond our objective
of the present work.
Thus, in terms of the relationship between 2 and 1 , as well as that between f (m, q)n/2
and , it is reasonable to consider that T is a function of 1 /, following the power law
relationship expressed in Eq. (34), which is then validated by the best power law-fitted result
in Fig. 6.


1
T =
.
(34)

The difference between  and the exponents proposed before (Henderson et al. 2010; Mota
et al. 2001; Liu and Masliyah 1996a, b) should be ascribed to the fractal behavior and the
geometrical shape of solid particles.
In terms of the characteristics of fractal porous media, the relationship between 1 / and
the size distribution of particles satisfies:

max dDf
1
=
.
(35)

min

123

Scaling Invariant Effects on the Permeability of Fractal Porous

447

Taking into account Eqs. (34) and (35), the hydraulic tortuosity for a fractal porous medium
is then defined as:

max (dDf )
= C T = C
,
(36)
min
where C is introduced for the geometrical considerations and spatial arrangement of solid
phase, with C > 1. In porous media filled with monosized solid particles, the relationship
C = 1 is satisfied.
Equation (36) interrelates the tortuosity with determined parameters with clear physical
meanings, such as 1 , and tortuosity of the prefractal solid distribution 1 in a power law
model, which can be used directly for porous media with solid particles randomly distributed
because the derivation process using parallelseries flow resistance mode without special
assumption. And Eq. (36) is in accord with the real situation:
1. If a porous medium is filled with monosized solid particles, its hydraulic tortuosity will
be mainly affected by the spatial arrangement of solid phase;
2. For the fractal porous media with the same Df , the larger the ratio between max and min ,
the higher the value of the hydraulic tortuosity;
3. For any porous media with Df d, the hydraulic tortuosity 1, because Df d
means that the pore space will be ultimately filled with void phase, consequently resulting
in a straight-channel flow;
4. Because max /min 1 and (d Df ) > 0, that hydraulic tortuosity satisfies 1 is
always true, consistent with the tortuous nature of fluid flow through porous media.
For visual comparison, the streamlines of some VmSq-type porous media are demonstrated
in Fig. 7. In the LBM simulations, no-slip boundary conditions were imposed on the top and
bottom walls and periodic boundary conditions were assumed at the inlet (left wall) and outlet
(right wall). The flow was driven by an external force field whose magnitude was chosen
so that the Reynolds number Re < 1 to ensure the Darcys flow. For short, the hydraulic
tortuosity of fluid flow is denoted by TVmSqLnRl .
In Fig. 7, the streamlines indicate that: (1) TV3S1L3Rl > TV3S1L2Rl > TV3S1L1Rl , which
is due to that the ratio of max to min in V3S1L3Rl-type porous media is larger than that
in V3S1L2Rl-type porous media, even if they share the same Df ; (2) TV3S1L2Rl > TV9S3L2Rl
although the porosity of V3S1L2Rl- and V9S3L2Rl-type porous media is same. That is because
the hydraulic tortuosity will alert with Df , and a large fractal dimension will result in a small
hydraulic tortuosity, as aforementioned; (3) For a porous medium filled with monosized
solid particles, as that in Fig. 7a, it could be denominated by any kind of V3xSxL1Rl-type
porous media, such as V3S1L1Rl-, V6S2L1Rl- and V9S3LiRl-type as well. So, the hydraulic
tortuosity of VmSqL1Rl-type porous media is determined by q/m, because q/m determines
the spatial arrangement of solid phase.
Obviously, Eq. (36) is consistent with the numerical result. Substituting Eqs. (17) and (36)
into Eq. (28) gives:
K =

1 132 2 2
min
C0 12 (1 1 )2

(37)

According to Eq. (35), we can then rewrite Eq. (37) as


K =

1 132 2
C0 12 (1 1 )2

2
Df d

2
max
.

(38)

123

448

Y. Jin et al.

Fig. 7 (Color) The fluid flow streamlines of some VmSqLn-type porous media when the LBM simulations
reached a stable condition. ad are the flow streamlines of V3S1L1Rl-, V3S1L2Rl-, V3S1L3Rl- and V9S3L2Rltype porous media, respectively. The color represents the relative magnitude of the dimensionless velocity
which was normalized by the max velocity of a flow

The correlation between the analytical permeability (K as ) by Eq. (38) without parameter
C0 and that calculated from the LBM simulations is investigated for different VmSqLntype porous media of various linear sizes (K ns , calculated from the LBM simulations).
The relationships between K ns and C0 K as all yield highly linear correlation(not shown
here), but the coefficient C0 will alter with the pore structure characterized by m, q and
n. Substituting the fitted C0 into Eq. (38), we note that the values of permeability found
in LBM simulations are in excellent agreement with those estimated by Eq. (38), some
small deviation should be ascribed to the numerical precision and calculation errors (Fig. 8).
The corresponding parameters of the porous media for comparison in Fig. 8 are listed in
Table 2.
As aforementioned, parameter C0 has been considered to be a constant in the porous
media filled by monosized particles. But in fractal porous media, C0 is found to alert with the
porosity and pore structure, as shown in Fig. 9. The fitted result shows that C0 is a function
of the porosity and Df , approximately follows:
1

C0 32 dDf

123

(39)

Scaling Invariant Effects on the Permeability of Fractal Porous

449

Fig. 8 (Color) Relationship between K as and K ns . The symbols represent the relationship between K as and
K ns from different VmSqLn-type porous media, and the solid line serves as a reference, where y = x indicates
that K as has an identical value to K ns . l.u. is the dimensionless lattice unit
Table 2 The corresponding
parameters of the VmSqLn-type
porous media in Fig. 8 for
validation of Eq. (38)

Porous media type

Df

max /min

V3S1

1.893

335

15
14

V4S2

1.792

444

V5S1

1.975

553

13

V6S2

1.934

663

13

1.946

992

12

V9S3

In a fractal porous medium, that Df = d means the Euclidean space is fully occupied by
void phase. In such a condition, the flow rate Q in a circular capillary is usually represented
by the classical HP equation, reads:
Q=

4 P
,
128L

(40)

where is the capillary diameter. Meanwhile, according to the Darcys law, Q interrelates
the permeability K by:
Q=K

AP
.
L

(41)

where A the total cross-sectional area. Substituting Eq. (40) into Eq. (41) yields:
K =

4
.
128A

(42)

Because that the Euclidean space is fully occupied by void space, thus A = 2 /4 and
= L are satisfied. Consequently, Eq. (42) is rearranged into K = L 2 /32. For a VmSqLnRld1
type porous medium, when Df = d, the specific area S = s LL d , T = 1 (because 1 = 1 and

123

450

Y. Jin et al.

Fig. 9 Relationship between dDf and the fitted coefficients C0 . Symbol markers represent the fitted C0
from different porous models, and the solid line is the best-fitted power law model

f (m, q) = 1), 2 = 1 (see Table 1, and because P 0 when Df d), and 1 = = 1


are all satisfied. Together with Eq. (28), the relation C0 = 32 is obtained, which is obviously
consistent with the fitted result that is approximately expressed by Eq. (39), where the small
errors should be ascribed to the numerical precision.
In the meantime, taking Eqs. (11) and (39) both into account, we can get C0 min /L,
which is consistent with that pointed out by Xu and Yu (2008).
Together with Eqs. (28), (34) and (35), the KC constant k approximately yields:

min 1 2
k 32
L

1+2(Df d)

2(Df d) min
= 32P
.
(43)
L
Equation (43) indicates that the KC constant k is a function of the fractal dimension, range
and the Lacunarity of pore size.
1. If Df d, then k 32, meaning the fluid flow in porous media is approaching a
Poiseuille flow;
2. If the scaling characteristics is weak or there is no fractal behavior of the pore size
distribution. Thus, 1 and then k 32 min
L ;
L
and Df d are always satisfied,
3. If the size range of pores is fixed, because P < min
meaning that k will decrease with the fractal dimension Df ; in terms of Eq. (11), the
porosity increases with fractal dimension monotonically if the pore size range is fixed
(Jin et al. 2013), thus the smaller the porosity is, the larger is the value of k.
The variation characteristic indicates that the KC constant is not a constant, which characterizes the difference of fluid flow in a porous media between that in a straight tube. A large
deviation of a fluid flow from the Poiseuilles flow will result in a small k, vice versa.

123

Scaling Invariant Effects on the Permeability of Fractal Porous

451

Even though the variation trend of the KC constant proposed here is consistent with that
in Xu and Yu (2008), their estimation expressions are not the same. The main difference is
that we use intersected pore space other than the nonintersecting assumption of fluid paths.

5 Conclusions
In this study, we investigate the scaling invariant characteristics of the fractal porous media.
Following the analyticalnumerical coupling solution, the scaling effects of pores on the permeability, hydraulic tortuosity and the KC constant are fully analyzed, and some conclusions
are drawn as follows:
(1) In fractal porous media, the permeabilityporosity relationship follows power law correlation with exponent being 2, not the cubic law described in KC equation;
(2) The hydraulic tortuosity of the fractal porous medium is a function of the size range of
solid particles, fractal dimension of pore size and the prefractal solid distribution; the
tortuosityporosity relation admits a fractal scaling law with tortuosity fractal dimension
approximately being 0.71 in two-dimensional context;
(3) The KC constant is not a constant, which characterizes the derivation of the fluid flow
from the Poiseuille flow. If the pore size fractal dimension is deterministic, a large value
of min /L will yield a small KC constant; and KC constant will decrease with the fractal
dimension if the range of pore size is fixed;
(4) When the Euclidean space is occupied fully by the void phase, the fluid flow in porous
media will turn into the Poiseuille flow; while if there is no fractal behavior of pores
size, the medias permeability follows KC equation characterizing the fluid flow in the
pore space filled by monosized particles. The shift of these two kinds of fluid flow is
controlled by the fractal dimension and the range of pore size distribution.

References
Adler, P.M., Thovert, J.F.: Fractal porous media. Transp. Porous Med. 13, 4178 (1993)
Adler, P.M., Thovert, J.: Real porous media: local geometry and macroscopic properties. Appl. Mech. Rev.
51, 537585 (1998)
Ahmadi, M.M., Mohammadi, S., Hayati, A.N.: Analytical derivation of tortuosity and permeability of monosized spheres: a volume averaging approach. Phys. Rev. E 83, 026312 (2011)
Bear, J.: Dynamics of Fluids in Porous Media. Elsevier, New York (1972)
Bear, J., Verruijt, A.: Modeling Groundwater Flow and Pollution. Springer, New York (1987)
Cai, J.C., Yu, B.M., Zou, M.Q., Mei, M.F.: Fractal analysis of surface roughness of particles in porous media.
Chin. Phys. Lett. 27, 024705 (2010)
Carman, P.C.: Fluid flow through granular beds. Trans. Inst. Chem. Eng. 15, 150166 (1937)
Carman, P.C.: Permeability of saturated sands, soils and clays. J. Agric. Sci. 29, 262273 (1939)
Carman, P.C.: Flow of Gases Through Porous Media. Butterworths Scientific Publications, London (1956)
Chen, Q., Zhang, X.B., Zhang, J.F.: Improved treatments for general boundary conditions in the lattice Boltzmann method for convectiondiffusion and heat transfer processes. Phys. Rev. E 88, 033304 (2013)
Chen, S.Y., Doolen, G.D.: Lattice Boltzmann method for fluid flows. Annu. Rev. Fluid Mech. 30, 329364
(1998)
Collins, R.E.: Flow of Fluids Through Porous Materials. Reinhold Pub. Corp, New York (1961)
Comiti, J., Renaud, M.: A new model for determining mean structure parameters of fixed beds from pressure
drop measurements: application to beds packed with parallelepipedal particles. Chem. Eng. Sci. 44,
15391545 (1989)
Costa, A.: Permeabilityporosity relationship: a reexamination of the KozenyCarman equation based on a
fractal pore-space geometry assumption. Geophys. Res. Lett. 33, L2318 (2006)

123

452

Y. Jin et al.

Croce, G., DAgaro, P., Nonino, C.: Three-dimensional roughness effect on microchannel heat transfer and
pressure drop. Int. J. Heat Mass Transf. 50, 52495259 (2007)
Dathe, A., Thullner, M.: The relationship between fractal properties of solid matrix and pore space in porous
media. Geoderma 129, 279290 (2005)
Degruyter, W., Burgisser, A., Bachmann, O., Malaspinas, O.: Synchrotron X-ray microtomography and lattice
Boltzmann simulations of gas flow through volcanic pumices. Geosphere 6, 470481 (2010)
Duda, A., Koza, Z., Matyka, M.: Hydraulic tortuosity in arbitrary porous media flow. Phys. Rev. E 84, 036319
(2011)
Dullien, F.A.: Porous Media: Fluid Transport and Pore Structure. Academic Press, Salt Lake City (1991)
Dnweg, B., Schiller, U.D., Ladd, A.J.C.: Statistical mechanics of the fluctuating lattice Boltzmann equation.
Phys. Rev. E 76, 036704 (2007)
Feng, Y.F., Yu, B.M.: Fractal dimension for tortuous streamtubes in porous media. Fractals 15, 385390 (2007)
Ghanbarian, B., Hunt, A.G., Ewing, R.P., Sahimi, M.: Tortuosity in porous media: a critical review. Soil Sci.
Soc. Am. J. 77, 14611477 (2013)
Ghanbarian, B., Hunt, A.G., Sahimi, M., Ewing, R.P., Skinner, T.E.: Percolation theory generates a physically
based description of tortuosity in saturated and unsaturated porous media. Soil Sci. Soc. Am. J. 77,
19201929 (2013)
Ghanbarian-Alavijeh, B., Hunt, A.G.: Comments on More general capillary pressure and relative permeability
models from fractal geometry by Kewen Li. J. Contam. Hydrol. 140141, 2123 (2012)
Ghanbarian-Alavijeh, B., Hunt, A.: Unsaturated hydraulic conductivity in porous media: percolation theory.
Geoderma 187188, 77 (2012)
Guarracino, L.: Estimation of saturated hydraulic conductivity Ks from the van Genuchten shape parameter.
Water Resour. Res. 43, W11502 (2007)
Happel, J.: Viscous flow relative to arrays of cylinders. AiChE J. 5, 174177 (1959)
Henderson, N., Brttas, J.C., Sacco, W.F.: A three-parameter KozenyCarman generalized equation for fractal
porous media. Chem. Eng. Sci. 65, 44324442 (2010)
Jin, Y., Song, H.B., Hu, B., Zhu, Y.B., Zheng, J.L.: Lattice Boltzmann simulation of fluid flow through coal
reservoirs fractal pore structure. Sci. China Earth Sci. 56, 15191530 (2013)
Jin, Y., Dong, J.B., Li, X., Wu, Y.: Kinematical measurement of hydraulic tortuosity of fluid flow in porous
media. Int. J. Mod. Phys. C 26, 1550017 (2015)
Kandhai, D., Vidal, D.J.E., Hoekstra, A.G., Hoefsloot, H., Iedema, P., Sloot, P.M.A.: Lattice-Boltzmann and
finite element simulations of fluid flow in a SMRX Static Mixer Reactor. Int. J. Numer. Methods Fluids
31, 10191033 (1999)
Koponen, A., Kataja, M., Timonen, J.: Permeability and effective porosity of porous media. Phys. Rev. E 56,
33193325 (1997)
Kozeny, J.: Uber Kapillare Leitung Des Wassers in Boden. Stizungsber. Akad. Wiss. Wien 136, 271306
(1927)
Krohn, C.E., Thompson, A.H.: Fractal sandstone pores: automated measurements using scanning-electronmicroscope images. Phys. Rev. B. 33, 6366 (1986)
Ladd, A.J.C.: Numerical simulations of particulate suspensions via a discretized Boltzmann equation. Part 2.
Numerical results. J. Fluid Mech. 271, 311339 (1994)
Liu, S., Masliyah, J.H.: Single fluid flow in porous media. Chem. Eng. Commun. 148, 653732 (1996a)
Liu, S., Masliyah, J.H.: Steady developing laminar flow in helical pipes with finite pitch. Int. J. Comput. Fluid
Dyn. 6, 209224 (1996b)
Li, J., Yu, B.: Tortuosity of flow paths through a Sierpinski carpet. Chin. Phys. Lett. 28, 34701 (2011)
Mandelbrot, B.B.: The Fractal Geometry of Nature. Macmillan, New York (1983)
Matyka, M., Khalili, A., Koza, Z.: Tortuosityporosity relation in porous media flow. Phys. Rev. E 78, 026306
(2008)
Mauret, E., Renaud, M.: Transport phenomena in multi-particle systems. Limits of applicability of capillary
model in high voidage beds-application to fixed beds of fibers and fluidized beds of spheres. Chem. Eng.
Sci. 52, 18071817 (1997)
Mortensen, N.A., Okkels, F., Bruus, H.: Reexamination of HagenPoiseuille flow: shape dependence of the
hydraulic resistance in microchannels. Phys. Rev. E 71, 057301 (2005)
Mota, M., Teixeira, J.A., Bowen, W.R., Yelshin, A.: Binary spherical particle mixed beds porosity and permeability relationship measurement. Trans. Fit. Soc. 1, 101106 (2001)
Nasta, P., Vrugt, J.A., Romano, N.: Prediction of the saturated hydraulic conductivity from Brooks and Coreys
water retention parameters. Water Resour. Res. 49, 29182925 (2013)
Nithiarasu, P., Ravindran, K.: A new semi-implicit time stepping procedure for buoyancy driven flow in a fluid
saturated porous medium. Comput. Methods Appl. Mech. 165, 147154 (1998)

123

Scaling Invariant Effects on the Permeability of Fractal Porous

453

Panda, M.N., Lake, L.W.: Estimation of single-phase permeability from parameters of particle-size distribution.
AAPG Bull. 78, 10281039 (1994)
Perrier, E., Bird, N., Rieu, M.: Generalizing the fractal model of soil structure: the pore solid fractal approach.
Geoderma 88, 137164 (1999)
Perrier, E.M.A., Bird, N.R.A.: Modelling soil fragmentation: the pore solid fractal approach. Soil Tillage Res.
64, 9199 (2002)
Pitchumani, R., Ramakrishnan, B.: A fractal geometry model for evaluating permeabilities of porous preforms
used in liquid composite molding. Int. J. Heat Mass Transf. 42, 22192232 (1999)
Qian, Y.H., DHumires, D., Lallemand, P.: Lattice BGK models for NavierStokes equation. Europhys. Lett.
17, 479488 (1992)
Rahli, O., Tadrist, L., Miscevic, M., Santini, R.: Fluid flow through randomly packed monodisperse fibers: the
KozenyCarman parameter analysis. J. Fluids Eng. 119, 188192 (1997)
Rawls, W.J., Brakensiek, D.L., Logsdon, S.D.: Predicting saturated hydraulic conductivity utilizing fractal
principles. Soil Sci. Soc. Am. J. 57, 11931197 (1993)
Sahimi, M.: Flow phenomena in rocks: from continuum models to fractals, percolation, cellular automata, and
simulated annealing. Rev. Mod. Phys. 65, 13931534 (1993)
Sangani, A.S., Acrivos, A.: Slow flow past periodic arrays of cylinders with application to heat transfer. Int.
J. Multiph. Flow 8, 193206 (1982)
Smidt, J.M., Monro, D.M.: Fractal modeling applied to reservoir characterization and flow simulation. Fractals
6, 401408 (1998)
Succi, S.: The Lattice Boltzmann Equation for Fluid Dynamics and Beyond. Oxford, New York (2001)
Sukop, M.C., Thorne, D.T.: Lattice Boltzmann Modeling : An Introduction for Geoscientisits and Engineers.
Springer, New York (2007)
Tarafdar, S., Franz, A., Schulzky, C., Hoffmann, K.H.: Modelling porous structures by repeated Sierpinski
carpets. Phys. A 292, 18 (2001)
Thovert, J.F., Wary, F., Adler, P.M.: Thermal conductivity of random media and regular fractals. J. Appl. Phys.
68, 38723883 (1990)
Valds-Parada, F.J., Porter, M.L., Wood, B.D.: The role of tortuosity in upscaling. Transp. Porous Med. 88,
130 (2011)
Vita, M.C., De Bartolo, S., Fallico, C., Veltri, M.: Usage of infinitesimals in the Mengers Sponge model of
porosity. Appl. Math. Comput. 218, 81878195 (2012)
Wang, B.Y., Jin, Y., Chen, Q., Zheng, J.L., Zhu, Y.B., Zhang, X.B.: Derivation of permeabilitypore relationship
for fractal porous reservoirs using series-parallel flow resistance model and lattice Boltzmann method.
Fractals 22, 1440005 (2014)
Wheatcraft, S.W., Tyler, S.W.: An explanation of scale-dependent dispersivity in heterogeneous aquifers using
concepts of fractal geometry. Water Resour. Res. 24, 566578 (1988)
Xu, P., Yu, B.M.: Developing a new form of permeability and KozenyCarman constant for homogeneous
porous media by means of fractal geometry. Adv. Water Resour. 31, 7481 (2008)
Young, I.M., Crawford, J.W.: The fractal structure of soil aggregates: its measurement and interpretation. Eur.
J. Soil. Sci. 42, 187192 (1991)
Yu, B.M.: Comments on A fractal geometry model for evaluating permeabilities of porous preforms used in
liquid composite molding. Int. J. Heat Mass Transf. 44, 27872789 (2001)
Yu, B.M., Li, J.H., Li, Z.H., Zou, M.Q.: Permeabilities of unsaturated fractal porous media. Int. J. Multiph.
Flow 29, 16251642 (2003)
Yu, B., Zou, M., Feng, Y.: Permeability of fractal porous media by Monte Carlo simulations. Int. J. Heat Mass
Transf. 48, 27872794 (2005)
Yu, B., Cai, J., Zou, M.: On the physical properties of apparent two-phase fractal porous media. Vadose Zone
J. 8, 177186 (2009)
Yu, B.M., Li, J.H.: Some fractal characters of porous media. Fractals 9, 365372 (2001)
Zhou, H., Perfect, E., Li, B., Lu, Y.: Comments on On the physical properties of apparent two-phase fractal
porous media. Vadose Zone J. 9, 192 (2010)

123

Potrebbero piacerti anche