You are on page 1of 15

Journal of Natural Gas Science and Engineering 100 (2022) 104456

Contents lists available at ScienceDirect

Journal of Natural Gas Science and Engineering


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

A critical review of breakthrough pressure for tight rocks and


relevant factors
Cheng Zhang a, Milei Wang b, *
a
School of Environmental and Municipal Engineering, Qingdao University of Technology, Qingdao, 266520, China
b
Qingdao Institute of Marine Geology, China Geological Survey, Qingdao, 266237, China

A R T I C L E I N F O A B S T R A C T

Keywords: The breakthrough pressure of tight rocks is a critical parameter in the geological storage of CO2 and the
Breakthrough pressure extraction of unconventional natural gas resources. Breakthrough pressure determines the caprock sealing effi­
Geological carbon sequestration ciency and the generation and migration of gas phases, particularly in the presence of groundwater. Laboratory
Caprock
experiments are the most representative approach to determine the breakthrough pressure given the complexity
Relevant factors
of the real stratigraphic conditions underground. The related definitions of capillary pressure, experimental
methods, and influencing factors of breakthrough pressure are comprehensively reviewed in this study.
The definitions related to capillary pressure in two-phase flow and methods to obtain breakthrough pressure
are summarized, and relevant factors and mechanisms for breakthrough pressure are discussed in detail.
Breakthrough pressure is fundamentally determined by the interfacial tension (IFT), contact angle and capillary
radius. Other parameters (e.g., temperature; pressure; salinity; ion species and concentration; pore structures;
mineral composition) indirectly affect breakthrough pressure by modifying these three parameters. This study
provides the most detailed overview of the evaluation of the breakthrough pressure and is expected to provide
theoretical support for CO2 geological sequestration and shale gas exploitation projects.

1. Introduction projects must have a capillary pressure that is sufficiently high to pre­
vent CO2 from escaping and enter the upper aquifer to guarantee safe
With the rapid development of industrialization in recent decades, storage (Rezaeyan et al., 2010).
the discharge of greenhouse gases (mainly CO2) has exacerbated climate Shale gas, an unconventional natural gas, has attracted the attention
change (De Silva et al., 2015; Lv et al., 2015; Norhasyima and Mahlia, of countries worldwide due to the shortage of conventional resources
2018). Geological carbon sequestration (GCS) has been widely recog­ such as coal, petroleum, and natural gas (Bocora, 2012; McGlade et al.,
nized as an important way to mitigate climate change worldwide in 2013; Yang et al., 2017). The significant proportion of nanopores in
recent years (Liu et al., 2019; Mikunda et al., 2021; Moore et al., 2021; shale makes it a low-porosity and low permeability medium
Omotilewa et al., 2021; Ranjith et al., 2013; Rezk et al., 2019; Sun et al., (Scotchman, 2016; Wang et al., 2016; Wang and Yu, 2019). The capil­
2021; Wen et al., 2021). Primary sequestration mechanisms include lary pressure generated from the small pore throat is considerable
stratigraphic structure sequestration, capillary residual sequestration, (Moghadam et al., 2020; Nojabaei et al., 2016; Verdugo and Doster,
dissolution sequestration, and mineralization sequestration (Amar­ 2022). When the gas pressure generated by organic matter cracking
asinghe et al., 2021; Li et al., 2017; Yekeen et al., 2021). The upward exceeds the capillary effect at the pore throat, the gas migrates along the
movement of CO2 plume is prevented when encounters the formation connected pathways to form a gas reservoir (Xu et al., 2015). During
with low permeability and high capillary pressure, and geologic struc­ shale gas exploitation, countercurrent circulation (imbibition of water
ture storage then formed (Cuéllar-Franca and Azapagic, 2015). Capillary forces out gas) is one of the critical mechanisms for gas displacement to
residual sequestration refers to CO2 injected underground entering pores achieve gas recovery enhancement (Babadagli and Hatiboglu, 2007;
with a relatively small pore size and is constrained in the rock pores (Li Ghaedi and Riazi, 2016; Moghadam et al., 2020; Verdugo and Doster,
and Elsworth, 2015) under the interfacial tension between the gas and 2022). The water saturation increases due to a large amount of water
liquid phases in the pore throat. Therefore, target formation for GCS injected into the hydraulic fracturing fluid, leading to increased

* Corresponding author.
E-mail address: wangmilei_415@163.com (M. Wang).

https://doi.org/10.1016/j.jngse.2022.104456
Received 28 October 2021; Received in revised form 18 December 2021; Accepted 30 January 2022
Available online 3 February 2022
1875-5100/© 2022 Elsevier B.V. All rights reserved.
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

capillary pressure (Verdugo and Doster, 2022). If the pressure difference


2σ cos θ
between gas and liquid in a pore does not reach the capillary pressure of Pc = Pw − Pg = (1)
r
the pore throat, the water-lock effect will form (An, 2019), preventing
gas from entering the connected pore channel (Zhang et al., 2011). The where σ is the interfacial tension (IFT) between the non-wetting phase
water lock effect will significantly reduce the permeability of shale gas (gas) and the wetting phase (brine); r is the equivalent radius of the pore
and seriously affect the productivity of shale gas, which is also one of the throat; and θ is the contact angle (CA) of the gas-liquid-solid system.
major problems of exploiting low permeability reservoirs (Liang et al.,
2021). Breakthrough pressure is also a significant factor in evaluating
2.1. Entry pressure (Pen)
the sealing efficiency of shale caprocks (Zhou et al., 2019).
The breakthrough pressure of low permeability rocks determines the
From an experimental perspective, the applied gas pressure must
sealing performance of geological storage of CO2 and nuclear waste and
reach a certain value to enter the pore network. The entry pressure refers
affects the migration of shale gas (Rezaeyan et al., 2015; Wu et al., 2020;
to the gas pressure to enter the largest connected pore and displace the
Xu et al., 2018). Therefore, gas breakthrough pressure has attracted
water (Guiltinan et al., 2018). From a macroscopic experimental
extensive attention in GCS, shale gas exploration, and underground
perspective, the entry pressure is the maximum pressure difference that
storage of nuclear waste projects (Xu et al., 2018). Many scholars have
exists across the gas-liquid interface before the non-wetting fluid begins
investigated breakthrough pressure; thus, it is important to review
to penetrate the connected pore space (Minardi et al., 2021). At this
existing research in terms of the definition of breakthrough pressure,
stage, however, due to local heterogeneities or the scale of observation,
experimental methods, and influencing factors.
the network invaded by the gas may not reach the other side of the
In this review article, we first review the definitions of pressure
sample, preventing breakthrough (Boulin et al., 2013). Thus, the entry
derived from capillary pressure in detail. Then, the different experi­
pressure is the pressure of a non-wetting fluid (Pg) that is required to
mental methods are described, and the advantages and limitations are
start to displace the wetting phase within the rock (Cui et al., 2019;
discussed. Finally, the effects of rock and fluid properties on break­
Guiltinan et al., 2018; Rezaeyan et al., 2015) (Fig. 2, a), and the satu­
through pressure are summarized and analyzed. This review can thus
ration of wetting fluid begins to decrease (Ferrand and Celia, 1992).
serve as a guide and inspiration for future research on breakthrough
pressure.
2.2. Threshold pressure (Pc)
2. Definitions of capillary pressures
In a multiphase system, the gas phase must overcome a certain
Many terminologies are currently used to describe capillary pressure capillary threshold pressure to displace the wetting phase from the pore
in a gas-liquid-solid system (Iglauer et al., 2015). The most basic pres­ system (Amann-Hildenbrand et al., 2012). The threshold pressure refers
sure is capillary pressure, which is the resistance of the non-wetting to the gas pressure at which water is first expelled from the most con­
phase to penetrate the pore system within a porous medium in the nected pores of a water-saturated sample (Boulin et al., 2013) and equals
presence of a wetting fluid (Andrew et al., 2014; Rezaeyan et al., 2015) the gas pressure at which a continuous water flow comes out of the rock;
(Fig. 1). Capillary pressure is the pressure difference between the wet­ however, there are no continuous gas flow forms (Thomas et al., 1968)
ting phase and the non-wetting phase, which exists as a result of surface (Fig. 2, b). The threshold pressure is a critical parameter when evalu­
forces between the two phases (Smith, 1966), and the capillary pressure ating the ability of a caprock to contain gas. Experimentally, the pres­
for a pore throat can be expressed as (Laplace, 1806; Washburn, 1921a, sure must be increased for gas breaking through the rock to measure the
1921b; Young, 1805): breakthrough pressure.

2.3. Breakthrough pressure (Pb)

Breakthrough pressure is most commonly used in both evaluations of


GCS efficiency and reservoir gas sealing efficiency, and is defined as the
minimum gas pressure required to overcome the capillary pressure in
the largest pore channels (Aggelopoulos et al., 2010; Heath et al., 2012;
Li et al., 2005; Schowalter, 1979; Song and Zhang, 2012; Zheng and
Espinoza, 2021). Thus, the breakthrough pressure is reached when the
percolation threshold is exceeded, and the continuous gas flow can be
formed and arrive at the outlet of the rock core through connected paths
across the system of interest (Heath et al., 2012; Song et al., 2016)
(Fig. 2, c). The breakthrough pressure is a macroscopic parameter that
can be used to evaluate the sealing efficiency of core plugs or entire seal
units (Heath et al., 2012).
Breakthrough pressure corresponds to the capillary pressure of the
most prominent interconnected pores in the rock sample (Boulin et al.,
2013). From an experimental perspective, breakthrough pressure refers
to the pressure difference between the inlet and outlet when the gas flow
is formed in a water-saturated sample (Zhou et al., 2019).

2.4. Snap-off pressure (Psnp)

After gas breakthrough, the gas pressure decreases and leads to water
moving from the downstream of the flow path upstream and stopping in
the narrowest pores. This process is referred to as water imbibition and
Fig. 1. Diagram of Pc in a pore throat with three solid-liquid-gas phases continues until all of the interconnected pathways are blocked (Fig. 2,
(Iglauer et al., 2015; Li et al., 2005; Schowalter, 1979; Zhang and Yu, 2019). d). During water imbibition, the relative permeability of the gas phase

2
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Fig. 2. Stages of gas breakthrough and reimbibition in water-saturated rock samples: (a) Pg = Pen, (b) Pg = Pc, (c) Pg = Pb, (d) Pg = Psnp.

will decrease due to the reduction of the effective diameter of pore


throats (Hildenbrand et al., 2002). Ultimately, a remaining pressure
difference that is lower than the breakthrough pressure between the
inlet and the outlet of the rock sample is defined as the snap-off pressure
(Busch and Amann-Hildenbrand, 2013), which is also called the shut-off
pressure or the residual pressure (Egermann et al., 2006). After this
point, gas flow is blocked by the water in the pore throats, and only
diffusional transport of gas occurs (Wollenweber et al., 2010). In gen­
eral, the snap-off pressure is 20%–50% of the breakthrough pressure
(Song and Zhang, 2012; Zweigel et al., 2005).

3. Experimental methods

Breakthrough pressure can be affected by many factors (e.g., tem­


perature, pressure, fluid composition) (Amann-Hildenbrand et al., 2013;
Gao et al., 2015). The breakthrough pressure can be determined by
direct and indirect methods at the core scale (Minardi et al., 2021;
Vilarrasa and Makhnenko, 2017). The direct method primarily includes
the step-by-step method (SBS), continuous injection method, dynamic
method, and residual capillary pressure method (RCP) (Busch and
Müller, 2011; Ito et al., 2011). The indirect method is the mercury in­
jection porosimetry (MIP) method. Fig. 3. Pressures upstream and downstream during the breakthrough process.

the upstream chamber will increase linearly until the gas flows out of the
3.1. SBS method
rock sample. However, the upstream pressure will fluctuate in a small
range during this process due to the heterogeneity of pores (Egermann
The SBS method was first proposed by Thomas et al. (1968). et al., 2006). The pressure difference between the upstream and
Although it is time-consuming for low permeability materials (Busch downstream when gas flow from the core is observed is recognized as
and Müller, 2011), the SBS method is the most representative approach the breakthrough pressure. If the rate of pressure increase is too fast, the
of in situ hydrocarbon migration through caprocks (Boulin et al., 2013;
gas breakthrough time is insufficient, resulting in an overestimation of
Zhang and Yu, 2019). The principle of this approach relies on the defi­ the breakthrough pressure.
nition of capillary pressure and thus is also recognized as the standard
method (Egermann et al., 2006). Before the breakthrough experiment,
3.3. Dynamic method
the rock sample is prefabricated to a certain water saturation according
to the research needs. The injection pressure of the non-wetting phase at
Before the experiment, the inlet part of the system is filled with brine
the inlet is imposed and maintained by a stepwise pressure increase. The
and then gas is introduced to promote the water into the rock sample
pressure difference between the upstream and downstream when gas
(Egermann et al., 2006). The pressure of the injected gas is at a constant
flow is formed in the outlet is recognized as the breakthrough pressure.
△Pt higher than the expected threshold pressure. The fluid (wetting and
Thus, the accuracy of this method depends on the pressure step to some
non-wetting) rate can be obtained from the separator at the outlet of the
extent (Ito et al., 2011), and the measured breakthrough pressure is
experimental apparatus. During the first period, only the wetting fluid
often higher than the real value (Tanai et al., 1997). Preliminary
flows in the sample under the applied gas pressure (△Pt), and the gas is
experimental data show that the pressure changes throughout the
still located in the inlet chamber of the system. The relationship between
breakthrough process (Fig. 3).
the brine flow rate (Qw, abs) and △Pt can be described according to
Darcy’s law (Equation (2)) (Boulin et al., 2013). Then, as soon as the
3.2. Continuous injection method non-wetting fluid reaches the inlet face of the core, the effective pressure
difference caused by the threshold pressure to displace the brine will
The upstream gas in the water-saturated rock sample is injected at a decrease. This pressure difference (△Pw) can be described by Equation
slow constant flow rate (Rudd and Pandey, 1973), and the pressure in (3). There exists both wetting and non-wetting phase flow in the core,

3
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

and the brine flow rate drops suddenly when the single-phase flow
transforms to the two-phase flow (Fig. 4). The threshold pressure (Pc)
can be calculated by Equation (4):
ΔPt kA
Qw,abs = (2)
μw L

ΔPw kA
Qw,eff = (3)
μw L

Pc = ΔPt − ΔPw (4)

where Qw, abs and Qw, eff are the flow rates of the brine in the single-phase
and two-phase flow processes, respectively; μw is the dynamic viscosity
of brine; L and A are the length and cross-sectional area of the core
sample, respectively; and k is the intrinsic permeability (absolute
permeability).
The breakthrough time varies from weeks to months with the SBS
and continuous injection methods, as shown in many previous break­
through studies (Boulin et al., 2013; Li et al., 2005). Egermann et al.
(2006) showed that the dynamic method is as accurate as the SBS
method and as fast as the RCP method under reservoir conditions, thus it
Fig. 5. Pressure variation of the RCP method.
is often adopted.

media (Mayer and Stowe, 1965). Mercury is forced into the pores of the
3.4. RCP method evacuated solid under pressure, and the variation curve of injected
mercury saturation with pressure is recorded in Fig. 6. An entire capil­
The RCP method, which is also known as the impulse method (Gao lary pressure curve consisting of more than 20 points can be determined
et al., 2015), was first proposed by Hildenbrand et al. (2002). The in a matter of hours rather than weeks; thus, this method is convenient
principle of the RCP method is to apply a gas pressure that is well above and quick. Another advantage of this method is that it can be conducted
the anticipated capillary pressure across the rock and monitor the using small and irregularly shaped pieces of rock instead of core plugs
change in differential pressure between upstream and downstream to (Amann-Hildenbrand et al., 2013; Busch and Amann-Hildenbrand,
obtain the breakthrough pressure (Fig. 5). The entire pressure change 2013; Purcell, 1949; Zheng and Espinoza, 2021). Fig. 6a exhibits the
process for the upstream and downstream is described in Fig. 5. Eger­ mercury intrusion volume change with the injection pressure. The
mann et al. (2006) found excellent consistency of the RCP and dynamic pressure corresponds to the intersection of the tangent to the sigmoidal
method results; however, the RCP method leads to a systematic under­ curve at the inflection point and the logarithmic pressure axis is the
estimation of the real threshold capillary pressure value (Ito et al., 2011; breakthrough pressure (displacement pressure) (Schlmer and Krooss,
Kawaura et al., 2013). This approach measures the snap-off pressure, 1997). However, taking the tangent line may lead to large errors from
which is the minimum capillary displacement pressure. the injection curve (Wu et al., 2020), the injection pressure at 10% Hg
saturation is always considered an estimate for the displacement pres­
3.5. Mercury injection porosimetry (MIP) sure (Fig. 6b) (Amann-Hildenbrand et al., 2013; Busch and
Amann-Hildenbrand, 2013; Schowalter, 1979). This pressure should be
The mercury injection porosimetry method was first proposed by transferred to the displacement pressure of the water-gas system by
Washburn (1921b) for measuring the pore size distribution in porous Equation (5) (Hildenbrand et al., 2002; Krevor et al., 2012; Purcell,
1949; Schowalter, 1979):
γgw ⋅cos θgw
Pcgw = ⋅Pcma (5)
γma ⋅cos θma

where Pcgw , γ gw , θgw are the capillary pressure, interfacial tension, and
contact angle for the gas-water system, respectively; and Pcma , γ ma , θma
are the capillary pressure, interfacial tension, and contact angle for the
air-mercury system, respectively. Previous research has shown that γma
= 485 mN/m and θma = 140◦ (Vavra et al., 1992).
Each of these methods has advantages and limitations regarding
accuracy, time consumption, in situ representativeness, and fluid
properties (Table 1).

4. Factors relevant to breakthrough pressure

Breakthrough pressure can be influenced by the gas-liquid systems’


interface properties, rock media properties, liquid saturation, non­
wetting phase type, and other factors.

4.1. Interfacial properties


Fig. 4. Variation in wetting fluid flow volume in the dynamic experiments
(modified from Egermann et al. (2006)). The capillary pressure in a pore throat is closely related to the gas-

4
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Fig. 6. Capillary pressure curve from the MIP experiment. (a) Intrusion volume versus log (Pressure) (modified from (Schlmer and Krooss, 1997)); (b) Pressure versus
intrusion volume fraction.

Broseta, 2012; Feng et al., 2021). The IFT is dependent on many factors,
Table 1
including fluid type, brine salinity, and external temperature and pres­
Characteristics of the different methods of measuring breakthrough pressure
sure. In this study, we used a series of gas/brine systems as the reference
(Busch and Müller, 2011).
case considering both the capillary sealing efficiency for geological CO2
Method Duration Rock Accuracy Comments storage and shale gas extraction. A summary of some typical studies of
morphology
the dependence of pressure, temperature, salinity, and gas mixtures on
SBS Long Bulk/Core High Standard method IFT is shown in Table 2.
plug
Continuous Short Bulk/Core Medium/ Overestimation
injection plug Bad 4.1.1.1. Effects of pressure and temperature on IFT. From Table 2, a
Dynamic Short Bulk/Core Good Overestimation consensus conclusion has arrived that the IFT of the gas-brine system
plug decreases strongly with increasing pressure until reaching an asymptotic
RCP Short Bulk/Core Low Underestimation
value (plateau) at high pressures regardless of which gas or brine was
plug
MIP Short Cuttings Medium Data conversion used (Arif et al., 2016a; Bachu and Bennion, 2009; Chiquet et al., 2007;
error Espinoza and Santamarina, 2010; Liu et al., 2017; Sarmadivaleh et al.,
2015). The limiting pressure beyond which the IFT ceases to decrease is
defined as the stabilization pressure (Bikkina et al., 2011; Liu et al.,
liquid-solid contact angle and IFT between gas and brine according to 2017), which shows a positive correlation with temperature (Aggelo­
Equation (1) (Espinoza et al., 2011). These two critical interfacial pa­ poulos et al., 2010; Chiquet et al., 2007). At high pressures, where the
rameters determine fluid permeability and distribution (Bachu and IFT has reached its asymptotic value, the IFT does not depend on the
Stefan, 2015; Chalbaud et al., 2009), which are critical to capillary pressure or temperature.
sealing in CO2 geological storage and the enhancement of unconven­ The pressure increase results in higher solubility of the gas phase (Liu
tional natural gas recovery (Aggelopoulos et al., 2010; Song and Zhang, et al., 2017) and higher concentration of aqueous species (Espinoza and
2012; Yu et al., 2020). The maximum gas height (CO2 or hydrocarbon) Santamarina, 2010; Yekeen et al., 2021). IFT decreases with pressure
that can be stored under a certain caprock is calculated by the following due to the reduction in surface energy caused by the increase in solu­
equation (Aggelopoulos et al., 2010, 2011; Chiquet et al., 2007, 2010; bility of the gas in the aqueous phase (Aggelopoulos et al., 2010; Yekeen
Song and Zhang, 2012): et al., 2021). Due to compressibility and low density of gas phase, the
2σ cos θ density increases with increasing pressure, which leads to a lower den­
H =( ) (6) sity difference between the gas and liquid phases. Thus, IFT will
ρw − ρg gr
decrease with the decreasing density difference between the gas and
where g is gravity, and ρw and ρg are the densities of water and gas, brine systems (Song and Zhang, 2012).
respectively. Therefore, gas/water IFT is critical to the storage of gas Table 2 highlights that temperature has a positive effect on IFT for
under a caprock. different gas liquid systems. Previous experiments generally selected
only three or fewer temperature points as variables (Aggelopoulos et al.,
4.1.1. Brine-gas interfacial tension 2011), which does not give a good indication of the temperature
IFT plays a vital role in the mechanism of multiphase flow in porous dependence of IFT. Since comprehensive studies on the temperature
media (Nojabaei et al., 2016; Rashid et al., 2017; Rezaeyan et al., 2010). dependence of IFT are focused on CO2 water systems, the following
Brine-gas IFT is generated because the liquid molecules in the interface discussion is mainly focused on CO2 water systems. In a systematic study
are differently attracted by the liquid molecules with the liquid bulk and of IFT for CO2 brine systems under 11 different temperatures (300–353
the surface gas (Khosharay, 2015; Lin et al., 2020). From Equation (1), K) and 10 different pressures (3–12 MPa), Liu et al. (2017) found that
the capillary pressure can be increased by increasing the IFT, and the IFT increases with temperature. This increasing trend becomes more
sealing efficiency of the caprock consequently improves (Duchateau and pronounced with increasing pressure until the critical pressure is

5
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Table 2
Summary of the studies on IFT in the brine-gas system.
Authors Gas Aqueous phase Salinities Temperature Pressure IFT variation

Chalbaud et al. CO2 NaCl solution 5,100, 200 and 27, 71, and 100 45–255 bars IFT increased with increasing salt concentration and
(2009) 300 g/L ◦
C temperature.
IFT decreased with pressure and then stabilized.
Chiquet et al. CO2 Water and NaCl 20 g/L 308–383 K 5–45 MPa IFT decreases with pressure below 25 MPa and then
(2007) solution stabilized; the effects of both salinity and temperature are
negligible.
Sarmadivaleh CO2 Water 0 296, 323 and 343 0–20 MPa IFT increased with temperature.
et al. (2015) K IFT decreased with pressure up to a pressure of ~10 MPa.
Saraji et al. (2014) CO2 Brine 0.2–5 M 50–100 ◦ C 2000- The effect of pressure is not significant.
4000Psi IFT decreases marginally with temperature.
Arif et al. (2016a) CO2 NaCl solution 20 wt% 308, 323, 343K 0.1–20 MPa IFT decreased with pressure.
IFT increased with temperature.
Li et al. (2012) CO2 NaCl + KCl 1.98 mol/kg 298–473K 0–50 MPa IFT decreased with pressure.
solution IFT increased linearly with molar concentration.
Fukuzawa et al. CO2–H2 Water 0 271–280 K 0–7 MPa IFT decreased with pressure.
(2018) mixtures The temperature dependency of the IFT could not be
observed.
Yekeen et al. CO2 NaCl solution 0, 3, 5, 7 wt% 80–180 ◦ C 8–22 MPa IFT increased with temperature and salinity but decreased
(2021) with pressure.
Aggelopoulos et al. CO2 NaCl + CaCl2 0.045–1.5 m for 27, 71, 100 ◦ C 50–250 bar IFT decreased with pressure, and the descent gradually
(2011) mixed solution each salt slowed to vanished.
IFT increased linearly with molar concentration.
IFT increased with temperature between 27 and 71 ◦ C and
remained constant thereafter.
Yan et al. (2001) CH4+N2 Water 0 298–373 K 1–30 MPa For the different gas-water systems, IFT decreased with
CO2+N2 pressure and increased with N2 mole fractions.
The temperature dependence of IFT is rather complex.
Liu et al. (2016) CO2+CH4 NaCl solution 0–200000 ppm 77 F–257 F 15-5027 pisa IFT decreased with CO2 mole fraction but increased with
mixtures salinity.
IFT decreased with pressure until a plateau was reached.
Bachu and CO2 Brine 0–334000 mg/L 20–125 ◦ C 2–27 MPa IFT steeply decreases with pressure and then toward a
Bennion (2009) plateau value for high pressures.
IFT increases with increasing water salinity.
IFT tends to rise, then fall, and then rise with temperature.
Aggelopoulos et al. CO2 CaCl2 solution 0.0045–2.7 m/L 27, 71, 100 ◦ C 50–250 bar IFT decreased with pressure until it reached the stabilized
(2010) pressure (Pst), depending on the temperature.
IFT increased with temperature from 27 to 71 ◦ C. No
increase of IFT was found for temperature from 71 to 100

C.
IFT increases linearly with salinity, and the increase is
proportional to the cation valence.
Liu et al. (2017) CO2 NaCl solution 0–1.8 m/kg 300–353K 3–12 MPa IFT decreased with pressure until critical pressure was
reached and then became constant. IFT increased with
temperature and salinity.
Mutailipu et al. CO2 NaCl + KCl 1–4.9 mol/kg 298–373K 3–15 MPa IFT increases with salinity and temperature and decreases
(2019) solutions with pressure until reaching a plateau.
Al-Yaseri et al. CO2、N2、 NaCl 5000 ppm 333K 13 MPa IFT of the N2-brine system is higher than that of CO2-brine
(2015) N2+CO2 and (N2+CO2)-brine systems.
Lun et al. (2012) CO2 Reservoir brine 0、14224.2、 45 ◦ C、97.53 ◦ C 0.1–36 MPa IFT decreased with pressure (~7.38 MPa) first and then
21460.6 mg/L unordered, and decreased (15Mpã) toward a plateau.
Bikkina et al. CO2 Water 0 298.15–333.15K 1.48–20.76 IFT decreases with pressure and then becomes stabilized.
(2011) MPa
Al-Anssari et al. CO2 NaCl, KCl, CaCl2 0-5 wt% 296–343K 0–20 MPa IFT increased with temperature and salinity.
(2018) IFT strongly decreased with pressure before reaching
nearly a pseudo plateau at approximately 12 MPa
Liu et al. (2016) CH4+CO2 NaCl ~200000 ppm 77-257 ◦ F 15-5027 psi The presence of CO2 decreases the IFT of the CH4/brine
system.
A higher salinity leads to an increased IFT for a given
system.
IFT decreases with pressure and then stabilizes at a specific
pressure.

reached, due to the conversion of CO2 from gas to the supercritical range (Al-Anssari et al., 2018; Mutailipu et al., 2019) (Fig. 7a). The
phase. Yekeen et al. (2021) also reached a consistent conclusion IFT decreases with the temperature at high temperature (T > 343 K) and
experimentally with a wide temperature variation (353–453 K) under low pressure (<5 MPa), particularly when the pressure is lower than 3
10 MPa. The increase in IFT with temperature is due to the decrease in MPa, and the temperature is higher than 373 K. Also, research from
the solubility of CO2 in the aqueous phase (Liu et al., 2017). Chiquet et al. (2007) showed that the IFT fluctuated with temperature
However, other researchers have found different results by under pressure below 15 MPa and gradually declined with temperature
measuring the IFT variation of the CO2water system under different when the pressure exceeded 20 MPa (Fig. 7b). In addition, some studies
temperature and pressure conditions. The IFT was found to increase have shown that the water gas (mixture of CH4 and CO2) IFT is inde­
with temperature in the low-temperature and high-pressure range and pendent of temperature (Ren et al., 2000). Thus, the literature research
decrease with temperature in the high-temperature and low-pressure indicates that the dependence of IFT on temperature is complex

6
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Fig. 7. CO2 water IFT variations for temperature under different pressures. (a) Results from Al-Anssari et al. (2018); (b) results from (Chiquet et al., 2007).

(Aggelopoulos et al., 2011; Li et al., 2012), which can be attributed to


the complications of the CO2 phase change and dissolution effects into
the brine affected by temperature and pressure (Mutailipu et al., 2019).
At higher temperatures (e.g., above 333 K), Bachu and Bennion
(2009) reported that the combined effect of temperature and pressure
dependence of IFT for CO2 can be expressed by the following equation
without considering the salinity (Busch and Amann-Hildenbrand, 2013):
( )
σ = 40.98 − 1.582 ⋅ P + 0.032 ⋅ P2 + 0.100 T − Tref (7)

where P and T are pressure and temperature, respectively; and Tref is the
reference temperature (348 K in this equation).

4.1.1.2. Effects of salt species and concentration on IFT. The valence of


cations strongly affects the interfacial tension (Aggelopoulos et al.,
2011). Salts improve the ionic strength gradient around the gas-aqueous
interface (Al-Anssari et al., 2018). The IFT of the CaCl2 solution CO2
system is approximately twice that of the NaCl CO2 system for all
pressures and temperatures at a given salinity (0.33, 1 and 2 mol/L)
(Chen et al., 2017) due to the stronger repulsion of the divalent cations
(e.g., Ca2+) from the interface; they have a higher hydration than the
monovalent cation Na+ (Manciu and Ruckenstein, 2003; Peter and
Robert, 1996). Similar relationships exist for other chlorides under Fig. 8. Comparison of the interfacial tension increase of CO2/NaCl solution,
ambient conditions, and the ranking of the influence of cations (at the CO2/CaCl2 solution, and CO2/(NaCl + CaCl2) solution as a function of each
same concentration conditions) on the increase in IFT is as follows salt’s molality (Aggelopoulos et al., 2011).
(Chalbaud et al., 2009; Li et al., 2017; Ralston and Healy, 1973):
water and gas, respectively; m is the molality of the single solute; and λ is
Cs+ < Rb+ < NH+ + + +
4 < K < Na < Li < Ca
2+
< Mg2+ (8)
the slope. Another empirical equation for evaluating the IFT for a gas/
The evolution of CO2/brine IFT is reported to increase linearly with mixed solution of different valence cation systems is as follows:
salinity and with different slopes using different salts (Aggelopoulos
σ = A[m+ ] + B (10)
et al., 2010) (Fig. 8). The increased salinity causes the specific gravity of
water to increase and, consequently, the density difference between gas ∑
n
and liquid to become greater, leading to a higher IFT for the gas/brine m+ = zi m i (11)
system (Liu et al., 2016). This increase can also be attributed to the 1

depletion of the solute from the interface (Liu et al., 2017). Bachu and
where mi and zi are the molalities and charge number of the ith cations,
Bennion (2009) attributed the positive dependence of IFT on salinity (5
respectively; and A and B are functions of temperature and pressure,
different concentrations from 0 to 334010 mg/L) to the increasing
which are polynomial relationships and vary marginally with different
aqueous electrolyte and decreasing CO2 solubility. The relationship
studies (Li et al., 2012; Mutailipu et al., 2019).
between the IFT and salinity can be described as:
σ = λm + σ 0 (9) 4.1.2. Wettability (contact angel)
In Equation (1), θ is the contact angle (CA), which is determined by
where σ and σ0 are IFT between brine and gas, and IFT between pure the wettability of a water-gas-rock system (Chalbaud et al., 2009).

7
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Wettability refers to the tendency of the fluid to spread on a surface in 4.2. Inherent properties of reservoir formation
the presence of other immiscible fluids and is caused by intermolecular
forces (Iglauer et al., 2015). In a rock-brine-gas system, the brine will 4.2.1. Formation properties
occupy the smallest pores and spread on the pore walls as water films According to the hydrostatic equation (Equation (13)), the hydro­
under the effect of water wettability. The CA is ultimately formed under static pressure of the formation underground increases with increasing
the balance of interfacial tension between brine-gas, brine-solid, and burial depth (Fig. 11). This additional pressure leads to an increase in
solid-gas phases (Fig. 9). The CA can be described by Young’s equation CO2 breakthrough pressure (Li et al., 2011), and the breakthrough
(Young, 1805): pressure of CO2 can be calculated by Equation (14) (Heath et al., 2012).
γ sg − γ sl However, few studies have shown that the variation in breakthrough
cosθ = (12) pressure with depth is sometimes controlled by diagenetic evolution.
γlg
The diagenesis process can also influence breakthrough pressure by
where γsg , γsl , γlg are the IFTs of the solid-gas (rock/mineral-CO2), solid- affecting the porosity and permeability. Most primary pores have been
brine (rock/mineral-brine), and brine-gas (brine-CO2) interfaces, lost by mechanical compaction or mineral cementation, and the reduc­
respectively. tion of porosity by mechanical compaction is more significant than that
According to Equation (1), the capillary pressure can be increased by by cementation (Aplin et al., 1995; Lai et al., 2017). Compaction and
reducing the CA of a rock-gas-brine system; thus, a higher CA leads to a cementation will decrease the porosity and permeability (Duan et al.,
lower capillary pressure (Sarmadivaleh et al., 2015; Yekeen et al., 2018; Fu et al., 2016). The different nature of the minerals generated in
2021). The wettabilities of different brine-solid-gas systems were the clay mineral transformation and the primary minerals also con­
analyzed by many specialists from different research areas (Table 3). tributes to the pore structure variations (Henning and Dypvik, 1983).
All studies in the literature, without exception, concluded that the Porosity of shale increases with increasing maturity of organic matter
CA for a CO2-brine-rock system increases with increasing pressure due to the thermal degradation process (Bernard et al., 2012; Eseme
(Al-Yaseri et al., 2015; Chiquet et al., 2010; Iglauer, 2017; Pan et al., et al., 2012; Hu, 2013).
2018; Sarmadivaleh et al., 2015; Yekeen et al., 2021). This result can be Psaline = ρgh (13)
explained by the increased intermolecular interactions between CO2 and
rock substrates (Al-Yaseri et al., 2015). However, some researchers PCO2 = Pc + Psaline (14)
attributed the CA increment to the influence of pressure on fluid density,
as well as fluid-fluid and solid-liquid interfacial tension (Sarmadivaleh where h is the burial depth of the formation, and Psaline is the hydrostatic
et al., 2015; Yekeen et al., 2021). pressure of the overlying formation brine.
A similar agreement is that higher brine salinity leads to higher CA Caprock thickness is an essential parameter in caprock sealing effi­
(Chiquet et al., 2010; Iglauer, 2017; Pan et al., 2018; Sarmadivaleh ciency evaluation and has been researched by scholars for years (Ma
et al., 2015; Yekeen et al., 2021), and the CAs for divalent cations are et al., 2019). The thick formations are not easily to generate
higher than those for monovalent cations (Al-Yaseri, 2016; Arif et al., cross-cutting faults and thus have better integrity (Li et al., 2008). When
2017a; Pan et al., 2018; Roshan et al., 2016). Zeta potential, one of the caprock acts as a permeable barrier, thick caprocks achieve better
important electrokinetic properties of minerals, is defined as the elec­ sealing performance than thin caprocks (Li et al., 2008). Besides, thicker
trical potential developed at the solid-liquid interface in response to the caprock constitutes a longer distance of gas bubble migration in the
relative movement of solid and liquid (Bartlett and Sparks, 1986; Kaya caprock (Zhang et al., 2012b), as a consequence, thicker formations
and Yukselen, 2005). Less negative zeta potentials lead to less polar could have higher gas storage capacity (Raziperchikolaee et al., 2019).
surfaces and thus de-wetting of the surface (Adamczyk et al., 2010). As a When the pressure reaches the mechanical strength of the caprock,
result, the increase in contact angle with salinity is due to the less thickness is essential in ensuring the integrity of a hydraulic seal. (Song
negative zeta potential values at higher salinities (Arif et al., 2016a, and Zhang, 2012). Formations with connected, open fractures generally
2017a, 2017b; Sarmadivaleh et al., 2015). have lower breakthrough pressures, and CO2 or hydrocarbons will
With regard to temperature, its effect on CA is unclear due to a lack of migrate up the fractures rather than accumulate in the formation (Smith,
systematic investigations (Iglauer, 2017; Sarmadivaleh et al., 2015; 1966). Heterogeneity is an important characteristic of caprocks because
Yekeen et al., 2021). Temperature influences CA by affecting the den­ of depositional heterogeneities and geomechanical properties during the
sity, van der Waals forces (solid-CO2), and hydrogen bonds (e.g., on the sedimentary process (Gelhar, 1986), which is always represented by
water-wet surface); however, van der Waals forces decrease, and permeability heterogeneity, porosity heterogeneity and capillary pres­
hydrogen bonds are increasingly broken on water-wet surfaces with sure heterogeneity (Krevor et al., 2015). In addition, the distributions of
increasing temperature (Pan et al., 2018). impermeable clay lenses also impose the rock heterogeneity (Hesse and
Woods, 2010). Thus, nearly all fluid transport in a rock matrix is het­
erogeneous (Ferrand and Celia, 1992), and the flow path of least resis­
tance for CO2 leaks is closely related to the heterogeneity of formation
(Cavanagh and Haszeldine, 2014; Meckel et al., 2015).

4.2.2. Pore structure characteristics


Typically, the capillary pressure in a water-saturated pore throat is
determined by the pore throat radius, which is related to the pore
structure (Espinoza and Santamarina, 2010; Zhang and Yu, 2016, 2019;
Zhao and Yu, 2017, 2018; Zheng and Espinoza, 2021). Thus, pore
structure characteristics play an important role in the gas breakthrough
pressure in a gas-liquid-solid system. According to the Washburn
equation (Equation (1)), the breakthrough pressure decreases with
macropore (diameter≥50 nm) proportions and increases with micropore
(diameter≤2 nm) and mesopore (2 nm<diameter<50 nm) volumes
(Zhang and Yu, 2016; Zhou et al., 2019); thus, pore size influences the
magnitudes of breakthrough pressure (Heath et al., 2012). Pore size
Fig. 9. CA in a gas-rock-liquid system.

8
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Table 3
Summary of the typical studies on contact angle in a rock-gas-brine system.
Reference Solid Gas Brine Salinities Temperature Pressure Conclusion

Yekeen et al. Shale CO2 NaCl 0~7 wt% 80–180 C ◦


10 MPa CA increase with temperature, pressure, and salt
(2021) solutions concentration.
Chiquet et al. Mica, quartz CO2 NaCl 0.0.1–1 M ambient 1–11 MPa CA increase with pressure; the increment for
(2010) solutions mica–CO2–brine system is more significant than that of
quartz–CO2–brine system.
An increase in NaCl concentration increases the CA
Mutailipu Quartz, CO2 NaCl + KCl 1.98 mol/kg 298–373K 3–15 MPa The rock samples became less water-wet when the CO2
et al. (2019) sandstone, solutions phase state changed from subcritical to supercritical.
limestone
Al-Yaseri et al. Quartz CO2、 NaCl 5000 ppm 333K 13 MPa The CA is ordered CO2> N2> N2+CO2.
(2015) N2、 CA increase with pressure.
N2+CO2
Pan et al. Shale CO2、 NaCl、 0 wt%、1 wt 298–343K 0.1–30 The CA increased with pressure and salinity but decreased
(2018) CH4 CaCl2、 %、5 wt% MPa with temperature. The CAs were larger for divalent ions
MgCl2、 than for monovalent ions.
Gholami et al. Shale CO2 NaCl、 0–10000 ppm 50 ◦ C 15 MPa The CA showed a trend of increasing and then decreasing
(2021) CaCl2、 with salinity.
Arif et al. Mica CO2 MgCl2、 20 wt% 323K 0–20 MPa The CA increases with salinity.
(2017a) CaCl2
Stevar et al. Calcite CO2 NaHCO3 1 mol/kg 298–373K 0–30 MPa The CA increases with pressure until the pressure reaches
(2019) 10 MPa and begins to decrease and eventually stabilizes.
The CA decreases with temperature between 298 and
373K.
Roshan et al. Shale Air NaCl, KCl, 0.1、0.5、1 35 ◦ C、70 ◦ C 0.1、10、 The CA for bivalent ions is larger than for monovalent ions,
(2016) MgCl2, CaCl2 M 20 MPa and it increases with both increase of pressure and
temperature.
Yong et al. Shale CO2+CH4 Water – 300–400K 0–80 MPa The CA increases with the CO2 mole fraction.
(2021)
Arif et al. Quartz, mica, CO2 Water – 343K 0–20 MPa The CA increases with pressure.
(2016b) coal

1998). Small pores with a high specific surface in clay-rich caprocks give
rise to a high capillary entry pressure and high viscous drag that hinder
CO2 migration (Espinoza and Santamarina, 2017; Guiltinan et al., 2018;
Zheng and Espinoza, 2021). There exists an inverse relationship be­
tween breakthrough pressure and pore size (Espinoza and Santamarina,
2017), and the breakthrough pressure increases as the sediment specific
surface increases and the porosity decreases.

4.2.3. Mineral composition


Once CO2 is injected into a formation, the water in the formation
generally becomes more acidic (Naderi and Simjoo, 2018) and increases
the ability of groundwater to dissolve carbonate minerals (Equation
(15)) (Foroutan et al., 2021; Lu et al., 2012; Luquot et al., 2014; Oliveira
de Moura e Silva et al., 2021), which increases the porosity of the rock
(Rosenbauer et al., 2005; Yoksoulian et al., 2013). The dissolution of the
carbonate minerals caused by the dissolution of CO2 change the me­
chanical and physical properties of the rock to provide more channels
for CO2 transport (Busch et al., 2008; Espinoza and Santamarina, 2017;
Luquot et al., 2014; Wigley et al., 2017), and thus the capillary pressure
decreases (Lü et al., 2014).

Fig. 10. Advancing contact angle variation of brine (NaCl solution)-rock-CO2


CO2 + H2 O ↔ H2 CO3 ↔ H + + HCO3 (15a)
systems at constant pressure (10 MPa) (Yekeen et al., 2021).
CaCO3 + CO2 + H2 O = Ca2+ + 2HCO−3 (15b)
distribution plays a key role in the phase behavior of fluid and gas
CaMg(CO3 )2 + 2CO2 + 2H2 O = Ca2+ + Mg2+ + 4HCO−3 (15c)
transport (Aplin and Macquaker, 2011; Lyu et al., 2021). Continuous gas
flow tends to be formed along with connected pores that are larger than The mineral composition has an important influence on the geo­
the mean pore size (Espinoza and Santamarina, 2010). Compared to mechanical properties of fined-grained lithologies (Li et al., 2021; Sun
micropores, gas migration tends to occur in larger microfractures (Vol­ et al., 2017). Brittleness index (BI) is proposed based on the mineral
tattorni et al., 2015). The capillary pressure generated in the micro­ composition of the rock to evaluate the brittleness of rock (Altamar and
fractures and macropores is lower than the minor pore throats, which Marfurt, 2014; Jarvie et al., 2007) (Equation (16)). The high content of
accounts for the decrease in breakthrough pressure with the increase in brittle minerals, such as quartz, is important for the development of
the proportion of macropores (Zhou et al., 2019). Previous studies natural and created fractures (Zhou et al., 2022). The hardness growth
indicated that the specific area of pores decreases with increasing pore with increasing quartz content also contributes to the brittleness of the
size (Table 4). Pores smaller than 20 nm make up a significant propor­ rock (Li et al., 2021). The above discussion refers mainly to detrital
tion of the clay-rich rocks (Aplin and Moore, 2016; Yang and Aplin, quartz, however, quartz cement is often a major cause of porosity

9
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Fig. 11. Pressure and temperature as a function of depth (Pcr and Tcr are the critical pressure and temperature for CO2) (Espinoza et al., 2011).

connectivity of the pore structure and thus leads to an increase in the


Table 4
breakthrough pressure (Zhang and Yu, 2016). The breakthrough pres­
Specific surface area of pores of different sizes in the proposed previous studies
sure increases by approximately half an order of magnitude with in­
(unit m2/g).
crements of 10 wt% clay into a synthetic fault gouge (Zheng and
References Sample Micropore Mesopore Macropore Espinoza, 2021) as Equation (17). Thus, the higher the clay mineral
Zhang and Yu (2016) C012 6.372 2.497 0.0249 content is, the better the capping properties of the rock (Lü et al., 2017).
C022 12.446 1.940 0.0173 The mineral composition determines the breakthrough pressure by
C039 4.754 2.483 0.7721
affecting the pore structure characteristics of the matrix (Zhou et al.,
C040 18.035 5.733 0.8182
Zhang and Yu (2019) No.1 3.200 2.274 0.5704 2022). Besides, clays are always water-wet, and this contribute to the
No.2 3.545 1.178 0.4568 capillary pressure of a water saturated pore throat. Wettability of
Yang and Yu (2020) C01 5.133 2.542 0.173 sandstones increases with increasing clay content (Sayyouh et al., 1990).
C02 4.318 2.473 0.063
Thus, clay wettability plays a key role in the process of fluid spontane­
C03 6.138 3.233 0.454
ously imbibes into the rock matrix (Pan et al., 2020). From this point of
view, clay minerals also make for the breakthrough pressure
reduction in the reservoirs (Zhang et al., 2012a). In addition to its enhancement.
importance in controlling porosity of the rocks, quartz cement may 10
occlude fractures and contributes to strengthen the rock frame work Pc = 14.51⋅wf +8.12
(17)
1 + e−
(Makowitz et al., 2006; Vidar et al., 2005). It has also been shown that
the TOC content also contributes to the brittleness of the rock (Altamar where wf is the clay weight fraction, and e is the natural constant.
and Marfurt, 2014). From a view of mechanic, the rock is more brittle
with a lower Poisson’s Ration, while the situation for Young’s Modulus 4.2.4. Intrinsic permeability
is opposite (Rickman et al., 2008). The natural fractures generated under The intrinsic or absolute permeability (kabs) represents the conduc­
the mineral brittleness are more likely to occur near high structural tivity characteristics of a rock sample and is independent of the fluid
curvature areas and faults. type (Busch and Amann-Hildenbrand, 2013; Moghadam and Chala­
Qz turnyk, 2017; Yang and Aplin, 2010). It is a characteristic parameter of a
BI = (16) rock that determines the rate of flow of a fluid through it (Abbas et al.,
Qz + Ca + Cly
1999; Hildenbrand et al., 2002; Moghadam and Chalaturnyk, 2017), and
where Qz is the quartz content, Ca is the calcite content, and Cly is the it does not depend on temperature, pressure, or fluid type (Amann-­
clay content. Hildenbrand et al., 2012). The dominant pore radii of a rock are partly
Clay minerals control the transport properties of the fluid in clay-rich controlled by grain size (Dewhurst et al., 1998; Yang and Aplin, 2007).
rocks (Espinoza and Santamarina, 2012). Clay minerals are strongly At a given porosity, pore radii, and thus permeability are expected to
hydrophilic, with a water film bound on their surface (Li et al., 2016). increase as grain size increase (Ghanizadeh et al., 2015; Yang and Aplin,
Clay-bound water accounts for approximately 2.63–7.19% of the total 2010). According to a previous analysis, absolute permeability has a
pore volume of shale samples (Boyer et al., 2006; Tan et al., 2021). In negative impact on the breakthrough pressure (Pb) (Zhou et al., 2022)
addition, clay particles are fine-grained and have a large specific surface (Equation (18)):
area to absorb water and swell (Aksu et al., 2015; Gholami et al., 2021; − 0.407
Pb = 0.0252 × kabs + 0.5383 (18a)
Graham et al., 2002; Liu et al., 2021; Zan and Yu, 2020). For example,
montmorillonite expands by approximately 9% when exposed to CO2 in logPb = − 0.8112 × logkabs + 1.1549 (18b)
the presence of water within its interlayer structure (Giesting et al.,
2012; Johnston, 2010). Thus, the swelling of clay minerals reduces the

10
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

4.3. Saturation of the wetting phase clay minerals that encounter water also contributes to the breakthrough
pressure increase, particularly in clay-rich shales. According to the
Pore wetting is important to the migration of immiscible fluids (Li Laplace equation, the reduction of the effective pore throat caused by
and Fan, 2014). Fine-grained sediments are sensitive to water saturation both the bound water films and the swelling of clay minerals leads to an
(Espinoza and Santamarina, 2010). Most previous studies of break­ increase in the breakthrough pressure.
through pressure were performed on water-saturated rock samples.
Even if there are studies on the effect of water saturation, they typically
investigated clay minerals rather than real cores (Graham et al., 2002). 4.4. Non-wetting phase
However, rock formations are partially water-saturated underground
(Birdsell et al., 2015; Haghighi and Ahmad, 2013; Li et al., 2016). We Most breakthrough pressure studies have been performed as part of
conducted a series of breakthrough experiments to systematically GCS projects, and therefore, CO2 is typically used as the non-wetting
investigate the effect of water saturation on breakthrough pressure. phase (Busch and Müller, 2011; Espinoza and Santamarina, 2017; Gao
Whether using carbon dioxide, methane or even mixed gas, the break­ et al., 2014; Kim and Santamarina, 2013; Rezaeyan et al., 2010; Skurt­
through pressure increased exponentially with the water saturation in veit et al., 2012; Zhao and Yu, 2017; Zweigel et al., 2005). Some studies
both shales and sandstones (Equation (19)) (Zhang and Yu, 2016, 2019; compared the breakthrough pressure of CO2, N2, and CH4 in the same
Zhao and Yu, 2017, 2018): rock samples. Hildenbrand et al. (2004) found that the breakthrough
pressure for CO2 and N2 is lower than that for CH4 due to the IFT and
Pb = α × eβ×Sw (19) wettability difference of solid-liquid-gas systems. The temperature used
in that study was 30 ◦ C for CH4 and 50 ◦ C for CO2 and N2. Considering
where α and β are constants for a given sample, and Sw is the water the conclusion of Section 4.1.1, the IFT between CH4 and water should
saturation. be marginally larger at 50 ◦ C than at 30 ◦ C. Thus, the order of break­
The water saturation process of a dry rock is shown in Fig. 12 (Hil­ through pressure should be CO2 < N2 < CH4. Rezaeyan et al. (2015)
denbrand et al., 2002; Li et al., 2015; Zhang and Yu, 2016). With low derived that the breakthrough pressures for CO2, N2, and CH4 are 2.76,
water saturation, a small amount of water is attracted to the surface of 4.96, and 5.52 MPa at 90 ◦ C, respectively, because the IFTs of the
the pore walls and spreads along the particle surfaces under surface N2-brine and CH4-brine systems are higher than that of the CO2-brine
tension. When water passes the dead-end pores, the water is absorbed system (Busch and Amann-Hildenbrand, 2013; Li et al., 2006). Li et al.
and cannot move. Only a few dead-end pores are saturated with water, (2005) also found that CO2 breakthrough pressure is proportionally
and parts of the pore surface are attached to a thin water film. The lower than CH4 and N2. Kawaura et al. (2013) and Ono et al. (2013)
largest connected flow path is nearly unaffected during this stage. With showed that the measured value of the threshold capillary pressure for
increasing water saturation, the water film on the pore surface thickens supercritical CO2 is two-thirds smaller than that for N2 gas.
and thus forms a “water bridge” in the narrowest pore throat (the red The breakthrough pressure of gas mixtures for the low permeability
boxes in Fig. 12b). A proportion of the pore water can be adsorbed or rock samples was performed recently by the authors. Using different
bonded to mineral surfaces, making it difficult to be displaced (Har­ mole fractions of CO2–CH4 mixtures, the decrease in breakthrough
rington and Horseman, 1999). As a result, the breakthrough pressure pressure is proportional to the increase mole fraction of CO2 in the gas
required for the non-wetting phase increases as the capillary pressure in mixtures (Equation (20)) (Zhang and Yu, 2019; Zhao and Yu, 2018).
the pore throat increases. Until the water saturation reaches approxi­ This results was attributed to the IFT of the CH4-water system being
mately 60%, an increasing number of pore throats are sealed by the larger than that of the CO2-water system. In addition, the diameter of
“water bridge”, and the flow paths filled with water are caused by the CO2 molecules is marginally smaller than that of CH4 and can thus more
increased thickness of the bound water film on the pore surface easily form a continuous gas flow in low-porosity media for CO2:
(Fig. 12c). An immovable water film is formed and constrained on the
surface of the solid phase due to the attraction between the water Pb = − A⋅MCO2 + B (20)
molecules and the mineral particles (Yang and Yu, 2020), and many
narrow pores are sealed off; thus, the connectivity of the flow path is where MCO2 is the mole fraction of the CO2 in the gas mixtures, and A
reduced markedly. When the water saturation is above 85%, an and B are fitting parameters for a sample at a certain water saturation
increasing number of connected pathways are broken by the bound and are controlled by the pore structures and the mineral composition.
water, corresponding with a sharp decrease in the effective pore radius
(r in Equation (1)). Thus, the breakthrough pressure increases drastically 5. Conclusions
(Graham et al., 2002). As discussed in Section 4.2.3, the swelling of the
Breakthrough pressure is critical to consider when evaluating the

Fig. 12. Schematic diagram of the water saturation process of rocks in a connected flow path.

11
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

caprock sealing efficiency of greenhouse gas control storage and energy Aggelopoulos, C.A., Robin, M., Vizika, O., 2011. Interfacial tension between CO2 and
brine (NaCl+CaCl2) at elevated pressures and temperatures: the additive effect of
gas production. In this review, the terms of the capillary pressure in two-
different salts. Adv. Water Resour. 34 (4), 505–511.
phase flow in a porous media, the experimental methods for break­ Aksu, I., Bazilevskaya, E., Karpyn, Z.T., 2015. Swelling of clay minerals in
through pressure, and relevant factors were discussed in detail. unconsolidated porous media and its impact on permeability. Geo. Res. J. 7, 1–13.
Entry pressure is the pressure difference that exists across the gas- Al-Anssari, S., Barifcani, A., Keshavarz, A., Iglauer, S., 2018. Impact of nanoparticles on
the CO2-brine interfacial tension at high pressure and temperature. J. Colloid
liquid interface at the point when the non-wetting fluid begins to Interface Sci. 532, 136–142.
penetrate the connected pore space. The threshold pressure equals the Al-Yaseri, A., 2016. Receding and advancing (CO2 + brine + quartz) contact angles as a
gas pressure at which a continuous water flow comes out before function of pressure, temperature, surface roughness, salt type and salinity. J. Chem.
Thermodyn. 93, 416–423.
continuous gas flow forms. The breakthrough pressure is the pressure Al-Yaseri, A., Sarmadivaleh, M., Saeedi, A., Lebedev, M., Barifcani, A., Iglauer, S., 2015.
difference between the upstream and downstream across the core when N2+CO2+NaCl brine interfacial tensions and contact angles on quartz at CO2 storage
continuous gas flow is observed at the outlet and is widely applied site conditions in the Gippsland basin, Victoria/Australia. J. Petrol. Sci. Eng. 129,
58–62.
because it is easier to understand than the other three pressures. After Altamar, R.P., Marfurt, K., 2014. Mineralogy-based brittleness prediction from surface
gas breakthrough, the pressure decline of gas leads to water imbibition, seismic data: application to the Barnett Shale. Interpretation 2 (4), 255–271.
which causes the gas flow path to be blocked by imbibed water. At this Amann-Hildenbrand, A., Bertier, P., Busch, A., Krooss, B.M., 2013. Experimental
investigation of the sealing capacity of generic clay-rich caprocks. Int. J. Greenh. Gas
point, the pressure difference between the inlet and outlet is defined as Control 19, 620–641.
the snap-off pressure. The magnitudes of these pressures are arranged as Amann-Hildenbrand, A., Ghanizadeh, A., Krooss, B.M., 2012. Transport properties of
follows: breakthrough pressure > entry pressure > snap-off pressure. unconventional gas systems. Mar. Petrol. Geol. 31 (1), 90–99.
Amarasinghe, W.S., Fjelde, I., Flaata, A.M.N., 2021. Visual investigation of CO2
However, it is not easy to distinguish these pressures precisely due to
dissolution and convection in heterogeneous porous media at reservoir temperature
experimental complexities. and pressure conditions. Greenh. Gas. Sci. Technol. 11 (2), 342–359.
Direct methods include the SBS method, continuous injection An, Z., 2019. Water lock prevention/removal technology used in the “three-high” tight
method, dynamic method, and RCP method, and the MIP method is the sandstone gas reservoirs of the Tarim Oilfield. Oil Drill. Prod. Technol. 41 (6),
763–767.
most representative indirect method. Each of these methods has ad­ Andrew, M., Bijeljic, B., Blunt, M.J., 2014. Pore-by-pore capillary pressure measurements
vantages and limitations (Table 1). Overall, direct methods are more using X-ray microtomography at reservoir conditions: curvature, snap-off, and
representative than the MIP method, and the effective gas permeability remobilization of residual CO2. Water Resour. Res. 50 (11), 8760–8774.
Aplin, A.C., Macquaker, J.H.S., 2011. Mudstone diversity: origin and implications for
can be calculated by direct methods after gas breakthrough in a core source, seal, and reservoir properties in petroleum systems. AAPG Bull. 95 (12),
flooding system. 2031–2059.
The breakthrough pressure is determined by the capillary pressure of Aplin, A.C., Moore, J.K.S., 2016. Observations of Pore Systems of Natural Siliciclastic
Mudstones. The Clay Minerals Society Workshop Lectures Series, pp. 33–44.
the narrowest pore throat in the connected pathways and is influenced Aplin, A.C., Yang, Y., Hansen, S., 1995. Assessment of β the compression coefficient of
by the IFT, contact angle, and pore radius (Equation (1)). Environmental mudstones and its relationship with detailed lithology. Mar. Petrol. Geol. 12 (8),
factors such as temperature, pressure, ion type, ion concentration, and 955–963.
Arif, M., Al-Yaseri, A.Z., Barifcani, A., Lebedev, M., Iglauer, S., 2016a. Impact of pressure
non-wetting phase type affect the breakthrough pressure by determining and temperature on CO2-brine-mica contact angles and CO2-brine interfacial tension:
the interfacial properties (i.e., IFT and CA). The IFT increases with implications for carbon geo-sequestration. J. Colloid Interface Sci. 462, 208–215.
increasing salinity, cation valence, and temperature, and decreases with Arif, M., Barifcani, A., Iglauer, S., 2016b. Solid/CO2 and solid/water interfacial tensions
as a function of pressure, temperature, salinity and mineral type: implications for
decreasing pressure. The CA of a gas-brine-rock system increases with
CO2 -wettability and CO2 geo-storage. Int. J. Greenh. Gas Control 53, 263–273.
pressure, salinity and cation valence. Regarding the dependence of Arif, M., Jones, F., Barifcani, A., Iglauer, S., 2017a. Electrochemical investigation of the
temperature on CA, there is no agreement in the literature yet. The effect of temperature, salinity and salt type on brine/mineral interfacial properties.
different properties of the various gases also lead to differences in IFT Int. J. Greenh. Gas Control 59, 136–147.
Arif, M., Jones, F., Barifcani, A., Iglauer, S., 2017b. Influence of surface chemistry on
and wettability, thus resulting in different breakthrough pressures. interfacial properties of low to high rank coal seams. Fuel 194, 211–221.
With the increase of the burial depth and thickness of the formation, Babadagli, T., Hatiboglu, C.U., 2007. Analysis of counter-current gas–water capillary
the rock breakthrough pressure shows an increasing trend. The break­ imbibition transfer at different temperatures. J. Petrol. Sci. Eng. 55 (3–4), 277–293.
Bachu, Stefan, 2015. Review of CO2 storage efficiency in deep saline aquifers. Int. J.
through pressure increases as the sediment specific surface increases and Greenh. Gas Control 40, 188–202.
the porosity and absolute permeability decrease. The breakthrough Bachu, S., Bennion, D.B., 2009. Interfacial tension betweeen CO2, freshwater, and brine
pressure increases with increasing specific surface area and decreases in the range of pressure from (2 to 27) MPa, temperature from (20 to 125) ◦ C, and
water salinity from (0 to 334000) mg▪L-1. J. Chem. Eng. Data 54, 765–775.
porosity and permeability. Mineral composition affects breakthrough Bartlett, R.J., Sparks, D.L., 1986. Soil Redox Behavior, Soil Physical Chemistry. CRC
pressure by influencing the pore structure (e.g., brittle minerals tend to Press, Boca Raton.
form fractures; clay minerals prone to swell in the presence of water). Bernard, S., Horsfield, B., Schulz, H.M., Wirth, R., Schreiber, A., Sherwood, N., 2012.
Geochemical evolution of organic-rich shales with increasing maturity: a STXM and
The pressure increases with increasing water saturation because the TEM study of the Posidonia Shale (Lower Toarcian, northern Germany). Mar. Petrol.
presence of bound water dramatically reduces the pore connectivity. Geol. 31 (1), 70–89.
Different gas-liquid-solid system interface characteristics will also lead Bikkina, P.K., Shoham, O., Uppaluri, R., 2011. Equilibrated interfacial tension data of the
CO2–water system at high pressures and moderate temperatures. J. Chem. Eng. Data
to differences in breakthrough pressure.
56 (10), 3725–3733.
Birdsell, D.T., Rajaram, H., Lackey, G., 2015. Imbibition of hydraulic fracturing fluids
Declaration of competing interest into partially saturated shale. Water Resour. Res. 51 (8), 6787–6796.
Bocora, J., 2012. Global prospects for the development of unconventional gas. Procedia-
Sci. Behav. Sci. 65, 436–442.
The authors declare that they have no known competing financial Boulin, P.F., Bretonnier, P., Vassil, V., Samouillet, A., Fleury, M., Lombard, J.M., 2013.
interests or personal relationships that could have appeared to influence Sealing efficiency of caprocks: experimental investigation of entry pressure
the work reported in this paper. measurement methods. Mar. Petrol. Geol. 48, 20–30.
Boyer, C., Kieschnick, J., Suarez-Rivera, R., Lewis, R.E., Waters, G., 2006. Producing gas
from its source. Oilfield Rev. 18, 36–49.
References Busch, A., Alles, S., Gensterblum, Y., Prinz, D., Dewhurst, D., Raven, M., Stanjek, H.,
Krooss, B., 2008. Carbon dioxide storage potential of shales. Int. J. Greenh. Gas
Abbas, A., Carcasses, M., Ollivier, J.-P., 1999. Gas permeability of concrete in relation to Control 2 (3), 297–308.
its degree of saturation. Mater. Struct. 32 (215), 3–8. Busch, A., Amann-Hildenbrand, A., 2013. Predicting capillarity of mudrocks. Mar. Petrol.
Adamczyk, Z., Zaucha, M., Zembala, M., 2010. Zeta potential of mica covered by colloid Geol. 45, 208–223.
particles: a streaming potential study. Langmuir 26 (12), 9368–9377. Busch, A., Müller, N., 2011. Determining CO2/brine relative permeability and capillary
Aggelopoulos, C.A., Robin, M., Perfetti, E., Vizika, O., 2010. CO2/CaCl2 solution threshold pressures for reservoir rocks and caprocks: recommendations for
interfacial tensions under CO2 geological storage conditions: influence of cation development of standard laboratory protocols. Energy Proc. 4, 6053–6060.
valence on interfacial tension. Adv. Water Resour. 33 (6), 691–697. Cavanagh, A.J., Haszeldine, R.S., 2014. The Sleipner storage site: capillary flow
modeling of a layered CO2 plume requires fractured shale barriers within the Utsira
Formation. Int. J. Greenh. Gas Control 21, 101–112.

12
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Chalbaud, C., Robin, M., Lombard, J.M., Martin, F., Egermann, P., Bertin, H., 2009. Haghighi, M., Ahmad, M., 2013. Water Saturation Evaluation of Murteree and Roseneath
Interfacial tension measurements and wettability evaluation for geological CO2 Shale Gas Reservoirs Cooper Basin Australia Using Wire-Line Logs Focused Ion Beam
storage. Adv. Water Resour. 32 (1), 98–109. Milling and Scanning Electron Microscopy. SPE Unconventional Resources
Chen, C., Chai, Z., Shen, W., Li, W., Song, Y., 2017. Wettability of supercritical Conference, Brisbane, Australia, pp. 1–20.
CO2–brine–mineral: the effects of ion type and salinity. Energy Fuels 31 (7), Harrington, J.F., Horseman, S.T., 1999. Gas transport properties of clays and mudrocks.
7317–7324. Geol. Soc. London. Spec. Publ. 158 (1), 107–124.
Chiquet, P., Broseta, D., Thibeau, S., 2010. Wettability alteration of caprock minerals by Heath, J.E., Dewers, T.A., McPherson, B.J.O.L., Nemer, M.B., Kotula, P.G., 2012. Pore-
carbon dioxide. Geofluids 7 (2), 112–122. lining phases and capillary breakthrough pressure of mudstone caprocks: sealing
Chiquet, P., Daridon, J.-L., Broseta, D., Thibeau, S., 2007. CO2/water interfacial tensions efficiency of geologic CO2 storage sites. Int. J. Greenh. Gas Control 11, 204–220.
under pressure and temperature conditions of CO2 geological storage. Energy Henning, Dypvik, 1983. Clay mineral transformations in tertiary and mesozoic sediments
Convers. Manag. 48 (3), 736–744. from north sea. AAPG Bull. 67 (1), 160–165.
Cuéllar-Franca, R.M., Azapagic, A., 2015. Carbon capture, storage and utilisation Hesse, M.A., Woods, A.W., 2010. Buoyant dispersal of CO2 during geological storage.
technologies: a critical analysis and comparison of their life cycle environmental Geophys. Res. Lett. 37 (1), L01403.
impacts. J. CO2 Util. 9, 82–102. Hildenbrand, A., Schlömer, S., Krooss, B.M., 2002. Gas breakthrough experiments on
Cui, L., Ye, W., Wang, Q., Chen, Y., Chen, B., Cui, Y., 2019. Investigation on gas fine-grained sedimentary rocks. Geofluids 2 (1), 3–23.
migration in saturated bentonite using the residual capillary pressure technique with Hildenbrand, A., Schlömer, S., Krooss, B.M., Littke, R., 2004. Gas breakthrough
consideration of temperature. Process Saf. Environ. Protect. 125, 269–278. experiments on pelitic rocks: comparative study with N2, CO2 and CH4. Geofluids 4
De Silva, G.P.D., Ranjith, P.G., Perera, M.S.A., 2015. Geochemical aspects of CO2 (1), 61–80.
sequestration in deep saline aquifers: a review. Fuel 155, 128–143. Hu, H., 2013. Porosity evolution of the organic-rich shale with thermal maturity
Dewhurst, D.N., Aplin, A.C., Sarda, J.-P., Yang, Y., 1998. Compaction-driven evolution of increasing. Acta Pet. Sin. 34 (5), 820–825.
porosity and permeability in natural mudstones: an experimental study. J. Geophys. Iglauer, S., 2017. CO2–water–rock wettability: variability, influencing factors, and
Res. Solid Earth 103 (B1), 651–661. implications for CO2 geostorage. Acc. Chem. Res. 50 (5), 1134–1142.
Duan, W., Li, C.-F., Luo, C., Chen, X.-G., Bao, X., 2018. Effect of formation overpressure Iglauer, S., Pentland, C.H., Busch, A., 2015. CO2 wettability of seal and reservoir rocks
on the reservoir diagenesis and its petroleum geological significance for the DF11 and the implications for carbon geo-sequestration. Water Resour. Res. 51 (1),
block of the Yinggehai Basin, the South China Sea. Mar. Petrol. Geol. 97, 49–65. 729–774.
Duchateau, C., Broseta, D., 2012. A simple method for determining brine–gas interfacial Ito, D., Akaku, K., Okabe, T., Takahashi, T., Tsuji, T., 2011. Measurement of threshold
tensions. Adv. Water Resour. 42, 30–36. capillary pressure for seal rocks using the step-by-step approach and the residual
Egermann, P., Lombard, J.-M., Bretonnier, P., 2006. A Fast and Accurate Method to pressure approach. Energy Proc. 4, 5211–5218.
Measure Threshold Capillary Pressure of Caprocks under Representative Conditions. Jarvie, D.M., Hill, R.J., Ruble, T.E., Pollastro, R.M., 2007. Unconventional shale-gas
International Symposium of the Society of Core Analysts, Trondheim, Norway, systems: the Mississippian Barnett Shale of north-central Texas as one model for
pp. 1–14. thermogenic shale-gas assessment. AAPG Bull. 91 (4), 475–499.
Eseme, E., Krooss, B.M., Littke, R., 2012. Evolution of petrophysical properties of oil Johnston, C.T., 2010. Probing the nanoscale architecture of clay minerals. Clay Miner. 45
shales during high-temperature compaction tests: implications for petroleum (3), 245–279.
expulsion. Mar. Petrol. Geol. 31 (1), 110–124. Kawaura, K., Akaku, K., Nakano, M., Ito, D., Takahashi, T., Kiriakehata, S., 2013.
Espinoza, D.N., Kim, S.H., Santamarina, J.C., 2011. CO2 geological storage - geotechnical Examination of methods to measure capillary threshold pressures of pelitic rock
implications. KSCE J. Civ. Eng. 15 (4), 707–719. samples. Energy Proc. 37, 5411–5418.
Espinoza, D.N., Santamarina, J.C., 2010. Water-CO2 -mineral systems: interfacial tension, Kaya, A., Yukselen, Y., 2005. Zeta potential of clay minerals and quartz contaminated by
contact angle, and diffusion-Implications to CO2 geological storage. Water Resour. heavy metals. Can. Geotech. J. 42 (5), 1280–1289.
Res. 46 (7), W07537. Khosharay, S., 2015. Linear gradient theory for modeling investigation on the surface
Espinoza, D.N., Santamarina, J.C., 2012. Clay interaction with liquid and supercritical tension of (CH4+H2O), (N2+H2O) and (CH4+N2)+H2O systems. J. Nat. Gas Sci. Eng.
CO2: the relevance of electrical and capillary forces. Int. J. Greenh. Gas Control 10, 23, 474–480.
351–362. Kim, S., Santamarina, J.C., 2013. CO2 breakthrough and leak-sealing – experiments on
Espinoza, D.N., Santamarina, J.C., 2017. CO2 breakthrough-caprock sealing efficiency shale and cement. Int. J. Greenh. Gas Control 19, 471–477.
and integrity for carbon geological storage. Int. J. Greenh. Gas Control 66, 218–229. Krevor, S., Blunt, M.J., Benson, S.M., Pentland, C.H., Reynolds, C., Al-Menhali, A.,
Feng, D., Bakhshian, S., Wu, K., Song, Z., Ren, B., Li, J., Hosseini, S.A., Li, X., 2021. Niu, B., 2015. Capillary trapping for geologic carbon dioxide storage – from pore
Wettability effects on phase behavior and interfacial tension in shale nanopores. Fuel scale physics to field scale implications. Int. J. Greenh. Gas Control 40, 221–237.
290, 119983. Krevor, S.C.M., Pini, R., Zuo, L., Benson, S.M., 2012. Relative permeability and trapping
Ferrand, L.A., Celia, M.A., 1992. The effect of heterogeneity on the drainage capillary of CO2 and water in sandstone rocks at reservoir conditions. Water Resour. Res. 48
pressure-saturation relation. Water Resour. Res. 28 (3), 859–870. (2), W02532.
Foroutan, M., Ghazanfari, E., Amirlatifi, A., Perdrial, N., 2021. Variation of pore- Lai, J., Wang, G., Chai, Y., Xin, Y., Wu, Q., Zhang, X., Sun, Y., 2017. Deep burial
network, mechanical and hydrological characteristics of sandstone specimens diagenesis and reservoir quality evolution of high-temperature, high-pressure
through CO2-enriched brine injection. Geomec. Energy Environ. 26, 100217. sandstones: examples from Lower Cretaceous Bashijiqike Formation in Keshen area,
Fu, M.-y., Song, R.-c., Xie, Y.-h., Zhang, S.-n., Gluyas, J.G., Zhang, Y.-z., Zhang, Y., 2016. Kuqa depression, Tarim basin of China. AAPG Bull. 101 (6), 829–862.
Diagenesis and reservoir quality of overpressured deep-water sandstone following Laplace, P.S., 1806. Traité de Mécanique Céleste; Supplément au dixième livre, vol. 349.
inorganic carbon dioxide accumulation: upper Miocene Huangliu Formation, Sur l’Action Capillaire. Courcier, Paris.
Yinggehai Basin, South China Sea. Mar. Petrol. Geol. 77, 954–972. Li, J., Li, X., Wang, X., Li, Y., Wu, K., Shi, J., Yang, L., Feng, D., Zhang, T., Yu, P., 2016.
Fukuzawa, K., Watanabe, K., Yasuda, K., Ohmura, R., 2018. Interfacial tension Water distribution characteristic and effect on methane adsorption capacity in shale
measurements in the (CO2 + H2 ) gas mixture and water system at temperatures clay. Int. J. Coal Geol. 159, 135–154.
from 271.2 K to 280.2 K and pressures up to 7.0 MPa. J. Chem. Thermodyn. 119, Li, J., Yan, Q., Zhang, Y., Liu, G., Wang, X., 2008. The special sealing mechanism of
20–25. caprock for Quaternary biogenetic gas in Sanhu area, Qaidam Basin, China. Sci.
Gao, S., Wei, N., Li, X., 2015. Review of CO2 breakthrough pressure measurement China, Ser. Dokl. Earth Sci. 51, 45–52.
methods on caprocks. Rock Soil Mech. 36 (9), 2716–2727. Li, Q., Li, J., Duan, L., Tan, S., 2021. Prediction of rock abrasivity and hardness from
Gao, S., Wei, N., Li, X., Wang, Y., Wang, Q., 2014. Cap rock CO2 breakthrough pressure mineral composition. Int. J. Rock Mech. Min. Sci. 140, 104658.
measurement apparatus and application in Shenhua CCS project. Energy Proc. 63, Li, S., Dong, M., Z, L.I., Huang, S., Qing, H., Nickel, E., 2005. Gas breakthrough pressure
4766–4772. for hydrocarbon reservoir seal rocks: implications for the security of long-term CO2
Gelhar, L., 1986. Stochastic subsurface hydrology from theory to applications. Water storage in the Weyburn field. Geofluids 5 (4), 326–334.
Resour. Res. 22 (9), 135–145. Li, S., Wo, Y., Yan, Z., Liu, W., 2011. Controlling factors affect sealing capability of well-
Ghaedi, M., Riazi, M., 2016. Scaling equation for counter current imbibition in the developed muddy cap rock. Acta Geol. Sin. 85 (10), 1691–1697.
presence of gravity forces considering initial water saturation and SCAL properties. Li, X., Boek, E., Maitland, G.C., Trusler, J.P.M., 2012. Interfacial tension of (Brines +
J. Nat. Gas Sci. Eng. 34, 934–947. CO2): (0.864 NaCl + 0.136 KCl) at temperatures between (298 and 448) K, pressures
Ghanizadeh, A., Clarkson, C.R., Aquino, S., Ardakani, O.H., Sanei, H., 2015. between (2 and 50) MPa, and total molalities of (1 to 5) mol⋅kg–1. J. Chem. Eng. Data
Petrophysical and geomechanical characteristics of Canadian tight oil and liquid-rich 57 (4), 1078–1088.
gas reservoirs: I. Pore network and permeability characterization. Fuel 153, Li, X., Elsworth, D., 2015. Geomechanics of CO2 enhanced shale gas recovery. J. Nat. Gas
664–681. Sci. Eng. 26, 1607–1619.
Gholami, R., Raza, A., Andersen, P., Escalona, A., Cardozo, N., Marín, D., Li, X., Fan, X., 2014. Pore wetting and its effect on breakthrough pressure in water-wet
Sarmadivaleh, M., 2021. Long-term integrity of shaly seals in CO2 geo-sequestration and oil-wet pores. Int. J. Chem. Eng. Appl. 5 (4), 359–362.
sites: an experimental study. Int. J. Greenh. Gas Control 109, 103370. Li, Y., Li, X., Wang, Y., Yu, Q., 2015. Effects of composition and pore structure on the
Giesting, P., Guggenheim, S., Koster van Groos, A.F., Busch, A., 2012. Interaction of reservoir gas capacity of Carboniferous shale from Qaidam Basin, China. Mar. Petrol.
carbon dioxide with Na-exchanged montmorillonite at pressures to 640bars: Geol. 62, 44–57.
implications for CO2 sequestration. Int. J. Greenh. Gas Control 8, 73–81. Li, Y., Ranjith, P.G., Perera, M.S.A., Yu, Q., 2017. Residual water formation during the
Graham, J., Halayko, K.G., Hume, H., Kirkham, T., Oscarson, D., 2002. A capillary- CO2 storage process in deep saline aquifers and factors influencing it: a review.
advective model for gas break-through in clays. Eng. Geol. 64 (2–3), 273–286. J. CO2 Util. 20, 253–262.
Guiltinan, E.J., Espinoza, D.N., Cockrell, L.P., Cardenas, M.B., 2018. Textural and Li, Z., Dong, M., Li, S., Huang, S., 2006. CO2 sequestration in depleted oil and gas
compositional controls on mudrock breakthrough pressure and permeability. Adv. reservoirs—caprock characterization and storage capacity. Energy Convers. Manag.
Water Resour. 121, 162–172. 47 (11–12), 1372–1382.

13
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Liang, X., Liang, T., Zhou, F., Wang, C., Zhang, D., 2021. Impact of shut-in time on Ono, M., Kameya, H., Hosoda, K., Kamidozono, Y., Azuma, H., 2013. Experimental
production after hydraulic fracturing in fractured shale gas formation: an measurements of threshold pressure for modeling saline aquifers in Japan. Energy
experimental study. J. Nat. Gas Sci. Eng. 88, 103773. Proc. 37, 5456–5463.
Lin, M., Wang, Y., Cao, Y., Wang, Y., Wang, X., Xi, K., 2020. Experimental study of the Pan, B., Li, Y., Wang, H., Jones, F., Iglauer, S., 2018. CO2 and CH4 wettabilities of
influence of oil-wet calcite cements on oil migration and implications for clastic organic-rich shale. Energy Fuels 32 (2), 1914–1922.
reservoirs. Mar. Petrol. Geol. 118, 104427. Pan, B., Yin, X., Iglauer, S., 2020. A review on clay wettability: from experimental
Liu, C., Sang, S., Zhang, K., Song, F., Wang, H., Fan, X., 2019. Effects of temperature and investigations to molecular dynamics simulations. Adv. Colloid Interface Sci. 285,
pressure on pore morphology of different rank coals: implications for CO2 geological 102266.
storage. J. CO2 Util. 34, 343–352. Peter, K.W., Robert, J.P., 1996. Surface tension of aqueous solutions of electrolytes:
Liu, D., Ren, D., Du, K., Qi, Y., Ye, F., 2021. Impacts of mineral composition and pore relationship with ion hydration, oxygen solubility, and bubble coalescence.
structure on spontaneous imbibition in tight sandstone. J. Petrol. Sci. Eng. 201, J. Colloid Interface Sci. 184 (2), 550–563.
108397. Purcell, W.R., 1949. Capillary pressures-Their measurement using mercury and the
Liu, Y., Li, H.A., Okuno, R., 2016. Measurements and modeling of interfacial tension for calculation of permeability therefrom. J. Petrol. Technol. 186 (2), 39–48.
CO2/CH4/Brine systems under reservoir conditions. Ind. Eng. Chem. Res. 55 (48), Ralston, J., Healy, T.W., 1973. Specific cation effects on water structure at the air-water
12358–12375. and air-octadecanol monolayer-water interfaces. J. Colloid Interface Sci. 42 (3),
Liu, Y., Tang, J., Wang, M., Wang, Q., Tong, J., Zhao, J., Song, Y., 2017. Measurement of 629–644.
interfacial tension of CO2 and NaCl aqueous solution over wide temperature, Ranjith, P.G., Perera, M.S.A., Khan, E., 2013. A study of safe CO2 storage capacity in
pressure, and salinity ranges. J. Chem. Eng. Data 62 (3), 1036–1046. saline aquifers: a numerical study. Int. J. Energy Res. 37 (3), 189–199.
Lu, J., Kharaka, Y.K., Thordsen, J.J., Horita, J., Karamalidis, A., Griffith, C., Hakala, J.A., Rashid, S., Harimi, B., Hamidpour, E., 2017. Prediction of CO2-Brine interfacial tension
Ambats, G., Cole, D.R., Phelps, T.J., Manning, M.A., Cook, P.J., Hovorka, S.D., 2012. using a rigorous approach. J. Nat. Gas Sci. Eng. 45, 108–117.
CO2–rock–brine interactions in Lower Tuscaloosa Formation at Cranfield CO2 Raziperchikolaee, S., Kelley, M., Gupta, N., 2019. A screening framework study to
sequestration site, Mississippi. U.S.A. Chem. Geol. 291, 269–277. evaluate CO2 storage performance in single and stacked caprock–reservoir systems of
Lü, X., Qu, Y., Yu, H., Lan, X., 2014. Sealing capacity of carbonate cap rocks:A case study the Northern Appalachian Basin. Greenh. Gas. Sci. Technol. 9 (3), 582–605.
of Ordovician in northern slope of central Tarim Basin. Petrol. Geol. Exp. 36 (5), Ren, Q.Y., Chen, G.J., Yan, W., Guo, T.M., 2000. Interfacial Tension of (CO2+CH4) plus
532–538. water from 298 K to 373 K and pressures up to 30 MPa. J. Chem. Eng. Data 45 (4),
Lü, X., Wang, Y., Yu, H., Bai, Z., 2017. Major factors affecting the closure of marine 1004–1020.
carbonate caprock and their quantitative evaluation: a case study of Ordovician Rezaeyan, A., Tabatabaei-Nejad, S.A., Khodapanah, E., Kamari, M., 2015. A laboratory
rocks on the northern slope of the Tazhong uplift in the Tarim Basin, western China. study on capillary sealing efficiency of Iranian shale and anhydrite caprocks. Mar.
Mar. Petrol. Geol. 83, 231–245. Petrol. Geol. 66, 817–828.
Lun, Z., Fan, H., Wang, H., Luo, M., Pan, W., Wang, R., 2012. Interfacial tensions Rezaeyan, A., Tabatabaei-Nejad, S.A., Khodepanah, E., 2010. Parametric analysis of
between reservoir brine and CO2 at high pressures for different salinity. Energy Fuels caprock integrity in relation with CO2 geosequestration: capillary breakthrough
26 (6), 3958–3962. pressure of caprock and gas effective permeability. Greenh. Gas. Sci. Technol. 4 (3),
Luquot, L., Roetting, T.S., Carrera, J., 2014. Characterization of flow parameters and 714–731.
evidence of pore clogging during limestone dissolution experiments. Water Resour. Rezk, M.G., Foroozesh, J., Zivar, D., Mumtaz, M., 2019. CO2 storage potential during CO2
Res. 50 (8), 6305–6321. enhanced oil recovery in sandstone reservoirs. J. Nat. Gas Sci. Eng. 66, 233–243.
Lv, G., Li, Q., Wang, S., Li, X., 2015. Key techniques of reservoir engineering and Rickman, R., Mullen, M., Petre, E., Grieser, B., Kundert, D., 2008. A Practical Use of Shale
injection–production process for CO2 flooding in China’s SINOPEC Shengli Oilfield. Petrophysics for Stimulation Design Optimization: All Shale Plays Are Not Clones of
J. CO2 Util. 11, 31–40. the Barnett Shale. SPE Annual Technical Conference and Exhibition, Denver,
Lyu, Y., Dasani, D., Tsotsis, T., Jessen, K., 2021. Characterization of shale using Helium Colorado, USA.
and Argon at high pressures. J. Petrol. Sci. Eng. 206, 108952. Rosenbauer, R.J., Koksalan, T., Palandri, J.L., 2005. Experimental investigation of
Ma, C., Lin, C., Dong, C., Luan, G., Zhang, Y., Sun, X., Liu, X., 2019. Quantitative CO2–brine–rock interactions at elevated temperature and pressure: implications for
relationship between argillaceous caprock thickness and maximum sealed CO2 sequestration in deep-saline aquifers. Fuel Process. Technol. 86 (14–15),
hydrocarbon column height. Nat. Resour. Res. 29 (3), 2033–2049. 1581–1597.
Makowitz, A., Lander, R.H., Milliken, K.L., 2006. Diagenetic modeling to assess the Roshan, H., Al-Yaseri, A.Z., Sarmadivaleh, M., Iglauer, S., 2016. On wettability of shale
relative timing of quartz cementation and brittle grain processes during compaction. rocks. J. Colloid Interface Sci. 475, 104–111.
AAPG Bull. 90 (6), 873–885. Rudd, N., Pandey, G.N., 1973. Threshold pressure profiling by continuous injection. In:
Manciu, M., Ruckenstein, E., 2003. Specific ion effects via ion hydration: I. Surface 48th Annual Fall Meeting of the Society of Petroleum Engineers of AIME. Society of
tension. Adv. Colloid Interface Sci. 105 (1–3), 63–101. Petroleum Engineering, las Vegas.
Mayer, R.P., Stowe, R.A., 1965. Mercury porosimetry—breakthrough pressure for Saraji, S., Piri, M., Goual, L., 2014. The effects of SO2 contamination, brine salinity,
penetration between packed spheres. J. Colloid Sci. 20, 893–911. pressure, and temperature on dynamic contact angles and interfacial tension of
McGlade, C., Speirs, J., Sorrell, S., 2013. Unconventional gas – a review of regional and supercritical CO2/brine/quartz systems. Int. J. Greenh. Gas Control 28, 147–155.
global resource estimates. Energy 55, 571–584. Sarmadivaleh, M., Al-Yaseri, A.Z., Iglauer, S., 2015. Influence of temperature and
Meckel, T.A., Bryant, S.L., Ravi Ganesh, P., 2015. Characterization and prediction of CO2 pressure on quartz-water-CO2 contact angle and CO2-water interfacial tension.
saturation resulting from modeling buoyant fluid migration in 2D heterogeneous J. Colloid Interface Sci. 441, 59–64.
geologic fabrics. Int. J. Greenh. Gas Control 34, 85–96. Sayyouh, M.H., Dahab, A.S., Omar, A.E., 1990. Effect of clay content on wettability of
Mikunda, T., Brunner, L., Skylogianni, E., Monteiro, J., Rycroft, L., Kemper, J., 2021. sandstone reservoirs. J. Petrol. Sci. Eng. 4 (2), 119–125.
Carbon capture and storage and the sustainable development goals. Int. J. Greenh. Schlmer, S., Krooss, B.M., 1997. Experimental characterization of the hydrocarbon
Gas Control 108, 103318. sealing efficiency of cap rocks. Mar. Petrol. Geol. 14 (5), 565–580.
Minardi, A., Stavropoulou, E., Kim, T., Ferrari, A., Laloui, L., 2021. Experimental Schowalter, T.T., 1979. Mechanics of secondary hydrocarbon migration and entrapment.
assessment of the hydro-mechanical behaviour of a shale caprock during CO2 AAPG Bull. 63 (5), 723–760.
injection. Int. J. Greenh. Gas Control 106, 103225. Scotchman, I.C., 2016. Shale gas and fracking: exploration for unconventional
Moghadam, A., Vaisblat, N., Harris, N.B., Chalaturnyk, R., 2020. On the magnitude of hydrocarbons. Proc. Geologists’ Assoc. 127 (5), 535–551.
capillary pressure (suction potential) in tight rocks. J. Petrol. Sci. Eng. 190, 107133. Skurtveit, E., Aker, E., Soldal, M., Angeli, M., Wang, Z., 2012. Experimental investigation
Moghadam, A.A., Chalaturnyk, R., 2017. Rate dependency of permeability in tight rocks. of CO2 breakthrough and flow mechanisms in shale. Petrol. Geosci. 18 (1), 3–15.
J. Nat. Gas Sci. Eng. 40, 208–225. Smith, D.A., 1966. Theoretical considerations of sealing and non-sealing faults. AAPG
Moore, J., Holcomb, P., Crandall, D., King, S., Choi, J.-H., Brown, S., Workman, S., 2021. Bull. 50 (2), 363–374.
Rapid determination of supercritical CO2 and brine relative permeability using an Song, J., Zhang, D., 2012. Comprehensive review of caprock-sealing mechanisms for
unsteady-state flow method. Adv. Water Resour. 153, 103953. geologic carbon sequestration. Environ. Sci. Technol. 47 (1), 9–22.
Mutailipu, M., Liu, Y., Jiang, L., Zhang, Y., 2019. Measurement and estimation of CO2- Song, Y., Davy, C.A., Troadec, D., 2016. Gas breakthrough pressure (GBP) through
brine interfacial tension and rock wettability under CO2 sub- and super-critical claystones: correlation with FIB/SEM imaging of the pore volume. Oil. Gas. Sci.
conditions. J. Colloid Interface Sci. 534, 605–617. Tech. Rev. IFP Energy Nouvelles 71 (51), 1–16.
Naderi, S., Simjoo, M., 2018. Numerical study of Low Salinity Water Alternating CO2 Stevar, M.S.P., Böhm, C., Notarki, K.T., Trusler, J.P.M., 2019. Wettability of calcite under
injection for enhancing oil recovery in a sandstone reservoir: coupled geochemical carbon storage conditions. Int. J. Greenh. Gas Control 84, 180–189.
and fluid flow modeling. J. Petrol. Sci. Eng. 173, 279–286. Sun, W., Wang, L., Wang, Y., 2017. Mechanical properties of rock materials with related
Nojabaei, B., Siripatrachai, N., Johns, R.T., Ertekin, T., 2016. Effect of large gas-oil to mineralogical characteristics and grain size through experimental investigation: a
capillary pressure on production: a compositionally-extended black oil formulation. comprehensive review. Front. Struct. Civ. Eng. 11 (3), 322–328.
J. Petrol. Sci. Eng. 147, 317–329. Sun, Y., Liu, J., Xue, Z., Li, Q., Fan, C., Zhang, X., 2021. A critical review of distributed
Norhasyima, R.S., Mahlia, T.M.I., 2018. Advances in CO₂ utilization technology: a patent fiber optic sensing for real-time monitoring geologic CO2 sequestration. J. Nat. Gas
landscape review. J. CO2 Util. 26, 323–335. Sci. Eng. 88, 103751.
Oliveira de Moura e Silva, G., Lamas, L.F., Soares, A.P., 2021. Characterization of Tan, Y., Zhang, S., Tang, S., Cui, G., Ma, Y., Sun, M., Pan, Z., 2021. Impact of water
carbonate facies and study of their relationship with dissolution and wettability in saturation on gas permeability in shale: experimental and modelling. J. Nat. Gas Sci.
CO2 injection. J. Petrol. Sci. Eng., 108938 Eng. 95, 104062.
Omotilewa, O.J., Panja, P., Vega-Ortiz, C., McLennan, J., 2021. Evaluation of enhanced Tanai, K., Kanno, T., Galle, C., 1997. Experimental study of gas permeabilities and
coalbed methane recovery and carbon dioxide sequestration potential in high breakthrough pressures in clays. Mater. Res. Soc. Symp. Proc. 465, 995–1002.
volatile bituminous coal. J. Nat. Gas Sci. Eng. 91. Thomas, L.K., Katz, D.L., Tek, M.R., 1968. Threshold pressure phenomena in porous
media. Soc. Petrol. Eng. J. 243, 174–184.

14
C. Zhang and M. Wang Journal of Natural Gas Science and Engineering 100 (2022) 104456

Vavra, C.L., Kaldi, J.G., Sneider, R.M., 1992. Geological applications of capillary- Yang, Y., Aplin, A.C., 2010. A permeability–porosity relationship for mudstones. Mar.
pressure: a review. AAPG Bull. 7 (66), 840–850. Petrol. Geol. 27 (8), 1692–1697.
Verdugo, M., Doster, F., 2022. Impact of capillary pressure and flowback design on the Yekeen, N., Padmanabhan, E., Abdulelah, H., Irfan, S.A., Okunade, O.A., Khan, J.A.,
clean up and productivity of hydraulically fractured tight gas wells. J. Petrol. Sci. Negash, B.M., 2021. CO2/brine interfacial tension and rock wettability at reservoir
Eng. 208, 109465. conditions: a critical review of previous studies and case study of black shale from
Vidar, Storvoll, Knut, Bjrlykke, Nazmul, H., Mondol, 2005. Velocity-depth trends in Malaysian formation. J. Petrol. Sci. Eng. 196, 107673.
mesozoic and cenozoic sediments from the Norwegian shelf. AAPG Bull. 89 (3), Yoksoulian, L.E., Freiburg, J.T., Butler, S.K., Berger, P.M., Roy, W.R., 2013.
359–381. Mineralogical alterations during laboratory-scale carbon sequestration experiments
Vilarrasa, V., Makhnenko, R.Y., 2017. Caprock integrity and induced seismicity from for the Illinois basin. Energy Proc. 37, 5601–5611.
laboratory and numerical experiments. Energy Proc. 125, 494–503. Yong, W., Derksen, J., Zhou, Y., 2021. The influence of CO2 and CH4 mixture on water
Voltattorni, N., Lombardi, S., Beaubien, S.E., 2015. Gas migration from two mine wettability in organic rich shale nanopore. J. Nat. Gas Sci. Eng. 87, 103746.
districts: the Tolfa (Lazio, Central Italy) and the Neves-Corvo (Baixo Alentejo, Young, T., 1805. An essay on the cohesion of fluids. Phil. Trans. Roy. Soc. Lond. 95,
Portugal) case studies. J. Geochem. Explor. 152, 37–53. 65–87.
Wang, J., Liu, H., Yu, W., Cao, F., Sepehrnoori, K., 2016. Necessity of porosity correction Yu, X., Li, J., Chen, Z., Wu, K., Zhang, L., Hui, G., Yang, M., 2020. Molecular dynamics
before simulation and re-understanding of the effects of gas adsorption on computations of brine-CO2/CH4-shale contact angles: implications for CO2
production in shale gas reservoirs. J. Petrol. Sci. Eng. 139, 162–170. sequestration and enhanced gas recovery. Fuel 280, 118590.
Wang, M., Yu, Q., 2019. A method to determine the permeability of shales by using the Zan, Y., Yu, Q., 2020. Experimental investigation of spontaneous water imbibition into
dynamic process data of methane adsorption. Eng. Geol. 253, 111–122. methane-saturated shales under different methane pressures. Energy Fuels 34 (11),
Washburn, E.W., 1921a. The dynamics of capillary flow. Phys. Rev. 17 (3), 273–283. 14356–14367.
Washburn, E.W., 1921b. Note on a method of determining the distribution of pore sizes Zhang, B., Zhang, J., Yan, S., Gu, Z., Wang, X., 2012a. Detrital quartz and quartz cement
in a porous material. Proc. Natl. Acad. Sci. U.S.A. 7 (4), 115–116. in Upper Triassic reservoir sandstones of the Sichuan basin: characteristics and
Wen, B., Shi, Z., Jessen, K., Hesse, M.A., Tsotsis, T.T., 2021. Convective carbon dioxide mechanisms of formation based on cathodoluminescence and electron backscatter
dissolution in a closed porous medium at high-pressure real-gas conditions. Adv. diffraction analysis. Sediment. Geol. 267–268, 104–114.
Water Resour. 154, 103950. Zhang, C., Yu, Q., 2016. The effect of water saturation on methane breakthrough
Wigley, M., Kampman, N., Dubacq, B., Bickle, M., 2017. Fluid-mineral reactions and pressure: an experimental study on the Carboniferous shales from the eastern
trace metal mobilization in an exhumed natural CO2 reservoir, Green River, Utah. Qaidam Basin, China. J. Hydrol. 543, 832–848.
Geology 40 (6), 555–558. Zhang, C., Yu, Q., 2019. Breakthrough pressure and permeability in partially water-
Wollenweber, J., Alles, S., Busch, A., Krooss, B.M., Stanjek, H., Littke, R., 2010. saturated shales using methane–carbon dioxide gas mixtures: an experimental study
Experimental investigation of the CO2 sealing efficiency of caprocks. Int. J. Greenh. of Carboniferous shales from the eastern Qaidam Basin, China. AAPG Bull. 103 (2),
Gas Control 4 (2), 231–241. 273–301.
Wu, T., Pan, Z.J., Connell, L.D., Liu, B., Fu, X.F., Xue, Z.Q., 2020. Gas breakthrough Zhang, G., Liu, X., Bi, Y., Pu, W., 2011. Experimental study of penetrant solution impact
pressure of tight rocks: a review of experimental methods and data. J. Nat. Gas Sci. on gas desorption. Procedia Eng. 26, 113–119.
Eng. 81, 103408. Zhang, L., Bai, G., Zhao, Y., 2012b. A method for eliminating caprock thickness influence
Xu, L., Ye, W.M., Chen, Y.G., Chen, B., Cui, Y.J., 2018. A new approach for determination on anomaly intensities in geochemical surface survey for hydrocarbons. Geomath.
of gas breakthrough in saturated materials with low permeability. Eng. Geol. 241, Geosci. 44 (8), 929–944.
121–131. Zhao, Y., Yu, Q., 2017. CO2 breakthrough pressure and permeability for unsaturated low-
Xu, L., Ye, W.M., Ye, B., Chen, B., Chen, Y.G., Cui, Y.J., 2015. Investigation on gas permeability sandstone of the Ordos Basin. J. Hydrol. 550, 331–342.
migration in saturated materials with low permeability. Eng. Geol. 197, 94–102. Zhao, Y., Yu, Q., 2018. Effect of CH4 on the CO2 breakthrough pressure and permeability
Yan, W., Zhao, G.Y., Chen, G.J., Guo, T.M., 2001. Interfacial tension of (methane + of partially saturated low-permeability sandstone in the Ordos Basin, China.
nitrogen) + water and (carbon dioxide + nitrogen) + water systems. J. Chem. Eng. J. Hydrol. 556, 732–748.
Data 46 (6), 1544–1548. Zheng, X., Espinoza, D.N., 2021. Multiphase CO2-brine transport properties of synthetic
Yang, C., Zhang, J., Tang, X., Ding, J., Zhao, Q., Dang, W., Chen, H., Su, Y., Li, B., Lu, D., fault gouge. Mar. Petrol. Geol. 129, 105054.
2017. Comparative study on micro-pore structure of marine, terrestrial, and Zhou, X., Lü, X., Quan, H., Qian, W., Mu, X., Chen, K., Wang, Z., Bai, Z., 2019. Influence
transitional shales in key areas, China. Int. J. Coal Geol. 171, 76–92. factors and an evaluation method about breakthrough pressure of carbonate rocks:
Yang, S., Yu, Q., 2020. Experimental investigation on the movability of water in shale an experimental study on the Ordovician of carbonate rock from the Kalpin area,
nanopores: a case study of carboniferous shale from the Qaidam basin, China. Water Tarim Basin, China. Mar. Petrol. Geol. 104, 313–330.
Resour. Res. 56 (8). Zhou, X., Lü, X., Sui, F., Wang, X., Li, Y., 2022. The breakthrough pressure and sealing
Yang, Y., Aplin, A.C., 1998. Influence of lithology and compaction on the pore size property of Lower Paleozoic carbonate rocks in the Gucheng area of the Tarim Basin.
distribution and modelled permeability of some mudstones from the Norwegian J. Petrol. Sci. Eng. 208, 109289.
margin. Mar. Petrol. Geol. 15, 163–175. Zweigel, P., Lindeberg, E., Moen, A., Wessel-Berg, D., 2005. Towards a methodology for
Yang, Y., Aplin, A.C., 2007. Permeability and petrophysical properties of 30 natural top seal efficacy assessment for underground CO2 storage. In: Rubin E.S. et al. (Eds.),
mudstones. J. Geophys. Res. 112 (B3). Greenhouse Gas Control Technologies 7. Elsevier Science Ltd, Oxford, 1323-1328.

15

You might also like