You are on page 1of 12

Journal of Applied Geophysics 162 (2019) 35–46

Contents lists available at ScienceDirect

Journal of Applied Geophysics

journal homepage: www.elsevier.com/locate/jappgeo

The effect of various lengths of pores and throats on the formation


resistivity factor
Haitao Wang a,b,⁎, Jinyan Zhang a
a
Well Logging Company of Shengli Oilfield Service Corporation SINOPEC, Dongying 257096, China
b
School of Geosciences, China University of Petroleum (East China), Qingdao 266580, China

a r t i c l e i n f o a b s t r a c t

Article history: Pore geometry, quantified by the length and size of the pore space, affects the electrical resistivity of the porous
Received 21 March 2018 media. However, the length described by the geometrical tortuosity and pore topology effect on the resistivity of
Received in revised form 30 December 2018 the model does not thoroughly consider the existing pore geometrical models, which only investigate the pore
Accepted 7 January 2019
size, throat length and throat size of the models. A modified capillary model including a pore body and several
Available online 11 January 2019
throats derived the formation factor from the coordination number, the cross-sectional area ratio and geometri-
Keywords:
cal tortuosity of the pores and throats in the model. The derived relationship between the formation factor and
Archie equation porosity of the capillary model compared with the formation factor versus porosity for the various rock types
Modified capillary model was shown in a plot. The geometrical tortuosity of each rock type almost followed the electrical tortuosity of
Tortuosity the related rock type, and the electrical tortuosity was larger than the geometrical tortuosity caused by the var-
Pore-to-throat cross-sectional area ratio iation in the size of the pore and throat. For low porosity, the resistivity of the real porous media was lower than
Porosity exponent the resistivity of the Archie equation due to the low values of the pore-to-throat cross-sectional area ratio and
Formation factor geometrical tortuosity. The sandstones and interparticle carbonates presented a porosity exponent close to 2.0
and an electrical tortuosity from 1.5 to 4.0 on the plot of the porosity exponent versus electrical tortuosity. The
fractured carbonate showed a porosity exponent from 1.0 to 1.6 and an electrical tortuosity ranging from 1.0
to 3.0, while the vuggy carbonates were widely distributed, with a porosity exponent larger than 2.0 and an elec-
trical tortuosity ranging from 2 to 8. This new model correlates the formation factor and porosity to the combi-
nation of the coordination number with the geometrical tortuosity and the throat and pore cross-sectional
area ratios, thus enhancing the understanding of the pore geometry and topology effect on the electrical resistiv-
ity of a porous media with various pore types.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction space. Therefore, there is a relationship between the electrical resistivity


and the pore structure of the porous media in the reservoir.
The quantification of the pore space of porous media is an important The famous relationship relating electrical resistivity to porosity is
issue in petrophysics (Vogel and Roth, 2001; Müller-Huber et al., 2015); known as Archie's first equation (Archie, 1942). The Archie's first equa-
the morphological properties of the pore space in porous media affect tion defines the formation factor, F, as the electrical resistivity ratio of
the reservoir properties such as hydraulic and electrical conductance. the fully saturated clean porous media by brine, R0, to the brine, Rw.
The electrical resistivity of the reservoir is an important physical param- Later, an exponent m is used to relate the formation factor, F, to the
eter, providing information about the formation types, porosity and porosity, ϕ, in the general equation in Eq. (1).
water saturation of the reservoir. The electrical resistivity of porous
media in varied formation types ranges over many orders of magnitude. R0 1
F¼ ¼ ð1Þ
In porous media without Fe-bearing or clay minerals, the electrical Rw ϕm
resistivity of brine-saturated rocks depends predominantly on parame-
ters such as the volume of the pore space swept by an electrical current, where the empirical exponent m is related to the pore geometry of the
the pore morphology characterized by the pore connectivity, the pore porous media and therefore reflects the properties of the pore struc-
shape and size, and the resistivity of the brine saturated in the pore tures (Schön, 2011), known as the cementation exponent or porosity
exponent (Archie, 1942).
⁎ Corresponding author at: Well Logging Company of Shengli Oilfield Service
In sandstones, the theoretical correlation of the formation factor to
Corporation SINOPEC, Dongying 257096, China. the porosity is usually done by introducing parameters in the numerator
E-mail address: wanght_2004@163.com (H. Wang). (Winsauer et al., 1952) or by investigating the sphere packs

https://doi.org/10.1016/j.jappgeo.2019.01.005
0926-9851/© 2019 Elsevier B.V. All rights reserved.
36 H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46

(Perez-Rosales, 1976, 1982). However, in carbonates, the pore space is aspect is the tortuosity of the capillary channel reflecting the real path-
much more complicated due to complex pore structures and various way (Clennell, 1997), defined as the length of the real pore channel
pore types caused by the diagenetic process (Choquette and Pray, flowed by electrical or hydraulic flow in the porous media to the length
1970). The Archie equation is only valid in the carbonate rocks with in- of the porous media. Mualem and Friedman (1991) propose a capillary
terparticle pores, which is similar to the sandstones with intergranular bundle model to calculate the soil electricity, assuming that the hydrau-
pores, while it fails in the carbonate with moldic or vuggy porosity. lic flow is equivalent to electrical flow, and investigate the effects of the
The method to solve this problem is to correlate the typical porosity ex- interface conductivity of the grain surfaces, pore size distributions and
ponent to the dominant pore types in the carbonate rocks. Focke and saturation on the electrical resistivity in soil samples and glass beads.
Munn (1987) point out that the porosity exponent of the generic pore Berg (2012) calculates the electrical conductivity of fully brine saturated
types determined by visual thin section analysis ranges from 2 to over porous medium by tortuosity and constriction of the pore microstruc-
5 according to the electrical resistivity and porosity data. Rasmus ture described by the X-ray computed tomography and applies the
(1983) determines the porosity exponent of the carbonate rocks model to the Bentheimer sandstone.
consisting of vuggy porosity and fracture system. The resistivity is pro- Pore network models based on high-resolution X-ray computed
portionally reduced with the increase of the fracture porosity, and the micro-tomography, which emphasize visual imaging and quantifying
porosity exponent is less than the matrix value of 2 for the interparticle the pore structure of the porous media, are gradually applied to simu-
pores and close to 1.0. Towle (1962) gives similar conclusions by calcu- late the electrical resistivity of the porous media (Bigake, 2000; Arns,
lating the electrical current flow in a synthetic pore system including 2002; Békri and Vizika, 2006; Han et al., 2008; Bauer et al., 2011;
interconnected tubes or planes with a low porosity exponent and Corbett et al., 2017). Abousrafa et al. (2009) theoretically propose an
vuggy pore systems with wide range of porosity exponents. The corre- isotropic network model consisting of a spherical pore node connected
lations between electrical resistivity and porosity of different pore by capillary throat bonds at an angle to derive the electrical resistivity
types derived from a synthetic pore systems are known as double- or (Fig. 1c). The numerical results show that the key parameter affecting
triple- porosity models (Aguilera and Aguilera, 2003; Al-Ghamdi et al., the resistivity is the throat bond radius. While the pore and throat
2011; Wang, 2018). The porosity models explicitly consider the pore radii are fixed, Müller-Huber et al. (Müller-Huber et al., 2015, 2016a,
type but not the pore morphology. 2016b) propose a modified capillary model that describes a rotated
A traditionally applied approach to simulate the electrical and hy- straight pore channel with a cross-sectional area varying between the
draulic flow in porous media is the capillary model (Kozeny, 1927; two extreme values of the Awt throat bond and the Awp pore node,
Carman, 1997).The geometry of the model is composed of a cylinder which follows an exponential function along the pore axis x (Fig. 1b).
pore channel with a cross-sectional area equal to Aw and the length The tortuosity is defined by the rotated angle as 1/cos (θ), which is
equal to L in a cube rock model with a length equal to L and cross- also used by Abousrafa et al. (2009) (Fig. 1c). This new approach derives
sectional area equal to A (Fig. 1a)(Carman, 1997). In this case, the poros- the ratio of the pore to throat radius based on the porosity and forma-
ity exponent derived from this capillary model is equal to 1 less than the tion factor, enhances the understanding of pore space geometries in
porosity exponent value of 2 in the experimental data, indicating that different formation types and obtains a physical explanation for the
the capillary model is an oversimplification of the pore geometry in at parameters used in permeability and formation factor prediction. How-
least two aspects. One aspect is the pore connectivity; the pore and ever, the rotated curve pore capillary model is straight and does not
throat radii should be different (Müller-Huber et al., 2015). The other include more complex cases where tortuous pore channels exist and

Fig. 1. Schematics representation of existing models of the pore space. (a) A simple capillary model presenting a pore channel of length, L, and cross-sectional area, Aw (Carman, 1997);
(b) a modified capillary model with curved pore geometry with a cross-sectional areal of Aw (Müller-Huber et al., 2015, 2016a, 2016b); (c) a spherical pore with a cross-sectional area
of Awp, and cylindrical throats with a cross-sectional area of Awt, connected to the pore body at an angle θ (Abousrafa et al., 2009); (d) a spherical pore with a cross-sectional area of
Awp, and cylinderical throats with a cross-sectional area of Awt, tortuous length of Lwt (Li et al., 2017), residing in a non-porous cube of rock of length, L and cross-sectional area A.
H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46 37

the tortuosity is larger than 1.41 (Müller-Huber et al., 2016a, 2016b). Li The modified capillary model is composed of two tortuous compo-
et al. (2017) present a pore-throat model that treats the pore space as a nents; one component is a throat series, and the other is a pore body.
series of spherical pores described by the pore size and two tortuous The throat is the tortuous channel with a small cross-sectional area,
throats described by the size ratio of the pore to the throat and the Awt, and the pore is the tortuous channel with a large cross-sectional
throat tortuosity (TTt) (Fig. 1d). They build a robust relationship be- area, Awp. A is the cross-sectional area of the capillary model. Lwt and
tween the formation factor and the porosity for the low- and middle- Lwp are the lengths of the tortuous throat and pore channel in the capil-
porosity sandstone formations. However, the tortuosity of the pore lary model, respectively. L is the length of porous capillary model. n is
space (TT), including the pore and throat, especially the tortuous pore, the coordination number, as the number of the throats connected to
is not thoroughly investigated in their model. the pore (Dong, 2007).
Numerous studies have been done to theoretically investigate the The calculation of the electrical resistivity of the modified pore
correlation of electrical resistivity of the porous media fully saturated geometry (Fig. 2) is carried out based on the assumption that the throats
by brine to pore geometrical information. However, the impacts of the are first parallel and then connect to the pore in series in their equiva-
length of the complex pore system and the coordination number on lent circuit (Aguilera and Aguilera, 2003). According to Ohm's law, the
the electrical resistivity do not incorporate the existing pore geometrical resistivity of the capillary model fully saturated by brine is Eq. (2).
models. The findings of the aforementioned investigations demonstrate 0 1
that pore geometry strongly affects electrical current flow and that a
L B Lwp 1 C
theoretically geometrical model concept urgently needs to introduce a R0 ¼ Rw B
@Awp þ Pn Awti A
C ð2Þ
variable tortuosity defined by the lengths of the tortuous pore and A
i¼1
throat, and a coordination number. This study aims to theoretically de- Lwti
rive a relationship between resistivity and porosity by introducing a
modified capillary model. The theoretical modelling results and the In this case, the porosity of the rock model is given by Eq. (3).
Archie first equation are applied to build diagrams to investigate the re- !
X
n
lationship between the formation factor and porosity, and the variation ϕ¼ Awp Lwp þ Awti Lwti =ðALÞ ð3Þ
of the tortuosity effect on the porosity exponent. A comparison with the i¼1
experimental data is presented to finally infer details about the impact
of the tortuosity on the electrical current flow for sandstones and According to the definition of the formation factor in Eq. (1), the for-
carbonates with different pore types. The modified capillary model pre- mation factor of the modified capillary model is Eq. (4).
sented in the next section is a first step to enhance our knowledge about
the pore space structures and resistivity and to reach the above-stated R0 Lwp =L 1
F¼ ¼ þ ð4Þ
aims. Rw Awp =A n Awti =A
∑i¼1
Lwti =L
2. A modified capillary model
The pore and throat cross-sectional area ratio, ψp, and ψt, are defined
A capillary tube model with a constant size or a straight channel does as.
not represent the complicated pore morphology of the porous media in
the reservoir; a more complicated pore geometrical model with various Awp Awt
ψp ¼ ψt ¼ ψp ; ψt ∈½0; 1 ð5Þ
radii and tortuous channels of pore and throat as shown in Fig. 2 is A A
investigated in this section.
The pore geometrical tortuosity, τgp, and throat geometrical tortuos-
ity, τgt, are defined as.

Lwp
τ gp ¼
L ð6Þ
Lwt
τ gt ¼
L

and τgp, τgt ∈ [0, + ∞].


Based on the definition of ψp, ψt, τgp and τgt, the formation factor and
porosity are expressed in Eq. (7) and Eq. (8), respectively.

R0 τgp 1
F¼ ¼ þ ð7Þ
Rw ψp n ψti
∑i¼1
τ gti

n
ϕ ¼ ψp τ gp þ ∑i¼1 ψti τgti ð8Þ

For a simpler case, only one pore and several identical throat chan-
nels exist in the capillary tube model, and the porosity and formation
factor can be calculated using Eq. (9) and Eq. (10).

R0 τgp τ gt
F¼ ¼ þ ð9Þ
Rw ψp nψt

ϕ ¼ ψp τ gp þ nψt τgt ð10Þ

Fig. 2. Schematic representation of the modified capillary model including several throats
and a pore channel in the rock model (a) and the equivalent electrical circuit of the Additionally, there is a limitation between the coordination number,
capillary rock model (b). n, the cross-sectional area ratios of the pore and throat, ψp, ψt, and the
38 H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46

geometrical tortuosity, τgp, τgt. and throat in Eq. (9). The formulation allows a theoretical analysis of the
coordination number, the cross-sectional area and the tortuosity of
0≤ψp τ gp þ nψt τ gt ≤1 ð11Þ the pore space's separate effect on the electrical formation factor. In
the following investigation, the tortuosity will be analysed for the differ-
Based on the pore and throat geometrical tortuosity in Eq. (6), the ent formation types.
geometrical tortuosity of the capillary tube is calculated by Eq. (12).
3. Calculated data from model and comparison with experimental
Pn
i¼1 Lwti data
Lwp þ P n
i¼1 i Lwp þ Lwt
τg ¼ ¼ ¼ τ gp þ τgt ð12Þ
L L First, the formation factor versus the fractional porosity plot is
directly calculated for the various geometrical tortuosities, TT (τg),
Therefore, the formation factor is given as the function of the coordi- with the same pore-to-throat cross-sectional area ratio, PTAR (ψp/ψt),
nation number, the cross-sectional area ratio and tortuosity of the pore and the coordination number equal to 1, in Fig. 3. The results are a series

Fig. 3. Plot of the electrical formation factor versus the fractional porosity. Black solid lines follow Archie's first equation with various porosity exponents, and the coloured curves are
simulated by the modified capillary model for various geometrical tortuosities and cross-sectional areas of the pore and throat. The solid coloured curves correspond to small pore
space, while the dotted coloured curves correspond to large pore space, but both series have the same PTAR of 10 and the same coordination number of 1. Experimental data include
those of the sandstones, SA (Sawyer et al., 2001), AR (Archie, 1942), LW (Li et al., 2017), IM1 and IM2 (Dong, 2007) in (a), and those of the carbonates, VE (Verwer et al., 2011), FM1–
3, FM4, FM5, FM6 (Focke and Munn, 1987), EU (Müller-Huber et al., 2015), and PET (Corbett et al., 2017) in (b).
H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46 39

of curves; the solid curves with different colours correspond to the case The formation factor versus the fractional porosity plot is also calcu-
in which PTAR is equal to 10 and CSARp (cross-sectional area ratio of the lated for the various coordination numbers with the same geometrical
pore, ψp) is equal to 0.1, and the dot curves correspond to the case in tortuosity (TT) and the same pore-to-throat cross-sectional area ratio
which PTAR is equal to 10 and CSARp is equal to 0.2. By increasing the (PTAR) in Fig. 4. The results obtained are also a series of curves; the
tortuosity in the geometry, the formation factor increases at a constant solid curves (ZTT1) with different colours correspond to the case in
porosity. The formation factor of the larger pore system (TTL) is less which PTAR is equal to 10 and TT is equal to 1, and the dotted curves
than the formation factor of the pore system with a smaller pore size (ZTT3) correspond to the case in which PTAR is equal to 10 and TT is
(TTS). Additionally, the straight lines for a different porosity exponent equal to 3. The formation factor decreases with increasing coordination
m following the Archie first equation are plotted. Clearly, using m = number at a constant porosity, which conforms to the proposed
2.5 as an example, a lower porosity relates to a smaller size of the relationship between formation factor and coordination number,
pore space and a higher geometrical tortuosity (e.g. tortuosity TT = 6 1/F∝(z-zc)γ (Bernabé et al., 2011), where zc is the critical coordination
for porosity equal to 8% for CSARp = 0.1), and a higher porosity relates number for percolation (zc ≈ 2 and 1.5 in two- and three-dimensional
to a larger size of the pore space and a lower tortuosity (e.g. tortuosity models, respectively) and γ (γ N 1.29) is related to the variation in the
TT = 2 for porosity equal to 30% for CSARp = 0.2). pore and throat size (Li et al., 2015). In this modified capillary model,

Fig. 4. Plot of the electrical formation factor versus the fractional porosity. Black solid lines follow Archie's first equation with various porosity exponents, and the coloured curves are
simulated by the modified capillary model for various coordination numbers and geometrical tortuosities. The solid coloured curves correspond to small tortuosities, while the dotted
coloured curves correspond to large tortuosities, but both series have the same PTAR of 10. Experimental data include those of sandstones, SA (Sawyer et al., 2001), AR (Archie, 1942),
LW (Li et al., 2017), IM1 and IM2 (Dong, 2007) in (a), and those of the carbonates, VE (Verwer et al., 2011), FM1–3, FM4, FM5, FM6 (Focke and Munn, 1987), EU (Müller-Huber et al.,
2015), and PET (Corbett et al., 2017) in (b).
40 H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46

when the coordination number is greater than zero, the capillary model Table 2
percolates, and then the critical coordination number zc = 0. The forma- The experimental samples and the associated ranges of formation factor, F, porosity and
porosity exponent.
tion factor of a much more tortuous pore system (ZTT3) is larger than
that of a pore system with less tortuosity (ZTT1). Using m = 2.0 as an Formation factor Porosity (fraction) Porosity exponent Samples
example, a lower porosity relates to a smaller coordination number of 5.96…93.72 0.116…0.402 1.53…2.47 AR
the pore space (e.g. coordination number ZTT3 = 1 for porosity equal 4.80…5.80 0.330…0.429 1.56…1.87 IM1
to 4% for tortuosity TT = 3), and a higher porosity relates to a larger co- 6.16…82.77 0.141…0.340 1.66…2.50 IM2
25.56…328.40 0.031…0.204 1.59…2.06 LW
ordination number (e.g. coordination number ZTT3 = 4 for porosity
19.70…96.20 0.094…0.213 1.67…2.09 SA
equal to 10% for tortuosity TT = 3). 22.45…608.90 0.006…0.062 1.01…1.64 EU
The experimental data (Table 1 and Table 2) are also presented in 11.17…1837.66 0.011…0.311 1.62…2.22 FM1–3
Fig. 3 and Fig. 4. The experimental data consist of sandstones and car- 14.74…764.74 0.060…0.374 2.12…5.69 FM4
bonates. The sandstones are from five formation types and are denoted 16.87…1466.4 0.049…0.360 2.01…3.83 FM5
10.65…505.58 0.041…0.328 1.61…2.37 FM6
as SA, AR, LW, IM1 and IM2. The carbonates consist of seven formation 45.72…285.82 0.040…0.195 1.70…3.00 PET
types and are denoted as VE, FM1–3, FM4, FM5, FM6, EU and PET. 11.00…4590.00 0.010…0.320 1.72…4.14 VE
Table 1 describes the formation types and the petrophysics of each sam-
Key: AR = (Archie, 1942); IM1, IM2 = (Dong, 2007); LW = (Li et al., 2017); SA = (Sawyer
ple and Table 2 summarizes the associated ranges of the formation fac- et al., 2001); EU = (Müller-Huber et al., 2015); FM1–3, FM4, FM5, FM6 = (Focke and
tor, porosity and porosity exponent of each sample. This shows that any Munn, 1987); PET = (Corbett et al., 2017); VE = (Verwer et al., 2011)

Table 1
Formation types and the petrophysical description of the experimental samples.
formation factor value can be expressed as the combination of a given
porosity, a corresponding geometrical tortuosity with a specific CSARp
Formation type and petrophysical description Sample and CSARt (cross sectional area ratio of the throat, ψt).
Consolidated sandstones from Nacatoch sand in the Bellevue area, AR Key: AR = (Archie, 1942); IM1, IM2 = (Dong, 2007); LW = (Li et al.,
Louisiana, which present a strongly linear relationship between the 2017); SA = (Sawyer et al., 2001); EU = (Müller-Huber et al., 2015);
logarithm of the porosity and the logarithm of the formation factor
FM1–3, FM4, FM5, FM6 = (Focke and Munn, 1987); PET = (Corbett
with a porosity exponent close to 2.0.
Low- and middle-porosity sandstone composed of quartz, calcite (1.81%), LW et al., 2017); VE = (Verwer et al., 2011);
siliceous (3.11%), and clays (9.11%) from the SHZ formation in the SLG In Fig. 3, for the sandstone and sand pack data, there is a clear tendency
gas field located in the Yishan Slope of the Ordos basin in China, which for a smaller pore size and higher geometrical tortuosity for the consoli-
shows a linear relationship between the formation factor and porosity dated sandstones (SA, AR, IM2, LW) compared to a larger pore size and
in the logarithm, with the curve bending down (the porosity exponent
reduces) as the porosity approaches 0.08.
lower geometrical tortuosity for the unconsolidated sediments (IM1). For
Sand pack from the LV64A-C, F42A-C and synthetic sand pack A1 from IM1 carbonate, the fractured carbonate (EU) corresponds to a lower porosity
the Imperial College, London, with a high porosity and high permeabil- and lower geometrical tortuosity, the carbonate with interparticle pores
ity reflecting a good connectivity. (FM1–3, FM6) shows the same tendency as the sandstone, and the carbon-
Sandstones S1 to S8 and Berea sandstone from the Imperial College, IM2
ate with moldic (FM4, FM5, VE) presents a larger geometrical tortuosity. In
London. Berea sandstone contains minor amounts of feldspar, dolomite
and clays in middle layer of the Waverly group, Ohio and Michigan Fig. 4, for the sandstone and sand pack data, there is a clear trend for a
Basin from the Mississippian geological period (360–325 million years smaller coordination number and lower porosity rather than a larger coor-
ago). dination number and higher porosity for the same geometrical series. Data
Appalachian sandstones from the Big Injun, Clinton, Bradford Third, SA for carbonates show the same trend as those for the sandstone. Take FM5
Venango Second and Gordon Formation, USA, do not follow the Archie
first equation. Significant anisotrophy of resistivity that is correlated to
as an example; its tortuosity is equal to 3.0, while the coordination number
permeability exists in the Bradford Third, Clinton and Gordon increases from 1 to 4 with the formation factor decreasing from 1000 to 10
sandstones. and the porosity increasing from 5% to 30%.
Upper Triassic Lagoonal Wetterstein dolomite with vuggy and bird's eye EU Additionally, for porous media with interparticle pores, including
porosity and fractures in Australia.
the carbonate and sandstone, the trend of the formation factor versus
The limestone and dolomite grainstones with intergranular and FM1-3
intercrystalline pores from Qatar; the formation factors versus the the porosity follows the Archie equation when the porosity is larger
porosities in the bi-logarithm plot is a straight line. At a low porosity than 0.1, while the formation resistivity factor of the Archie equation
(below 0.05), the line bends down, and the porosity exponent is b2.0. is greater than that of the real rocks in the low porosity range, known
Moldic limestones from Qatar with well-developed moldic pores formed FM4 as “non-Archie” phenomenon (Liu et al., 2009; Ziarani and Aguilera,
by cement that destroys the original pores and the dissolution of the
grains; the porosity exponent increases from 2.0 to 5.4 when the
2012; Li et al., 2017).
modlic porosity increases. The tortuosity calculation is related to the investigated property
Moldic dolomites from Qatar, in which recrystallization takes place in the FM5 (Clennell, 1997). The geometrical tortuosity has been extracted as a
dolomite. The porosity exponent values increase with the increasing pore space property (Clennell, 1997; Li et al., 2017), while the electrical
moldic porosity. The lower permeability samples generally have larger
or hydraulic tortuosity acts as a petrophysical parameter (Katsube,
values of the formation factor than the higher-permeability samples.
The mudstones and chalks with a matrix pore without any significant FM6 2010) affecting the electricity or permeability, respectively. The electri-
moldic, vuggy, fracture or fissure pores, and the porosity exponents are cal resistivity has been postulated to affect the separate vug porosity
approximately 2.0. (Lucia and Conti, 1987), the size of the pore space (Abousrafa et al.,
Calcirudite Coquina that includes well-connected matrix pores and PET 2009) and the electrical tortuosity (Saner et al., 1996). The increase of
infilled shell moulds, solution-seam related fracture-like pores, and
non-connected moldic pores, from the Morro do Chaves formation in
the tortuous path flowed by electrical current has been found to in-
the Sergipe-Alagoas Basin in Brazil. It is from the Cretaceous, and the crease the electrical resistivity. Due to the difficulty in quantifying the
porosity exponent increases with the increase of the permeability geometrical and electrical tortuosity (Verwer et al., 2011), several au-
Carbonates samples with less 2% non-carbonate minerals from the Khuff VE thors have empirically found the relationships between the electrical
and the Shu'aiba formations from the Permian and Cretaceous periods
tortuosity and the formation factor (Table 3). The Pirson's model
in the Middle east, from the Marion Plateau from the Miocene in
Australia, from the Maiella Mountain from the Cretaceous in Italy, and (Pirson, 1983) is usually applied to estimate the electrical tortuosity
from Holocene stromatolites from the Bahamas; the textures of these by the formation factor and porosity of the porous media at least in
carbonate samples range from coarse-grained grainstones, to interpar- terms of orders of magnitude (Ziarani and Aguilera, 2012). Hence, the
ticle pores, to oomoldic pores, to fine-grained wackestone, but they electrical tortuosities of the published cores were calculated by Pirson's
dominantly develop interparticle to intercrystalline pores.
model, which is a square root model, in this investigation.
H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46 41

Table 3 Second, the relationship between the formation factor and porosity
Electrical resistivity models for tortuosity. are determined by two approaches: the Archie's first equation in Eq.
Reference Equation (1) and the modified capillary model as given in Eq. (9) and (10).
Wyllie and Spangler (1952)
These equations are used to calculate the empirical porosity exponent
τ e 2 ¼ ðFϕÞ2
Winsauer et al. (1952)
m as a function of pore geometry by the cross-sectional area ratio and
τ e 2 ¼ ðFϕÞ1:2
Cornell and Katz (1953) τe = Fϕ
tortuosity of the pore and throat, and coordination number given by
Faris et al. (1954) τ e 2 ¼ ðFϕÞ1:41 Eq. (13),
Pirson (1983) τe = (Fϕ)0.5
Katsube (2010) τ e ¼ ðFϕ=b fÞ
0:5
!

1:5 sheet‐like connecting pores τ gp τgt  
bf ¼
3:0 circular connecting pores m ¼ −lo g þ =lo g ψp τgp þ nψt τgt ð13Þ
ψp nψt
bf depending on connecting pore shape

Fig. 5. Plot of the electrical porosity exponent versus the fractional porosity including the experimental data and the coloured curves simulated by the modified capillary model for various
coordination numbers and geometrical tortuosities. The solid coloured curves correspond to the small tortuosity, while the doted coloured curves correspond to large tortuosity, and both
series have the same PTAR of 10. Experimental data include those of the sandstones, SA (Sawyer et al., 2001), AR (Archie, 1942), LW (Li et al., 2017), IM1 and IM2 (Dong, 2007) in (a), and
those of the carbonates, VE (Verwer et al., 2011), FM1–3, FM4, FM5, FM6 (Focke and Munn, 1987), EU(Müller-Huber et al., 2015), and PET (Corbett et al., 2017) in (b).
42 H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46

Fig. 5 shows the calculated plot of the porosity exponent m versus (IM1) is generally associated with a low geometrical tortuosity b1.5
porosity for various coordination numbers and different geometrical and high porosities causing low porosity exponents close to 1.6. Sand-
tortuosities with a PTAR of 10, as well as the experimental data. For stones (SA, AR and IM2) possess the geometrical tortuosity of 1.5 to
the sand pack and sandstones, unconsolidated sand (IM1) with low 3.0 and a low to intermediate porosity resulting in an intermediate po-
geometrical tortuosity and sandstones (SA, AR and IM2) with geometri- rosity exponent m of 1.5 to 2.4. In contrast, for carbonates, fractured car-
cal tortuosity close to 2 both show that increasing coordination number bonates (EU) are usually associated with a low geometrical tortuosity
decreases the porosity exponent. The same trend is observed for ranging from 1 to 2 and low porosities contributing to a low porosity
carbonates. Fig. 6 shows the calculated plot of the porosity exponent exponent close to 1.3. Intergranular or intercrystalline carbonates
m versus porosity for different geometrical tortuosities with coordina- (FM1-3, FM6) present a geometrical tortuosity from 2 to 3 and a low
tion number equal to 1 in sandstone and carbonate as well as the exper- to intermediate porosity resulting in an intermediate porosity exponent
imental data. For the sand pack and sandstones, unconsolidated sand m of 1.8 to 2.0. The moldic or oolidic (vuggy) carbonates (FM4, FM5, PET

Fig. 6. Plot of the electrical porosity exponent versus the fractional porosity including the experimental data and the coloured curves simulated by the modified capillary model for various
geometrical tortuosities and cross-sectional areas of the pore and throat. The solid coloured curves correspond to small pore space, while the dotted coloured curves correspond to large
pore space, and both series have the same PTAR of 10 and the same coordination number of 1. Experimental data include those of the sandstones, SA (Sawyer et al., 2001), AR (Archie,
1942), LW (Li et al., 2017), IM1 and IM2 (Dong, 2007) in (a), and those of the carbonates, VE (Verwer et al., 2011), FM1–3, FM4, FM5, FM6 (Focke and Munn, 1987), EU(Müller-Huber
et al., 2015), and PET (Corbett et al., 2017) in (b).
H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46 43

and VE) occupy a high geometrical tortuosity larger than 3.0 and high Table 4
porosity resulting in a high porosity exponent m of 2.0 to 6.0. The pre- Summary of porosity and electrical tortuosity suggested in Klinkenberg (1951).

dicted values of the geometrical tortuosity of the different formation Formation types Porosity (fraction) Tortuosity,τe
types correspond to the electrical tortuosity of Klinkenberg (1951) pre- Unconsolidated sands and glass beads 0.10…0.47 1.43…1.96
sented in Table 4. Various consolidated sandstones 0.19…0.26 2.00…4.02
These findings explain the observations presented in earlier works: Unconsolidated sand packs 0.35…0.38 1.41…1.70
Focke and Munn (1987) and Verwer et al. (2011) found that carbonates Various limestones 0.11…0.34 3.40…3.70
with higher moldic porosity were associated with an increased porosity
exponent between 1.8 and 5.0, which is likely caused by their larger
pore size and low connectivity, and hence a high cross-sectional area tortuosity. The sand pack (IM1), sandstones with interparticle pores
ratio of the pore-to-throat (PTAR) and high tortuosity. The larger pore (SA, AR and IM2) and carbonates with intercrystalline pores (FM1-3,
size in the pore system leads to a greater potential volume for facilitat- FM6) seem to occupy well defined positions with porosity exponents
ing the electrical current flow and hence an overall lower resistivity of close to 2.0 and an electrical tortuosity from 1.5 to 4.0 on the plot. The
the porous medium. As a result, more tortuous paths are accessible to fractured carbonate (EU) locates in the area with a porosity exponent
electrical current at higher than at lower PTAR. According to Eq. (9), from 1.0 to 1.6 and the electrical tortuosity ranging from 1.0 to 3.0,
Eq. (10) and Eq. (12), as well as the empirical model of the electrical tor- while the vuggy carbonates (FM4, FM5, VE and PET) widely distribute,
tuosity in Table 3, the electrical tortuosity based on the formation factor with porosity exponents larger than 2.0 and the electrical tortuosity
and porosity of the capillary model was calculated in Eq. (14), almost ranging from 2 to 8.
! Additionally, it is also shown that each value of the porosity is related
2 nψt ψp to a specific combination of the porosity exponent, the size of the pore
F  ϕ ¼ τgp þ τ gp τgt þ þ τgt 2
ψp nψt space expressed by the PTAR and CSARp, and the tortuosity. Therefore,
! it is beneficial for petrophysicists to determine the porosity from the
nψt ψp
2
¼ τgp þ 2τ gp τgt þ τgt þ τ gp τ gt 2
þ −2 ð14Þ electrical resistivity according to the classification of the combination
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} ψp nψt
Part1 |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} of the porosity exponent, pore space size and tortuosity dominating in
Part2 a particular formation type or depth section.
≥τ gp 2 þ 2τgp τ gt þ τ gt 2 ¼ τg 2
4. Conclusions
Electrical tortuosity depends on two aspects: one is the length of the
tortuous pore space swept by the electrical flow in the geometry and the In this investigation, a new capillary model describing electrical cur-
geometrical tortuosity (part1), and the other is the electrical tortuosity rent flow in tortuous channels with several different cross-sectional
caused by the variation between the pore and throat in size reflected area ratios and various tortuous throat channels is presented. The vari-
by part2. From Eq. (14), it is clear that the electrical tortuosity is larger able cross-sectional area ratios include the cross-sectional area ratio of
than the geometrical tortuosity, which is caused by the PTAR. The larger the pore, CSARp, and the cross-sectional area ratio of the throat,
PTAR caused the larger ratio of the electrical tortuosity to the geometri- CSARt. The lengths of the tortuous pore and throat channels define the
cal tortuosity (Fig. 7). geometrical tortuosity of the model. The number of the tortuous throats
For the fracture porosity, Rasmus (1983) stated that the porosity ex- defines the coordination number. To theoretically investigate the rela-
ponent is lower than the ‘matrix’ value of 2.0 associated with interpar- tionship between the pore structure and electrical resistivity, the Archie
ticle porosity and may even approach 1.0. This behavior is likely porosity exponent m is presented as a function of the pore geometry
related to the shape of a fracture with low values of PTAR and geomet- expressed by the cross-sectional area ratios and tortuosities of the
rical tortuosity (Ziarani and Aguilera, 2012). The non-Archie phenome- pore and throat, and the pore topology expressed by coordination num-
non is obvious when the porosity is close to 10%; that the resistivity ber. The theoretical calculations of the formation factors and porosities
formation factor versus porosity does not follow Archie first equation were compared with the experimental formation factors and porosities
is due to the existence of the sheet pore (Ziarani and Aguilera, 2012) of the carbonates and sandstones. The comparison confirms that the im-
causing a smaller PTAR (Fig. 8) and smaller tortuosity and then reducing portance of introducing the various pore sizes and lengths instead of a
the electrical resistivity at the tight porous media. In Fig. 8, the geomet-
rical tortuosity is equal to 1.0 when the porosity is larger than 8%, and
the PTAR of the sandstones reduces from 100 to 4 with the porosity
reaching 50%. However, when the porosity is less than the 8%, the
PTAR for the Archie rock (Kennedy and Herrick, 2012), referred to as
the rock following the formation factor versus porosity trends described
by the Archie first equation in Eq. (1), increases from 100 to 400, while
the PTAR of the sandstones follows 100. Thus the pore geometry of the
Archie rock is complicated compared with the real rocks when the po-
rosity is small (Li et al., 2017).
Third, a diagram of the porosity exponent m versus the electrical
tortuosity,τe, is presented in Fig. 9. The solid lines representing the var-
ious porosities according to Eq. (15) were calculated by Pirson's Model
(Pirson, 1983) and the Archie's first Eq. (1).

m ¼ 1−2 logðτe Þ
ð15Þ
logðϕÞ

The electrical tortuosities of the investigated data sets were esti- Fig. 7. Plot of the tortuosity ratio of the electricity to geometry versus the fractional
porosity for two PTARs. The electrical tortuosity calculated by Pirson model (Pirson,
mated by Pirson's model by using the experimental formation factor 1983) with the formation factor and porosity from the proposed capillary model in Eq.
and porosity as input parameters. Fig. 9 highlights that there is no sim- (9) and Eq. (10) with coordination number equal to 1, and the geometrical tortuosity
ple relationship between the porosity exponent m and the electrical calculated by Eq. (12).
44 H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46

Fig. 8. Plot of the electrical formation factor versus the fractional porosity. Black solid lines follow Archie's first equation with various porosity exponents, and the coloured solid lines are
simulated by the modified capillary model for various PTARs with all geometrical tortuosity equal to 1.0 and coordination number equal to 1. Experimental data include those of the
sandstones, SA (Sawyer et al., 2001), AR (Archie, 1942), LW (Li et al., 2017), IM1 and IM2 (Dong, 2007), and those of the interparticle carbonates, FM1-3 and FM6 (Focke and Munn, 1987).

fixed size and length to a theoretical geometrical model to investigate Additionally, this model has a potential to illustrate the effects of the
the pore structure effect on the petrophysical properties. This modified coordination number on the electrical resistivity.
capillary model plays a role of a basic unit for composition of compli- The observed consistence between the geometrical tortuosity de-
cated models, such as the pore network model representing the pore rived from the modified capillary model and the values of the electrical
space of the porous media. The new capillary model's first result illus- tortuosity from the published literature for the different formation
trates the potential of quantifying the pore geometry by the cross- types need validation in the laboratory, including determination of the
sectional area ratio of the pore and throat (CASRp (ψp) and CSARt pore and throat size as well as the tortuosity via prevailing technologies
(ψt)) and the geometrical tortuosity of the pore and throat (TTp (τgp) such as nano or micro X-ray computed tomography, image processing,
and TTt (τgt)), and building a theoretical correlation between the numerical simulations (random walk simulation determining the tortu-
electrical formation factor and the pore geometrical parameters. osity) and a pore network analysis based on the extraction of the pore

Fig. 9. Plot of the electrical porosity exponent versus the electrical tortuosity of the experimental data including the solid black lines simulated with various porosities. The electrical
tortuosity of the experimental data were calculated by Pirson (Pirson, 1983) and the solid black lines were calculated by Eq. (15). The mean, minimum and maximum values of the
porosity exponent and electrical tortuosity of each investigated data sets are presented in the plot. Experimental data include those of the sandstones, SA (Sawyer et al., 2001), AR
(Archie, 1942), LW (Li et al., 2017), IM1 and IM2 (Dong, 2007), and those of the carbonates, VE (Verwer et al., 2011), FM1–3, FM4, FM5, FM6 (Focke and Munn, 1987), EU (Müller-
Huber et al., 2015), and PET (Corbett et al., 2017).
H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46 45

network from the reconstructed porous media from X-ray computed Békri, S., Vizika, O., 2006. Year, Pore-networkmodeling of rock transport properties: appli-
cation to a carbonate. International Symposium of the Society of Core Analysts,
tomography. Trondheim, Norway, 12–16, September.
Berg, C.F., 2012. Re-examining Archie's law, conductance description by tortuosity and
constriction. Phys. Rev. E 86 (4). https://doi.org/10.1103/PhysRevE.86.046314
046314,046311-046314,046319.
Nomenclature
Bernabé, Y., Zamora, M., Li, M., Maineult, A., Tang, Y.B., 2011. Pore connectivity, permeabil-
ity, and electrical formation factor: a new model and comparison to experimental
A Cross-sectional area of capillary model data. J. Geophys. Res. 116 (B11204), 1–15. https://doi.org/10.1029/2011JB008543.
Bigake, J., 2000. A study concerning the conductivity of porous rock. Phys. Chem. Earth
Awp Cross-sectional area of pore in the capillary model
Part A Solid Earth Geodesy 25 (2), 189–194. https://doi.org/10.1016/S1464-1895
Awt Cross-sectional area of throat in the capillary model
(00)00030-2.
CSARp Cross-sectional area ratio of the pore channel to the capillary model
Carman, P.C., 1997. Fluid flow through granular beds. Chem. Eng. Res. Des. 75, S32–S48.
CSARt Cross-sectional area ratio of the throat channel to the capillary model https://doi.org/10.1016/S0263-8762(97)80003-2 Supplement.
F Formation factor Choquette, P.W., Pray, L.C., 1970. Geologic nomenclature and classification of porosity in
L Length of the capillary model sedimentary carbonates. AAPG Bull. 54 (2), 207–250.
Lwp Length of the tortuous pore channel in the pore system of the capillary Clennell, M.B., 1997. Tortuosity: a guide through the maze. Geol. Soc. Lond., Spec. Publ.
model 122 (1), 299–344. https://doi.org/10.1144/GSL.SP.1997.122.01.18.
Lwt Length of the tortuous throat channel in the throat system of the capillary Corbett, P.W.M., Wang, H., Câmara, R.N., Tavares, A.C., Borghi de Almeida, L.F., Perosi, F.,
model Machado, A., Jiang, Z., Ma, J., Bagueira, R., 2017. Using the porosity exponent
m Porosity exponent (m) and pore-scale resistivity modelling to understand pore fabric types in coquinas
n Number of the throats connected to the pore channel in the capillary (Barremian-Aptian) of the Morro do Chaves Formation, NE Brazil. Mar. Pet. Geol. 88,
model 628–647. https://doi.org/10.1016/j.marpetgeo.2017.08.032.
PTAR Cross-sectional area ratio of the pore-to-throat Cornell, D., Katz, D.L., 1953. Flow of gases through consolidated media. Ind. Eng. Chem. 45
(10), 2145–2152. https://doi.org/10.1021/ie50526a021.
R0 Resistivity of the porous media fully saturated by brine
Dong, H., 2007. Micro CT Imaging and Pore Network Extraction. PhD. Imperical College
Rw Resistivity of the brine in the porous media
London.
x pore axis in the rotated curved pore channel Faris, S.R., Gournay, L.S., Lipson, L.B., Webb, T.S., 1954. Verification of tortuosity equations:
TT Geometrical tortuosity of the capillary model geological notes. AAPG Bull. 38 (10), 2226–2232.
TTL Geometrical tortuosity of the capillary model with large cross-sectional Focke, J.W., Munn, D., 1987. Cementation exponents in middle eastern carbonate reser-
area voirs. SPE Form. Eval., 155–167 https://doi.org/10.2118/13735-PA.
TTS Geometrical tortuosity of the capillary model with small cross-sectional Han, M., Tariel, V., Youssef, S., Rosenberg, E., Fleury, M., Levitz, P., 2008. the effect of
area the porous structure on resistivity index curves. An Experimental And Numer-
TTp Geometrical tortuosity of the pore channel ical Study. Paper presented at the 49th Annual Logging Symposium, Austin,
TTt Geometrical tortuosity of the throat channel Texas, 25–28, May.
z Coordination number Katsube, T.J., 2010. Review of formation resistivity factor equations related to new pore-
ZTT1 Coordination number of the capillary model with tortuosity equal to 1 structure concepts. Geol. Surv. Canada Curr. Res. 1–9 2010-7.
ZTT3 Coordination number of the capillary model with tortuosity equal to 3 Kennedy, D., Herrick, D.C., 2012. Conductivity models for Archie rocks. Geophysics 77 (3),
WA109–WA128. https://doi.org/10.1190/geo2011-0297.1.
zc Critical coordination number for percolation
Klinkenberg, L.J., 1951. Analogy between diffusion and electrical conductivity in porous
θ Angle between the pore body and the pore throat
media. Bull. Geol. Soc. Am. 62, 559–564.
ϕ Porosity of the capillary model
Kozeny, J., 1927. Über kapillare Leitung desWassers im Boden. Sitzungsber Akad. Wiss,
γ Exponential index Wien 136 (2a), 271–306.
τe Electrical tortuosity of the porous media calculated by models Li, M., Tang, Y.B., Bernabé, Y., Zhao, J.Z., Li, X.F., Bai, X.Y., Zhang, L.H., 2015. Pore connectiv-
τg Geometrical tortuosity of the capillary model ity, electrical conductivity, and partial water saturation: network simulations.
τgp Geometrical tortuosity of the pore channel J. Geophys. Res. 120 (6), 4055–4068. https://doi.org/10.1002/2014JB011799.
τgt Geometrical tortuosity of the throat channel Li, W., Zou, C., Wang, H., Peng, C., 2017. A model for calculating the formation resistivity
ψ Cross-sectional area ratio of the pore or throat to the capillary model factor in low and middle porosity sandstone formations considering the effect of
ψp Cross-sectional area ratio of the pore channel to the capillary model pore geometry. J. Pet. Sci. Eng. 152, 193–203. https://doi.org/10.1016/j.
ψt Cross-sectional area ratio of the throat channel to the capillary model petrol.2017.03.006.
Liu, X., Sun, J., Wang, H., 2009. Numerical simulation of rock electrical properties based on
digital cores. Appl. Geophys. 6 (1), 1–7. https://doi.org/10.1007/s11770-009-0001-6.
Lucia, F.J., Conti, R.D., 1987. Rock Fabric, Permeability, and Log Relationships in an
Upward-Shoaling, Vuggy Carbonate Sequence. University of Texas Bureau of Eco-
Acknowledgement nomic Geology, Austin, TX (Geological Circular).
Mualem, Y., Friedman, S.P., 1991. Theoretical prediction of electrical conductivity in satu-
This work is funded by the Ministry of Science and Technology of the rated and unsaturated soil. Water Resour. Res. 27 (10), 2771–2777. https://doi.org/
10.1029/91WR01095.
People's republic of China (No. 2016ZX05006-002), China Postdoctoral Müller-Huber, E., Schön, J., Börner, F., 2015. The effect of a variable pore radius on forma-
Science Foundation (No.2018M632716). China Postdoctoral Foundation tion resistivity factor. J. Appl. Geophys. 116, 173–179. https://doi.org/10.1016/j.
(No. GKB1611), and Shandong Province Postdoctoral Innovative Special jappgeo.2015.03.011.
Müller-Huber, E., Schön, J., Börner, F., 2016a. A pore body-pore throat-based capillary ap-
Fundation. The Imperial College London is appreciated for providing the
proach for NMR interpretation in carbonates using the coates equation. Paper presented
porous media from S1 to S9 (IM2), F42A, F42B, F42C, LV60A, LV60B, at the SPWLA 57th Annual Well Logging Symposium, Reykjavik, Iceland, 25–29, June.
LV60C and A (IM1). The author thanks the anonymous reviewers for Müller-Huber, E., Schön, J., Börner, F., 2016b. Pore space characterization in carbonate
their helpful comments. rocks — Approach to combine nuclear magnetic resonance and elastic wave velocity
measurements. J. Appl. Geophys. 127 (Supplement C), 68–81. https://doi.org/
10.1016/j.jappgeo.2016.02.011.
References Perez-Rosales, C., 1976. Generalization of the Maxwell equation for formation resistivity
factors. J. Pet. Technol. 28 (7), 819–824. https://doi.org/10.2118/5502-PA.
Abousrafa, E.M., Somerville, J.M., Hamilton, S.A., Olden, P.W.H., Smart, B.D.G., Ford, J., Perez-Rosales, C., 1982. On the relationship between formation resistivity factor and po-
2009. Pore geometrical model for the resistivity of brine saturated rocks. J. Pet. Sci. rosity. SPE J. 22 (4), 531–536. https://doi.org/10.2118/10546-PA.
Eng. 65 (3–4), 113–122. https://doi.org/10.1016/j.petrol.2008.12.009. Pirson, S., 1983. Geologic Well Log Analysis. Gulf Publishing, Houston, TX.
Aguilera, M.S., Aguilera, R., 2003. Improved models for petrophysical analysis of dual po- Rasmus, J.C., 1983. A variable cementation exponent, M, for fractured carbonates. Log.
rosity reservoirs. Petrophysics 44 (1), 21–35. Anal. 24 (06), 13–23.
Al-Ghamdi, A., Chen, B., Behmanesh, H., Qanbari, F., Aguilera, R., 2011. An improved triple- Saner, S., Al-Harthi, A., Htay, M.T., 1996. Use of tortuosity for discriminating electro-facies
porosity model for evaluation of naturally fractured reservoirs. SPE Reserv. Eval. Eng. to interpret the electrical parameters of carbonate reservoir rocks. J. Pet. Sci. Eng. 16
14 (04), 377–384. https://doi.org/10.2118/132879-PA. (4), 237–249. https://doi.org/10.1016/S0920-4105(96)00045-9.
Archie, G.E., 1942. The electrical resistivity log as an aid in determining some reservoir Sawyer, W.K., Lowe, R.B., Pierce, C.I., 2001. Electrical and hydraulic flow properties of ap-
characteristics. Trans. AIME 146, 54), 54–62. https://doi.org/10.2118/942054-G. palachian petroleum reservoir rocks. Petrohysics 42 (2), 71–82.
Arns, C.H., 2002. The Influence of Morphology on Physical Properties of Reservoir Rocks. Schön, J.H., 2011. Physical Properties of Rocks—A Work Book. Elsevier B.V, Oxford.
The University of New South Wales. Towle, G., 1962. An analysis of the formation resistivity factor-porosity relationship of
Bauer, D., Youssef, S., Han, M., Békri, S., Rosenberg, E., Fleury, M., Vizika, O., 2011. From some assumed pore geometries. Paper presented at the SPWLA 3rd Annual Logging
computed microtomography images to resistivity index calculations of heteroge- Symposium, Houston, Texas, 17–18, May.
neous carbonates using a dual-porosity pore-network approach: influence of perco- Verwer, K., Eberli, G.P., Weger, R.J., 2011. Effect of pore structure on electrical resistivity in
lation on the electrical transport properties. Phys. Rev. E 84 (011133), 1–12. carbonates. AAPG Bull. 95 (2), 175–190. https://doi.org/10.1306/06301010047.
46 H. Wang, J. Zhang / Journal of Applied Geophysics 162 (2019) 35–46

Vogel, H.J., Roth, K., 2001. Quantitative morphology and network representation of soil Wyllie, M.R., Spangler, M.B., 1952. Application of electrical resistivity measurements to
pore structure. Adv. Water Resour. 24 (3–4), 233–242. https://doi.org/10.1016/ problem of fluid flow in porous media. AAPG Bull. 36 (2), 359–403. https://doi.org/
S0309-1708(00)00055-5. 10.1306/3D934403-16B1-11D7-8645000102C1865D.
Wang, H., 2018. An improved dual-porosity model for the electrical analysis of fractured Ziarani, A.S., Aguilera, R., 2012. Pore-throat radius and tortuosity estimation from forma-
porous media based on the pore scale method. J. Appl. Geophys. 159, 497–505. tion resistivity data for tight-gas sandstone reservoirs. J. Appl. Geophys. 83, 65–73.
https://doi.org/10.1016/j.jappgeo.2018.09.032. https://doi.org/10.1016/j.jappgeo.2012.05.008.
Winsauer, W.O., Shearin, H.M., Masson, P.H., Williams, M., 1952. Resistivity
of brine- saturated sands in relation to pore geometry. AAPG Bull. 36 (2),
253–277.

You might also like