Next Article in Journal
Acid-Extracted Hydrocarbon Anomalies and Significance in the Chaoshan Depression of the Northern South China Sea
Previous Article in Journal
Analysis of Human-Related Incidents during Dynamic Positioning Operations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microscopic Characterization and Fractal Analysis of Pore Systems for Unconventional Reservoirs

1
Beijing International Center for Gas Hydrate, School of Earth and Space Sciences, Peking University, Beijing 100871, China
2
School of Environment and Energy, Peking University Shenzhen Graduate School, Shenzhen 518055, China
3
College of Chemical Engineering, Fuzhou University, Fuzhou 350116, China
4
National Engineering Research Center for Gas Hydrate Exploration and Development, Guangzhou 511466, China
*
Authors to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(6), 908; https://doi.org/10.3390/jmse12060908
Submission received: 25 April 2024 / Revised: 19 May 2024 / Accepted: 25 May 2024 / Published: 29 May 2024
(This article belongs to the Section Geological Oceanography)

Abstract

:
The complex pore structure of unconventional oil and gas reservoirs is one of the reasons for the difficulties in resource evaluation and development. Therefore, it is crucial to comprehensively characterize the pore structure, understand reservoir heterogeneity from multiple perspectives, and gain an in-depth understanding of fluid migration and accumulation mechanisms. This review outlines the methods and basic principles for characterizing microporous systems in unconventional reservoirs, summarizes the fractal analysis corresponding to the different methods, sorts out the relationship between the fractals and reservoir macroscopic physical properties (porosity, permeability, etc.) with the reservoir microscopic pore structures (pore structure parameters, pore connectivity, etc.). The research focuses on cutting-edge applications of characterization techniques, such as improved characterization accuracy, calibration of PSD ranges, and identification of different hydrogen compositions in pore systems for dynamic assessment of unconventional reservoirs. Fractal dimension analysis can effectively identify the quality level of the reservoir; complex pore-throat structures reduce permeability and destroy free fluid storage space, and the saturation of removable fluids is negatively correlated with Df. As for the mineral composition, the fractal dimension is positively correlated with quartz, negatively correlated with feldspar, and weakly correlated with clay mineral content. In future qualitative characterization studies, the application and combination of contrast agents, molecular dynamics simulations, artificial intelligence techniques, and 4D imaging techniques can effectively improve the spatial resolution of the images and explore the adsorption/desorption of gases within the pores, and also help to reduce the computational cost of these processes; these could also attempt to link reservoir characterization to research on supercritical carbon dioxide-enhanced integrated shale gas recovery, carbon geological sequestration, and advanced underground hydrogen storage.

1. Introduction

1.1. Unconventional Reservoirs

The microscopic pore system, which is the main control parameter for fluid movement, controls the reservoir’s macroscopic properties (porosity and permeability) [1,2]. Depositional environment and diagenetic alteration resulted in a diverse range of pore sizes, pore types, developed nanoscale pores, complex structures and massive nonhomogeneity. Exploration and exploitation of unconventional resources need a comprehensive understanding of the characteristics of the pore system since all of these factors impact the reservoirs’ storage performance [3,4,5].
Unconventional resource types come in different varieties (Table 1). The primary types of unconventional oil resources are tight oil, shale oil, thick oil, oil sands, and oil shale. Meanwhile, the main types of unconventional natural gas resources are tight sandstone [6,7], shale gas [2,8], coal bed methane [9], natural gas hydrate [10], etc.
Reservoirs are generally divided into clastic reservoir (sandstone, mudstone, shale), igneous reservoir and metamorphic reservoir. Originally loose and deposited in low-lying areas, the reservoir sediments were mainly caused by the damaging effects of weathering and erosion of primary rocks. After extended periods of sedimentary evolution, the overlying sediments gradually thickened, while the lower sediments were buried deeper, and eventually solidified into rocks as a result of compression, dehydration and cementation effects.

1.2. Pore System

The microscopic structure of pores or throats refers to the width, size distribution, geometric, topological properties and connectivity of pores, which affects the storage performance of reservoirs (Figure 1). The features of unconventional reservoir pore systems are loss of primary porosity, mainly developing secondary (dissolved) pores and micropores, and the existence of slits [11]. When characterizing micropores, it is necessary to select appropriate methods for joint characterization [12,13].
Currently, the most extensively used pore size classification systems are the Hodot pore size classification (micropore < 10 nm; micropore 10–100 nm; mesopore 100–1000 nm; macropore > 1000 nm) and the IUPAC pore classification (micropore < 2 nm; mesopore 2–50 nm; macropore > 50 nm) [11]. Furthermore, pores can be classified as intergranular pores, intragranular pores and microcracks, etc. (Figure 1) according to the pore morphology, the pores can also be classified into brittle mineral pores, clay mineral pores and other mineral pores based on the mineral components, spatial position and genesis (Table 2).
As the channel connecting pores, the size and shape of the throat and the way it connects with pores control the seepage capacity of the reservoir. Table 3 [14] lists the four different forms of throat.
For the purpose of identifying and evaluating reservoirs, estimating oil and gas capacity, and enhancing oil and gas recovery, pore system characterization is crucial [2,15,16]. Pore structure, quantitative parameter characterization, reservoir classification, and evaluation are the key areas of research. The geometry, size, pore size distribution, interconnection, and composition of pores and throats are all regarded as parts of pore structure. These characteristics may reflect the combination or interaction of various kinds of pores [3,17,18].

1.3. Methods for Reservoir Characterization

The reservoir micropore system can be qualitatively and quantitatively analyzed using a variety of techniques.
Qualitative methods include optical microscope (OM), casting thin section (CTS) [19], field emission scanning electron microscopy (FESEM) [2,16,20,21,22,23,24], transmission electron microscope (TEM) [25,26,27,28,29], focused ion beam scanning electron microscopy (FIBSEM) [30,31,32,33,34,35,36,37], focused ion beam helium ion microscope (FIBHIM) [38,39], atomic force microscope (AFM) [40,41,42,43,44,45,46], X-ray computed tomography (X-CT) [12,47,48,49,50,51,52,53,54]; the quantitative methods can be divided into mercury injection capillary pressure (MICP) [5,55,56,57], constant-rate mercury injection (CRMI) [58,59,60,61], nuclear magnetic resonance (NMR) [59,62,63,64,65,66,67,68,69], small angle X-ray scattering (SAXS), small-angle neutron scattering (SANS) [30,70,71,72,73,74], and gas adsorption techniques such as CO2 gas adsorption and N2 gas adsorption (CO2GA, N2GA) [17,75,76].
Researchers have used various methods to characterize sediment pores in different reservoirs (Figure 2). This paper summarizes the mainstream viewpoints on reservoir pores delineation in recent years and compiles them through the comprehensive application of the listed methods; it also briefly discusses the development trend and key points.
This paper summarizes the innovative applications of various techniques for reservoir characterization, focusing on data correction, image processing, and reservoir fluid discrimination. In addition, although there are many research works on fractals, there are fewer systematic discussions; this paper summarizes the principles and formulas of fractals of various characterization methods, which can be used as a reference for future characterization studies.

2. Materials and Methods Multi-Scale Characterization of Pores of Unconventional Reservoirs

2.1. Methods for Quantitative Characterization

2.1.1. Nuclear Magnetic Resonance (NMR)

In recent decades, NMR methods have been widely used in geological research, particularly for the characterization of pore structures, evaluation of wettability, fluid discrimination and prediction of porosity and permeability [77].
The T 2 of the fluid can be expressed by:
1 T 2 = 1 T 2 b + 1 T 2 s + 1 T 2 d
where T 2 is the relaxation time of the hydrogen-containing fluid in the pore; T 2 b is the bulk relaxation time; T 2 s is the surface relaxation time; T 2 d is the diffusion relaxation time [78]. T 2 b is higher for distilled water, so that the term 1/ T 2 b can be neglected, and diffusion relaxation can be reduced when a low and uniform magnetic field and shorter pulse intervals are used. Hence, for the evaluation of the pore sizes, the T 2 can be simplified as:
1 T 2 1 T 2 S = ρ S V p o r e
where ρ is the surface relaxivity, S V p o r e is the surface-to-volume ratio of pores.
Since S V p o r e of the pore is a function of the pore radius ( r ) and the pore geometry morphologic factor ( F s ), the T 2 can also be expressed as:
1 T 2 = ρ F s r
The constant value F s is influenced by the geometry of the pores, with 1, 2 and 3 corresponding to slit, cylindrical and spherical pores, respectively [79].
The T 2 curve distribution, which is derived through mathematical fitting of the NMR signal, provides an indirect representation of the sample’s pore size and distribution.
Bound water and bulk water represent the two categories of enriched water in the pore; NMR can be utilized to identify the borders between these two liquid types, enhancing the estimation of pore permeability. There are differences in the relaxation times of the free state, pore-confined, and adsorbed water, and a cut-off value (T2cutoff) is generally used to distinguish the different states [59].
The T2cutoff value (Figure 3) can separate bound water from bulk water; the T2cutoff value obtained from centrifugal tests divides the T2 spectrum into two parts, bound water pore volume is represented by the sum of T2 distributions smaller than the T2cutoff value and bulk water pore volume is represented by the sum of T2 distributions larger than the T2cutoff value [59,80].
The Coates model (Figure 4) is a frequently used model for permeability estimation. It can be used to estimate formations that include hydrocarbons and water:
K = C 2 F F I B V I 2
where is porosity; K is permeability; the constant C reflects the correlation between the pore throat and pore size as a function of pore geometry.
Different hydrogen fractions in complex pore structures can be more accurately identified by 2D NMR T1T2 correlation maps and T1T2 ratios (Figure 5). Furthermore, these identify a variety of fluids in different pore types, such as organic pores (OP) and inorganic pores (IP), including inter- and intra-particle pores, as well as detecting the geochemistry of organic pores.
The PSD of NMR can be calibrated by N2GA and SEM. To obtain a quantitative relationship between T2 and PSD, the two peaks of the T2 spectrum are combined with N2GA (microporous PSD) and SEM (mesoporous PSD), respectively. From this, linear and power index fitting equations were established to derive a linear model and a power index model to calibrate the PSD (Figure 6).

2.1.2. Nuclear Magnetic Resonance Cryoporometry

Nuclear magnetic resonance cryoporometry (NMRC) is an emerging method to analyze the PSD of nanoscale porous materials. The Gibbs–Thompson equation, which forms the basis of NMRC theory, determines the relationship between the melting temperature ( T m ) of a solid and pore size (x):
T m = T m T m x = K G T x
where T m is the melting temperature of the solid, T m is the decrease in melting temperature, T m x is the melting temperature of a crystal with pore diameter x, KGT is the Gibbs–Thompson constant [81,82].
PSD is the pore volume differential coefficient of pore size (x), and is expressed as dV/dx. Since KGT is an experimental empirical value, the increase in pore diameter x corresponds to T m .
The pore volume can be detected through NMRC by using the following expression:
d V d x d I d x = K G T X 2 × d I d T m x
The choice of probe material to characterize a geological sample can have a significant impact on the accuracy of the measurements. Probe materials with varying KGT values have a direct effect on the accuracy of pore size analysis at the same temperature control; the main probe materials currently used in the NMRC characterization technique include water, OMTCS, cyclohexane, and calcium chloride hexahydrate (Table 4).

2.1.3. Mercury Intrusion Capillary Pressure

The basic principle of MICP is based on the non-wetting property of mercury towards most solid interfaces.
Mercury enters the capillary without wetting the surface and overcomes the capillary pressure while expanding in the pores and throats, where the capillary pressure P c can be expressed as:
P c = 2 σ cos θ R
where R is pore radius, σ is interfacial tension, θ is static contact angle. The pore or throat radius distribution curve can be calculated by the equation, while the capillary pressure curve is generated from the entering mercury volume and the corresponding pressure obtained in the mercury pressure experiment [87].
This method has both advantages and limitations. One of the advantages is the wide range of experimental pressures, as shown in Equation (7), where different pressure values correspond to pores of a certain pore size; secondly, experimental capillary pressure curves for mercury rejection can be obtained to analyze the samples’ surface wettability; lastly, batch experiments can be easily conducted owing to the speed of the experiment.
The following are the limitations: (1) For samples with low porosity and low permeability, the resulting high pressure can create artificial cracks that introduce inaccuracies, which are especially noticeable in block samples; (2) The Washburn equation assumes that the sample pores are smooth cylindrical connected pores, but the pores of unconventional reservoirs have complex morphology and rough pore surfaces, which will cause errors in actual measurements; (3) R in Equation (7) is the maximum pore diameter to be measured, the existence of pore throats causes the measured PSD to deviate from the true value.
Nowadays, a good match between MICP-N2GA integrated porosity and helium porosity of the same sample validates some new data calibration methods, and an adequate correlation between the PSD curves recorded by these methods in their overlapping pore size range [5].

2.1.4. Constant-Rate Mercury Injection

Yuan and Swanson were the first to perform the RCMI experiments on the APEX (pore inspection equipment) pore analyzer.
The RCMI method injects mercury into pores at a very low velocity, obtains information about pore structure based on fluctuation of pressure controlled by the throat, and sequentially goes from one throat to another, thus distinguishing the pores and throats within the samples. The results can be visualized and quantitatively analyzed for parameters such as pore size, PSD, pore-throat radius ratio and throat channels.
As the mercury enters the main throat 1, the pressure gradually increases (Figure 7). Once it breaks through the throat, there is a sudden drop in pressure, known as the first pressure drop O(1); the mercury then fills the first hole gradually and moves on to the next throat, causing a second secondary pressure drop O(2); this process continues until all the holes controlled by the main throat are filled, and the pressure reaches the value of the main throat, completing the unit. The radius of the main throat is defined by the pressure at the breakthrough point, and the size of the orifice is determined by the volume of mercury feed. Hence, the size and number of throats can be distinctly observed in the typical intrusion and extrusion curve of RCMI.
Figure 6 depicts the experimental procedure. Initially, mercury enters throat I; once the pressure reaches a specific threshold, mercury breaks through the throat and enters pore 1, and the pressure drops; pore 1 then fills and the pressure rises. Mercury next breaks through throat II and enters pore 2, causing another drop in pressure; this cycle remains until the experiment ends. Compared with MICP, RCMI has a significantly lower maximum inlet pressure, leading to a higher minimum throat radius [58].
The MICP method is generally based on the capillary bundle model, which assumes that the porous media is composed of uniform capillary bundles of varying diameters. This method only provides information about the volume of pores controlled by the throat and is unable to measure the number of throats directly. In contrast, the RCMI method assumes that the porous media consists of throats and pores with different diameters, this method can simultaneously obtain information about both the pores and the throats (Figure 8). It is particularly suitable for low permeability and ultra-low permeability reservoirs with ranging properties of pore and throat. The completion of RCMI experiments usually takes 2–3 days due to the slow quasi-static mercury feed process. The contact angle θ closely approximates the static contact angle, thus in the experimentally obtained throat radius nearly equals the actual throat radius. Thereby, the RCMI method has the advantage of quantitatively characterizing the microscopic throat structure.

2.1.5. N2 Gas Adsorption

Based on the physical adsorption principle of gas and capillary condensation theory, the N2GA method can measure the pore volume (PV), specific surface area (SSA), pore structure, pore shape, and PSD of porous media. The PSD is usually calculated using the BJH (Barrett-Joiner-Halenda) model [88], which points out that the assumed hypotheses are the same as in the MICP technique: a bundle of capillary tubes (cylindrical) connected to the sample’s borders. For SSA calculations, the BET (Brunauer-Emmett-Teller) method is generally used [1,16,75,89].
According to the principle, the sample is placed in a nitrogen-helium gas mixture environment at liquid nitrogen temperature (77 K), and part of the nitrogen molecules are adsorbed on the samples’ pore surface. As the relative pressure increases, the thickness of the nitrogen adsorption layer rises until the pressure matches the corresponding pressure within the pores, then capillary condensation occurs, allowing the samples’ adsorption-desorption isotherms and nitrogen adsorption capacity at various pressures to be determined. When relative pressures (P/P0) are close to 1.0, coalescence occurs on all surfaces [9].
Adsorption or desorption isotherms are obtained by measuring the amount of adsorbed nitrogen and the equilibrium pressure. These are then divided into five categories (Figure 9: from I to V).
Types A, B, C and D hysteresis curves correspond to cylindrical pores, slit-like pores, t wedge-shaped pores, and bottle neck pores, respectively.
The advantages and limitations of the above quantitative characterization techniques are summarized in Table 5.

2.2. Qualitative Characterization Methods

2.2.1. Field Emission Scanning Electron Microscopy

FESEM’s ultra-high resolution of 0.5–2 nm enables image processing for surface morphology analysis of samples. With its X-ray spectrometer, the FESEM is one of the most useful tools for observing morphology and analyzing micro- and nanoscale pore structure (Figure 10). It can assess micro-area elements on sample surfaces both qualitatively and quantitatively; further, it can analyze the chemical composition and morphology of the samples.

2.2.2. Laser Scanning Confocal Microscopy

Laser scanning confocal microscopy (LSCM) utilizes a laser spot as a fluorescence excitation light for continuous scanning of samples, blocking the out-of-focus plane light through spatially conjugated diaphragms (pinholes) and imaging [91]. LSCM offers a high level of detail, with the ability to magnify up to 10,000 times and achieve an ultimate resolution of 0.15 μm. By employing LSCM, it is possible to rebuild 3D pictures and achieve both 3D imaging and structural reconstruction of the pore throat [92] (Figure 11).

2.2.3. Focused Ion Beam–Scanning Electron Microscopy

Seliger et al. pioneered the development of the world’s inaugural focused ion beam (FIB) system in 1978 at Hughes Research Labs (Malibu, CA, USA), utilizing liquid metal gallium (Ga) ions as the emission source. FEI subsequently manufactured the initial focused ion beam system in 1982 and the first electrostatic field-focused electron in 1988. Subsequently, the initial FIB-SEM system was effectively created by integrating a standard scanning electron microscope with the FIB system, positioning the ion beam at a specific angle relative to the electron beam. The FIB-SEM imaging principle is similar to that of SEM, as both use the detector to capture the excitation of the secondary electron imaging; the main difference is that FIB uses an ion beam as the irradiation source, which possesses greater power and mass compared to electrons.
FIB-SEM can analyze and simulate microscopic pores and perform nanoscale 3D reconstruction of unconventional reservoirs to gain insight into the microscopic pore structure inside the reservoir (Figure 12). Specifically, it can characterize the structure, orientation, grain morphology, size, distribution and other information of samples.

2.2.4. Transmission Electron Microscopy

Roska invented the transmission electron microscope (TEM) in 1932, which uses electron beams as the light source. The principle of TEM is to project an electron beam onto a very thin sample. When the electrons collide with the atoms in the sample, the trajectory of the electrons changes, resulting in solid angle scattering. The magnitude of the scattering angle is associated with the density and thickness of the sample; this relationship allows for the formation of bright and dark pictures, which may be viewed on imaging devices after being magnified and focused (Figure 13).
The resolution of TEM surpasses that of OM, achieving a range of 0.1~0.2 nm with magnifications ranging from tens of thousands to millions of times. Therefore, TEM can be used to observe the intricate morphology of samples, including the arrangement of just one column of atoms.
TEM contains three levels of lenses. The lenses include a focusing lens, an objective lens, and a projection lens. The focusing lens is used to mold the initial electron beam, while the objective lens is utilized to focus the electron beam as it traverses the sample, ensuring it passes through the samples. The magnification of TEM is determined by the ratio of the image plane distance of the sample to the objective lens. Cryo-microscopy typically involves integrating a sample freezing device into a conventional transmission electron microscope to cool the sample to liquid nitrogen temperature (77 K). This process minimizes electron beam damage and sample distortion, allowing for a more accurate representation of the sample’s shape.

2.2.5. Atomic Force Microscopy

Atomic force microscopy (AFM) is employed for studying the surface structure of porous media. By detecting the interatomic contact force between the sample surface and the micro force-sensitive components, the surface structure and characteristics of porous media are examined. The force modifies the state of the micro-cantilever, which is fixed at one end and has a micro-tip near the other end that interacts with the samples. With nanoscale resolution, the sensor will identify the change and determine the force distribution, providing information on the surface morphological structure and surface roughness [94,95,96,97].
Recently, AFM has been widely used for on the situ imaging in nanoscale and direct measurement of oil–rock interaction [98,99]. By using quantitative force curve data, it can be utilized to confirm the in situ wettability of the reservoirs and validate the adsorption of nanoparticles on the solid surfaces [99,100,101,102]. By measuring the thickness of the oil–water film directly on the sample surface, the force between the sample and the liquid film, and the roughness of the solid surface, an AFM probe may investigate changes in the wettability of the solid surface. Deng et al. [103] evaluated the wettability of reservoir rock mineral particles based on AFM; the study utilized AFM to measure micrometer-sized water droplets on the surface of quartz minerals of sandstone reservoir to calculate the water–quartz contact angle. Most studies evaluate the wettability-changing ability of wetting agents based on macroscopic phenomena such as contact angle. Three types of operating modes are distinguished according to how the tip interacts with the sample (Table 6).
The advantages are that the sample does not require pretreatment and AFM can directly image the surface morphology of nanoscale pores or pore throats. During the experiment, it is not easy to damage the sample and it can provide three-dimensional images of the surface. The main disadvantages include the image ranging limitation, its slow imaging speed and the probe’s substantial impact on the measurement results.
In previous studies, shale samples from the Longmaxi Formation in the Sichuan Basin were selected for AFM and N2GA experiments. The study proposes a new method to quantify shale porosity and a dual-threshold discrete integration method to estimate the pore contribution of major materials [44].
The advantages and limitations of the above qualitative characterization techniques are summarized in Table 7.

2.3. X-ray Radiation Method

2.3.1. Small Angle X-ray Scattering

The development of the SAXS technique is based on the idea that a porous medium consists of two phases (pore matrix), which has been consistently investigated by Guinier, Fournet and Porod. X-ray scattering is caused by the difference in electron density between the pore space and the reservoir rock matrix. The parameters of the pore structure in an ideal two-phase system are ascertained as follows:
I q = 4 π · ρ m ρ p 2 · φ · 1 φ · V · 0 r 2 γ 0 r sin q r q r d r
R g = 3 5 · R
q = 4 π sin θ λ
D = 2 π q
where γ 0 : normal correlation function of the reservoir sample; I q : scattering intensity; q: scattering vector; ρ m ρ p 2 : electron density difference between the sample matrix ( ρ m ) and the air in the pores ( ρ p ), proportional to the pore density); φ: volume fraction of the pores in the sample (total porosity); V: volume of the sample; R: pore radius; Rg: radius of gyration of the pores; λ: the wavelength of the X-rays; 2θ: scattering angle; D: pore diameter.
Based on Porod’s theory, the scattering intensity of SAXS is given by the following equation:
I q = 2 π I e · ρ m ρ p 2 · S q 4
lim q 4 · I q = K
where K: Porod constant; S: surface area; Ie: scattering intensity of electrons. When there are electron density fluctuations at the pore boundary, the Porod curve shows a positive deviation; when there is a fuzzy phase boundary in the pore, the Porod curve shows a negative deviation.
Q = 0 I · q 2 d q = I e · V · 2 π 2 φ · 1 φ · ρ m ρ p 2
Based on Equations (12) and (14), the SSA can be expressed as:
where Sv: specific surface area; Q: an integral invariant. The 2D scattering images can be converted to 1D scattering data by FIT2D software V12. Equation (13) determines the I(q)–q curve, and then the structural information such as Rg, pore size, fractal dimension, and thickness of the interfacial layer are calculated by post-processing.

2.3.2. X-ray Computed Tomography

Because of the considerable non-homogeneity of reservoir pores, it is important to precisely figure out how pores and fractures are developing to formulate a development program that would maximize returns from exploitation. In the 1980s, medical CT scanning technology was brought to the field of core analysis; as science and technology advanced and research methods were refined, this technology became widely applied in a variety of fields, including saturation and porosity measurement, core inhomogeneity characterization, and more.
The basis of CT imaging is the concept that CT involves using an X-ray beam to scan a sample via layers of a specific thickness; the X-ray attenuation difference is used to create two-dimensional slices. The computer system uses the “filtered-back projection” algorithm to rebuild the two-dimensional slices after converting the attenuation coefficients into CT numbers, which show up as gray values in the grayscale image. CT is equipped with an X-ray microscope, sample stage and slice reconstruction processing software, enabling non-destructive three-dimensional imaging of samples in their original state, and for determining the PSD, size, and connectivity of nano-submicron pores and throats in the reservoirs.
Equation (15) expresses how the X-ray beam attenuates during the scanning process:
I = I 0 e x p μ D
where I: the residual intensity of the ray after it passes through the sample, is the initial intensity of the ray; μ: ray attenuation coefficient, is related to the sample density and atomic coefficient. The larger the μ, the closer the point; μ represents the distribution of the density of the substance; the attenuation coefficient is usually converted to the CT number. D: thickness of the substance.
Currently, CT analysis methods are mainly divided into image direct observation, CT number analysis, and 3D reconstruction model methods (Table 8).
Table 8. Three Analysis Methods of XCT.
Table 8. Three Analysis Methods of XCT.
MethodImage Direct ObservationCT Number Analysis3D Reconstruction Model Method
principlethe internal structural features of the core can be directly observed;
in the scanned section, there are 256 levels of grayscale, with 0 representing the darkest (all black) and 255 representing the brightest (all white);
lower density material appears as black and higher density material appears as white (Figure 14).
CT values correlate with the attenuation coefficient and indirectly reflect the density of the substance;
for scanned samples, the larger mean CT value indicates a denser, more inhomogeneous substance.
threshold segmentation is performed by recognizing material with different densities and pores according to their respective CT number distribution intervals;
none standardized criteria for setting the threshold.
Figure 14. Porosity calculated from AFM-calculated values and N2GA-converted values, respectively [44].
Figure 14. Porosity calculated from AFM-calculated values and N2GA-converted values, respectively [44].
Jmse 12 00908 g014
In image visualization, the white part of grayscale represents mineral grains or high-density cement, such as silica; the black part represents micropores and microfractures; and the gray part represents the matrix components.
Mimics 21 software is used for processing CT statistics, Avizo 9.2 software performs 3D model reconstruction, and the researcher establishes the thresholds relying on their subjective judgment, which may lead to limited and inaccurate reconstruction outcomes (Figure 14, Figure 15, Figure 16 and Figure 17).
The advantages and limitations of the above X-ray characterization methods are summarized in Table 9:

2.4. Comprehensive Characterization PSD of Reservoir

Comprehending and defining the pore structure of reservoirs is vital for efficient hydrocarbon exploration. Reservoirs exhibit unique pore features, prompting the development of numerous methodologies to evaluate intricate pore systems. Table 7 outlines the benefits and constraints of these techniques; choosing the most suitable approach for characterizing reservoir pores is crucial for accurately evaluating reservoir potential and optimizing extraction procedures. The current challenge and primary focus of research in this field is how to harmonize multiple scales of PSD produced through several approaches onto a unified coordinate scale. Existing studies focus on quantitative characterization by stitching PSDs acquired by various methods (Figure 18, Figure 19, Figure 20, Figure 21, Figure 22 and Figure 23); however, there is still work to be done on the problems of image scale extension and feature point selection for quali–tative characterization methodologies [4,15,16,58,66,72,104,105,106].
The PSDs derived by the two SAXS models (Gaussian distribution model and maxi–mum entropy model) are different. The PSD obtained from the Gaussian distribution model for the range of 0–15 nm and the PSD obtained from the maximum entropy theory for the range of 15–48 nm were similar to the PSDs acquired from the NMRC.
The calculations between pore volume and pore size are different; while NMRC tests open pores, SAXS tests both open and closed pores, resulting in larger pore volumes. During NMRC experience, the interaction between water and the sample itself can cause pore dilatation and mineral dissolution. Specifically, the expansion of clay minerals owing to water interaction may increase the number of micropores. The bigger pore volume is correlated with the larger pore SSA, as shown by the strong positive association between TPV and SSA (Figure 18).
The combined analysis of CTS, SEM and HPMI contributed to a comprehensive study of the unconventional pore system (Figure 19).
Research in imaging is also innovating; for example, Du combines FESEM and XCT to establish three parameter categories, i.e., size, direction and morphology, through which the pore evolution characteristics of unconventional reservoirs at microscale depths can be investigated [109].
Based on the non-homogeneous nature of unconventional reservoirs, fluid flow in nanoscale microscopic pores has also become a major challenge in characterization [13], such as the nature of the initial interface between two miscible fluids and the control by peripheral magnetism on the convection of Casson fluids formed by the internal temperature variations in the microscopic pores [110,111,112,113].

2.5. Method Limitations and Challenges

The characterization of microporous structures in unconventional reservoirs still faces many challenges in both qualitative description and quantitative analysis, which need to be studied in depth.
In terms of qualitative characterization, the current results for nanopores inevitably contain errors. In order to improve the accuracy of characterization, there is an urgent need to develop higher precision techniques. At present, it is difficult for image observation techniques to take into account the needs of high resolution and large field of view. Optical microscopes and scanning electron microscopes can cover a wider field of view but lose resolution, while high-precision techniques suffer from the problems of small observation range, poor representation of the samples, long time-consumption, and high cost. In the 3D reconstruction technique, the threshold segmentation of the image is controlled by human factors, which leads to inaccurate hole throat localization. Therefore, combining the application of machine learning and imaging technology, further updating and improving the pore network modeling method, reducing the influence of human factors, and improving the characterization accuracy to solve the research bottleneck of high resolution and large field of view is an important development direction of unconventional reservoir characterization technology.
In terms of quantitative characterization techniques, the joint characterization of the full pore size distribution is the mainstream direction. The data splicing part of different quantitative experimental results is often contradictory. In future research, we need to analyze and compare the experimental principles and accuracy of various testing techniques in depth, find the number of unified parameters that can connect various techniques in series, and correct the errors effectively. This will help to understand the pore structure of unconventional reservoirs more comprehensively and provide more accurate data support for oil and gas exploration and development.

3. Characterization of Reservoir Inhomogeneity by Fractal Dimension (Df)

Previous studies defined fractals as self-similar objects independent of the level of magnification, characterized by the fractal dimension Df [114]. Since Mandelbrot (1977) proposed fractal theory, it has been widely used in petroleum exploration and development. Fractal theory is used to quantitatively characterize the inhomogeneity of reservoirs, explaining irregular, unstable, and highly complex pore structure features, and helping to relate macroscopic petrophysical parameters (porosity, permeability) to microscopic pore structure (pore diameter, PSD and pore throat connectivity).
Within the same scale range, the smaller Df represents a simpler reservoir pore structure, strong homogeneity, and the formation of the reservoir is conducive to the filling and enrichment of hydrocarbons [16,115,116]. The Df of the reservoir pore structure ranges from 2.0 to 3.0 (not included). A low Df (close to 2.0) indicates that the pore throat structure is regular and the pore surface is smooth; on the contrary, a Df close to 3.0 indicates that the pore throat structure is rough and the pore structure is complex. The change in pore morphology from regular to complex, which lowers permeability and obstructs pore fluid movement, is reflected in the increase in Df. Here are varies factors affecting the Df in Table 10.
According to fractal theory, the number of pores with a radius greater than r, N(>r), is related to the power function of the pore radius as follows:
N > r = r r m a x P r d r r D f
where r m a x : maximum pore radius; P(r): distribution density functions of pore radius.
MICP, N2GA, NMR and NMRC methods can be used to infer the Df of the reservoir.

3.1. Fractal Dimension (Df) of Mercury Intrusion Capillary Pressure

In MICP analysis, assuming that the pores of the reservoir sample are cylindrical, the number of pores with a pore radius greater than r can be expressed by Equation (16). The complex pore throat structure is considered as a series of interconnected irregular capillary networks [59,79,117,118,119,120,121,122,123,124].
Assuming that the pores in the reservoir sample are tubular (tube model), N(>r) can be expressed as:
N > r = V H g π r 2 l r D f
where V H g : cumulative volume of mercury at a specific capillary pressure; l: the length of the capillary.
According to the Yang–Laplace equation, the pressure of mercury injection can be expressed as:
P c = 2 σ cos θ r ,   V H g P c D f 2
The Df relating the mercury saturation S H g to the capillary pressure P c can be expressed by the following equation:
S H g = a P c D f 2
where a is a constant. Slope “ D f 2 ” can be obtained from the plot of log S H g vs. log P c :
log S H g = D f 2 log P c + log a
This equation describes the linear relationship between log S H g and log P c . The Df is calculated by Df = S + 2.
In addition to the tubular model, the ball-and-stick model (l = r) is also widely adopted by many scholars. Considering rminrmax, S v can be described as:
S v = r 3 D f r m i n 3 D f r m a x 3 D f r m i n 3 D f = r r m a c 3 D f
where S v : accumulated pore volume corresponding to the total pore volume less than the pore radius of r, i.e., the saturation of the wetting phase during mercury injection. Thus:
S v = 1 S H g
By the formula
P c = 2 σ cos θ r ,   V H g P c D 2
Equation (24) can be expressed as follows:
S v = P m i n P c 3 D f
The logarithm of both sides of the equation yields:
log S v = D f 3 log P c D f 3 log P m i n
D f = S + 3

3.2. Fractal Dimension (Df) of N2 Gas Adsorption

Qi et al. proposed that the Frenkel–Halsey–Hill equation (FHH model) can be used to calculate D f from N2GA data, which is described by the following equation:
ln V V 0 = A ln ln P P 0 + C
where C is a constant; A is derived from the slope of the ln V V 0 vs. ln P P 0 curve [16,108,109,125,126,127,128].
Figure 21. ln V vs. ln ln P 0 P curves [115].
Figure 21. ln V vs. ln ln P 0 P curves [115].
Jmse 12 00908 g021
The FHH plot of the coal samples is shown in the figure, indicating that there are different fractal regions on the coal surface. Based on the fitting coefficient R greater than 0.95, the “average” Df was used to describe the overall roughness of the coal surface.

3.3. Fractal Dimension (Df) of Nuclear Magnetic Resonance

In the fractal model, the number of pores N(r) with dimensions greater than r is calculated from the total fractal equation above. In NMR measurements, the cumulative porosity V p can be expressed by Equation (32) [19,78,79,129,130,131,132,133,134].
V p = r r m a x 3 D f r m i n r m a x 3 D f 1 r m i n r m a x 3 D f
where r m a x is the maximum pore size; r m i n is the minimum pore size. These two parameters correspond to the maximum and minimum T2 values, respectively.
Equation (28) describes the mass fractal model of the PSD in NMR measurements. The pore size r is directly related to the T 2 value; thus, Equation (29) can be obtained:
V p = T 2 T 2 m a x 3 D f T 2 m i n T 2 m a x 3 D f 1 T 2 m i n T 2 m a x 3 D f
where T 2 m a x and T 2 m i n are the maximum and minimum T2 values. Equation (29) quantifies the relationship between the T2 relaxation time distribution and the Df of the pore system.
Since the minimum value of the T2 ( T 2 m i n ) is small enough compared to the associated measured T2, Equation (30) is obtained:
V p = T 2 T 2 m a x 3 D f
where V p is the accumulative pore volume at T2 values; T 2 m a x is the maximum T2 value in the NMR T2 spectrum.
Df can be calculated from the slope 3 D f :
log V p = 3 D f log T 2 + D f 3 log T 2 m a x
where T 2 m a x is the maximum relaxation time, ms; V p is the cumulative volume.
Equation (31) describes the linear relationship between log V p and log T 2 ; Df can be calculated from the slope “ 3 D f ” of the log V p vs. log T 2 plot.
The research also needs to determine the demarcation line between capillary bound water and movable water—the T2cutoff value, T2c. The part of T2 smaller than T2c or larger than T2c characterizes the fractal features of bound and movable water, respectively.
Figure 22. The double-logarithm coordination showing the relationship between V p and T2 [134].
Figure 22. The double-logarithm coordination showing the relationship between V p and T2 [134].
Jmse 12 00908 g022

3.4. Fractal Dimension (Df) of Nuclear Magnetic Resonance Cryoporometry

Combining the principles of Df calculation with the NMRC results, it is found that NMRC data can also be used for the estimation of Df of unconventional reservoirs:
T m = T m T m x = K G T x
The following expression can be derived:
T m m i n = K G T x m a x
And then can be expressed as:
S v = T m m i n T m 3 D f
where S v is the cumulative volume of the pore with radius less than r; T m is calculated using raw NMR signal intensities at different temperature points. Taking the logarithm of the above equation [85,88].
log S v = D f 3 log T m + 3 D f log T m m i n
Figure 23. Df curves obtained from NMRC [93]. The fractal dimension ranges from 2.8063 to 2.9847 for segment I, and from 1.4181 to 2.3621 for segment II.
Figure 23. Df curves obtained from NMRC [93]. The fractal dimension ranges from 2.8063 to 2.9847 for segment I, and from 1.4181 to 2.3621 for segment II.
Jmse 12 00908 g023

3.5. Relationship between Df and Physical Properties of Reservoirs

Df has no obvious correlation with porosity, but is negatively correlated with permeability. When Df is larger, the pore structure is more complex, making it difficult for fluids to pass through and resulting in a decrease in permeability.
As shown in Figure 24, both Dg-s and Dg-b are negatively correlated with porosity and permeability. In terms of porosity, the coefficient of determination between Dg-b and Dg-s, shows that the porosity is mainly affected by the Df of large pore throats with regular shapes and smooth surfaces, while small pore throats only affect porosity to a certain extent. Regarding permeability aspects, the coefficient of determination between Dg-s, Dg-b and permeability are similar, which means that Df of both large and small pore throats has an impact on the permeability.

4. Conclusions and Future Research

This review outlines the methods and fundamentals of characterizing microporous systems in unconventional reservoirs, summarizes the fractal analysis methods corresponding to the different methods, and sorts out the relationship between the fractals and reservoir macroscopic physical properties (porosity, permeability, etc.) with the reservoir microscopic pore structures (pore structure parameters, pore connectivity, etc.). The conclusions are drawn as follows:
The scale of the pores in unconventional reservoirs varies greatly, ranging from nanometers to microns. When the characterization study of the reservoir focuses on the range of less than 10 nm, it is recommended to combine TEM, FIB-SEM, AFM and SAXS methods. For pores smaller than 100 nm, N2GA, NMRC and FESEM methods can be mainly used, and for the characterization of the range of hundreds of nanometers and micrometers, CTS, OM, MICP, CRMI, and NMR methods can be effectively combined. XCT (threshold segmentation) and CRMI techniques can be used to effectively differentiate between pores and throats, and the connectivity of pores can be combined with the pore connectivity analysis of XCT and the MICP technique; when the study involves the information of minerals and bioclastic debris, the OM and XCT techniques are the most intuitive, and the details can be further investigated by the FIBSEM and FESEM techniques; when obtaining pore diameter characterization while focusing on the information of fluids within the pore, it is recommended to use the NMR (T1, T1T2) and NMRC techniques.
The techniques and ideas for characterizing the pores are constantly advancing. This is mainly demonstrated through improvements reflected in the accuracy of characterization (the integrated approach of FIB-HIM and FIB-SEM extends organic pore imaging and quantitative analysis to less than 10 nm), calibration of the range of PSD (converting T2 spectra of NMR to full-scale PSD by using linear and power exponential models in combination with N2GA and SEM), identification of different hydrogen components in the pore system (T1/T2), and dynamic assessment of unconventional reservoirs (estimation of seepage resistance during water displacement based on pore throat size and distribution characteristics using a capillary bundle model, and modeling to dynamically monitor and evaluate the coproduction behavior of multilayer reservoirs).
The issue of balancing spatial resolution with the field of view remains unresolved. Destructive characterization approaches yield data with some inaccuracies, to ensure the reliability of the results, it is necessary to compare them with those obtained by non-destructive techniques.
The analysis of fractal dimension can effectively identify the quality level of reservoirs. Research on macroscopic petrophysical parameters shows that the saturation of movable fluid is negatively correlated with Df due to the complex pore-throat structure that reduces the permeability and destroys the storage space of free fluid. The mineral composition analysis reveals that the fractal dimension is positively correlated with quartz, negatively correlated with feldspar, and weakly correlated with clay mineral content, which needs to be analyzed specifically for each individual reservoir.
In future qualitative characterization research, the injection of contrast agents can be used as a new approach to study the pore structure and enhance the quality of imaging; moreover, the combination with molecular dynamics simulation and artificial intelligence techniques can effectively improve the spatial resolution of images and explore the adsorption/desorption of gases within the pores. This combination also helps in reducing the computational cost of these processes. Furthermore, the utilization of 4D imaging, which combines three-dimensional visualization with real-time monitoring, allows for the observation of the dynamic alterations in microstructure at the pore-scale level, this capability is of utmost importance for future research endeavors.
Regarding quantitative characterization, there are several emerging research areas that can be explored, such as supercritical CO2-enhanced integrated shale gas recovery, carbon geological sequestration, and advanced underground hydrogen storage; these areas aim to achieve the goals of long-term energy supply and net-zero carbon emissions.

Author Contributions

Conceptualization, H.L. and W.G.; writing—original draft preparation, W.G.; writing—review and editing, W.G., W.C., Z.L. and H.L.; visualization, W.G. and H.L.; supervision, H.L; project administration, H.L; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the DD20230063 and DD20221703 from the China Geological Survey and the Guangdong Major project of Basic and Applied Basic Research (No. 2020B0301030003).

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, H.; Shen, J.; Wang, G.; Li, K.; Fang, X. Experimental study on the effect of high-temperature nitrogen immersion on the nanoscale pore structure of different lithotypes of coal. Energy 2023, 284, 128596. [Google Scholar] [CrossRef]
  2. Xiang, J.; Zhu, Y.; Wang, Y.; Chen, S.; Jiang, Z. Structural deformation and its pore-fracture system response of the Wufeng-Longmaxi shale in the Northeast Chongqing area, using FE-SEM, gas adsorption, and SAXS. J. Pet. Sci. Eng. 2022, 209, 109877. [Google Scholar] [CrossRef]
  3. Xiong, J.; Liu, X.; Liang, L. Experimental study on the pore structure characteristics of the Upper Ordovician Wufeng Formation shale in the southwest portion of the Sichuan Basin, China. J. Nat. Gas Sci. Eng. 2015, 22, 530–539. [Google Scholar] [CrossRef]
  4. Jiang, F.; Chen, D.; Wang, Z.; Xu, Z.; Chen, J.; Liu, L.; Huyan, Y.; Liu, Y. Pore characteristic analysis of a lacustrine shale: A case study in the Ordos Basin, NW China. Mar. Pet. Geol. 2016, 73, 554–571. [Google Scholar] [CrossRef]
  5. Yu, Y.; Luo, X.; Wang, Z.; Cheng, M.; Lei, Y.; Zhang, L.; Yin, J. A new correction method for mercury injection capillary pressure (MICP) to characterize the pore structure of shale. J. Nat. Gas Sci. Eng. 2019, 68, 102896. [Google Scholar] [CrossRef]
  6. Liu, X.; Wang, J.; Ge, L.; Hu, F.; Li, C.; Li, X.; Yu, J.; Xu, H.; Lu, S.; Xue, Q. Pore-scale characterization of tight sandstone in Yanchang Formation Ordos Basin China using micro-CT and SEM imaging from nm- to cm-scale. Fuel 2017, 209, 254–264. [Google Scholar] [CrossRef]
  7. Qiao, J.; Zeng, J.; Jiang, S.; Feng, S.; Feng, X.; Guo, Z.; Teng, J. Heterogeneity of reservoir quality and gas accumulation in tight sandstone reservoirs revealed by pore structure characterization and physical simulation. Fuel 2019, 253, 1300–1316. [Google Scholar] [CrossRef]
  8. Iqbal, O.; Padmanabhan, E.; Mandal, A.; Dvorkin, J. Characterization of geochemical properties and factors controlling the pore structure development of shale gas reservoirs. J. Pet. Sci. Eng. 2021, 206, 109001. [Google Scholar] [CrossRef]
  9. Liu, D.; Qiu, F.; Liu, N.; Cai, Y.; Guo, Y.; Zhao, B.; Qiu, Y. Pore structure characterization and its significance for gas adsorption in coals: A comprehensive review. Unconv. Resour. 2022, 2, 139–157. [Google Scholar] [CrossRef]
  10. Sun, S.; Liu, C.; Ye, Y.; Liu, Y. Pore capillary pressure and saturation of methane hydrate bearing sediments. Acta Oceanol. Sin. 2014, 33, 30–36. [Google Scholar] [CrossRef]
  11. Lai, J.; Wang, G.; Wang, Z.; Chen, J.; Pang, X.; Wang, S.; Zhou, Z.; He, Z.; Qin, Z.; Fan, X. A review on pore structure characterization in tight sandstones. Earth-Sci. Rev. 2018, 177, 436–457. [Google Scholar] [CrossRef]
  12. Liu, Q.; Sun, M.; Sun, X.; Liu, B.; Ostadhassan, M.; Huang, W.; Chen, X.; Pan, Z. Pore network characterization of shale reservoirs through state-of-the-art X-ray computed tomography: A review. Gas Sci. Eng. 2023, 113, 204967. [Google Scholar] [CrossRef]
  13. Bera, A.; Shah, S. A review on modern imaging techniques for characterization of nanoporous unconventional reservoirs: Challenges and prospects. Mar. Pet. Geol. 2021, 133, 105287. [Google Scholar] [CrossRef]
  14. Peng, J.; Han, H.; Xia, Q.; Li, B. Fractal characteristic of microscopic pore structure of tight sandstone reservoirs in Kalpintag Formation in Shuntuoguole area, Tarim Basin. Pet. Res. 2020, 5, 1–17. [Google Scholar] [CrossRef]
  15. Yin, T.; Liu, D.; Cai, Y.; Zhou, Y.; Yao, Y. Size Distribution and Fractal Characteristics of Coal Pores through Nuclear Magnetic Resonance Cryoporometry. Energy Fuels 2017, 31, 7746–7757. [Google Scholar] [CrossRef]
  16. Zhang, J.; Tang, Y.; He, D.; Sun, P.; Zou, X. Full-scale nanopore system and fractal characteristics of clay-rich lacustrine shale combining FE-SEM, nano-CT, gas adsorption and mercury intrusion porosimetry. Appl. Clay Sci. 2020, 196, 105758. [Google Scholar] [CrossRef]
  17. Wang, X.; Wang, M.; Li, J.; Shao, H.; Deng, Z.; Wu, Y. Thermal maturity: The controlling factor of wettability, pore structure, and oil content in the lacustrine Qingshankou shale, Songliao Basin. J. Pet. Sci. Eng. 2022, 215, 110618. [Google Scholar] [CrossRef]
  18. Gao, Z.; Yang, X.; Hu, C.; Wei, L.; Jiang, Z.; Yang, S.; Fan, Y.; Xue, Z.; Yu, H. Characterizing the pore structure of low permeability Eocene Liushagang Formation reservoir rocks from Beibuwan Basin in northern South China Sea. Mar. Pet. Geol. 2019, 99, 107–121. [Google Scholar] [CrossRef]
  19. Zhao, D.; Hou, J.; Sarma, H.; Guo, W.; Liu, Y.; Xie, P.; Dou, L.; Chen, R.; Zhang, Z. Pore throat heterogeneity of different lithofacies and diagenetic effects in gravelly braided river deposits: Implications for understanding the formation process of high-quality reservoirs. Geoenergy Sci. Eng. 2023, 221, 111309. [Google Scholar] [CrossRef]
  20. Liu, A.; Liu, S.; Liu, Y.; Liu, B.; Liu, T. Characterizing mechanical heterogeneity of coal at nano-to-micro scale using combined nanoindentation and FESEM-EDS. Int. J. Coal Geol. 2022, 261, 104081. [Google Scholar] [CrossRef]
  21. Zhang, S.; Yan, J.; Hu, Q.; Wang, J.; Tian, T.; Chao, J.; Wang, M. Integrated NMR and FE-SEM methods for pore structure characterization of Shahejie shale from the Dongying Depression, Bohai Bay Basin. Mar. Pet. Geol. 2019, 100, 85–94. [Google Scholar] [CrossRef]
  22. Li, Z.; Liu, D.; Cai, Y.; Wang, Y.; Teng, J. Adsorption pore structure and its fractal characteristics of coals by N2 adsorption/desorption and FESEM image analyses. Fuel 2019, 257, 116031. [Google Scholar] [CrossRef]
  23. Pan, J.; Peng, C.; Wan, X.; Zheng, D.; Lv, R.; Wang, K. Pore structure characteristics of coal-bearing organic shale in Yuzhou coalfield, China using low pressure N2 adsorption and FESEM methods. J. Pet. Sci. Eng. 2017, 153, 234–243. [Google Scholar] [CrossRef]
  24. Deng, H.; Hu, X.; Li, H.A.; Luo, B.; Wang, W. Improved pore-structure characterization in shale formations with FESEM technique. J. Nat. Gas Sci. Eng. 2016, 35, 309–319. [Google Scholar] [CrossRef]
  25. Bosacka, A.; Zienkiewicz-Strzalka, M.; Derylo-Marczewska, A.; Sliwinska-Bartkowiak, M.; Sterczynska, A.; Sternik, D.; Rotnicki, K. The influence of chemical and thermal modifications of ordered mesoporous carbon on the melting processes of water confined in pores. Microporous Mesoporous Mater. 2023, 351, 112477. [Google Scholar] [CrossRef]
  26. Zhu, H.; Huang, C.; Ju, Y.; Bu, H.; Li, X.; Yang, M.; Chu, Q.; Feng, H.; Qiao, P.; Qi, Y.; et al. Multi-scale multi-dimensional characterization of clay-hosted pore networks of shale using FIBSEM, TEM, and X-ray micro-tomography: Implications for methane storage and migration. Appl. Clay Sci. 2021, 213, 106239. [Google Scholar] [CrossRef]
  27. Zhong, X.L.; Haigh, S.J.; Zhou, X.; Withers, P.J. An in-situ method for protecting internal cracks/pores from ion beam damage and reducing curtaining for TEM sample preparation using FIB. Ultramicroscopy 2020, 219, 113135. [Google Scholar] [CrossRef] [PubMed]
  28. Borrelli, M.; Campilongo, G.; Critelli, S.; Ida, D.P.; Perri, E. 3D nanopores modeling using TEM-tomography (dolostones—Upper Triassic). Mar. Pet. Geol. 2019, 99, 443–452. [Google Scholar] [CrossRef]
  29. Gaboreau, S.; Robinet, J.-C.; Prêt, D. Optimization of pore-network characterization of a compacted clay material by TEM and FIB/SEM imaging. Microporous Mesoporous Mater. 2016, 224, 116–128. [Google Scholar] [CrossRef]
  30. Sun, M.; Zhang, L.; Hu, Q.; Pan, Z.; Yu, B.; Sun, L.; Bai, L.; Fu, H.; Zhang, Y.; Zhang, C.; et al. Multiscale connectivity characterization of marine shales in southern China by fluid intrusion, small-angle neutron scattering (SANS), and FIB-SEM. Mar. Pet. Geol. 2020, 112, 104101. [Google Scholar] [CrossRef]
  31. Fang, H.; Sang, S.; Liu, S.; Du, Y. Methodology of three-dimensional visualization and quantitative characterization of nanopores in coal by using FIB-SEM and its application with anthracite in Qinshui basin. J. Pet. Sci. Eng. 2019, 182, 106285. [Google Scholar] [CrossRef]
  32. Liu, S.; Sang, S.; Wang, G.; Ma, J.; Wang, X.; Wang, W.; Du, Y.; Wang, T. FIB-SEM and X-ray CT characterization of interconnected pores in high-rank coal formed from regional metamorphism. J. Pet. Sci. Eng. 2017, 148, 21–31. [Google Scholar] [CrossRef]
  33. Li, Z.; Liu, D.; Cai, Y.; Ranjith, P.; Yao, Y. Multi-scale quantitative characterization of 3-D pore-fracture networks in bituminous and anthracite coals using FIB-SEM tomography and X-ray μ-CT. Fuel 2017, 209, 43–53. [Google Scholar] [CrossRef]
  34. Devarapalli, R.S.; Islam, A.; Faisal, T.F.; Sassi, M.; Jouiad, M. Micro-CT and FIB–SEM imaging and pore structure characterization of dolomite rock at multiple scales. Arab. J. Geosci. 2017, 10, 361. [Google Scholar] [CrossRef]
  35. Aslannejad, H.; Hassanizadeh, S.M.; Raoof, A.; de Winter, D.; Tomozeiu, N.; van Genuchten, M. Characterizing the hydraulic properties of paper coating layer using FIB-SEM tomography and 3D pore-scale modeling. Chem. Eng. Sci. 2017, 160, 275–280. [Google Scholar] [CrossRef]
  36. Zhou, S.; Yan, G.; Xue, H.; Guo, W.; Li, X. 2D and 3D nanopore characterization of gas shale in Longmaxi formation based on FIB-SEM. Mar. Pet. Geol. 2016, 73, 174–180. [Google Scholar] [CrossRef]
  37. Gu, X.; Cole, D.R.; Rother, G.; Mildner, D.F.R.; Brantley, S.L. Pores in Marcellus Shale: A Neutron Scattering and FIB-SEM Study. Energy Fuels 2015, 29, 1295–1308. [Google Scholar] [CrossRef]
  38. Wu, J.; Luo, C.; Zhong, K.; Li, Y.; Li, G.; Du, Z.; Yang, J. Innovative characterization of organic nanopores in marine shale by the integration of HIM and SEM. Energy 2023, 282, 128390. [Google Scholar] [CrossRef]
  39. Zhang, K.; Song, Y.; Jia, C.; Jiang, Z.; Han, F.; Wang, P.; Yuan, X.; Yang, Y.; Zeng, Y.; Li, Y.; et al. Formation mechanism of the sealing capacity of the roof and floor strata of marine organic-rich shale and shale itself, and its influence on the characteristics of shale gas and organic matter pore development. Mar. Pet. Geol. 2022, 140, 105647. [Google Scholar] [CrossRef]
  40. Yan, X.; Dai, C.; Wang, R.; Liu, H.; Meng, S.; Jin, X.; Hu, Y.; Wu, Y. Experimental study on countercurrent imbibition in tight oil reservoirs using nuclear magnetic resonance and AFM: Influence of liquid–liquid/solid interface characteristics. Fuel 2024, 358, 130026. [Google Scholar] [CrossRef]
  41. Zhang, W.; Ning, Z.; Gai, S.; Zhu, J.; Fan, F.; Liu, Z.; Wang, H. Fast and effective observations of the pore structure of tight sandstones at the same location by utilizing AFM and CF-SEM. J. Pet. Sci. Eng. 2022, 208, 109554. [Google Scholar] [CrossRef]
  42. Qiao, P.; Ju, Y.; Yu, K.; Ju, L.; Xiao, L.; Feng, H.; Yao, Y.; Nie, B.; Li, X.; Tian, J.; et al. Nanoscale quantitative characterization of microstructure evolution of partly graphitized high rank coal: Evidence from AFM and HRTEM. Fuel 2022, 324, 124802. [Google Scholar] [CrossRef]
  43. Pipintakos, G.; Hasheminejad, N.; Lommaert, C.; Bocharova, A.; Blom, J. Application of Atomic Force (AFM), Environmental Scanning Electron (ESEM) and Confocal Laser Scanning Microscopy (CLSM) in bitumen: A review of the ageing effect. Micron 2021, 147, 103083. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, S.; Li, X.; Chen, S.; Wang, Y.; Gong, Z.; Zhang, Y. A new application of atomic force microscopy in the characterization of pore structure and pore contribution in shale gas reservoirs. J. Nat. Gas Sci. Eng. 2021, 88, 103802. [Google Scholar] [CrossRef]
  45. Yesufu-Rufai, S.; Marcelis, F.; Georgiadis, A.; Berg, S.; Rucker, M.; van Wunnik, J.; Luckham, P. Atomic Force Microscopy (AFM) study of redox conditions in sandstones: Impact on wettability modification and mineral morphology. Colloids Surf. A Physicochem. Eng. Asp. 2020, 597, 124765. [Google Scholar] [CrossRef]
  46. Li, Y.; Yang, J.; Pan, Z.; Tong, W. Nanoscale pore structure and mechanical property analysis of coal: An insight combining AFM and SEM images. Fuel 2020, 260, 116352. [Google Scholar] [CrossRef]
  47. Yu, Y.; Xu, H.; Bai, Y.; Niu, W.; Tian, L.; Zhang, H. CT-based 3D pore-fracture network analysis of volcanic reservoirs of Lower Cretaceous Yingcheng formation in southern Songliao Basin, China: Impact on natural gas migration. Geoenergy Sci. Eng. 2023, 223, 211581. [Google Scholar] [CrossRef]
  48. Liu, C.; Buono, G.; Pappalardo, L.; Shan, X.; Yi, J.; Shi, Y.; Ventura, G. X-ray computed microtomography revealing the effects of volcanic, alteration, and burial processes on the pore structure of rocks from unconventional reservoirs (Songliao Basin, NE China). Geoenergy Sci. Eng. 2023, 226, 211781. [Google Scholar] [CrossRef]
  49. Li, C.; Tan, M.; Wang, Z.; Li, Y.; Xiao, L. Nuclear magnetic resonance pore radius transformation method and fluid mobility characterization of shale oil reservoirs. Geoenergy Sci. Eng. 2023, 221, 211403. [Google Scholar] [CrossRef]
  50. Xie, L.; You, Q.; Wang, E.; Li, T.; Song, Y. Quantitative characterization of pore size and structural features in ultra-low permeability reservoirs based on X-ray computed tomography. J. Pet. Sci. Eng. 2022, 208, 109733. [Google Scholar] [CrossRef]
  51. Su, Y.; Zha, M.; Jiang, L.; Ding, X.; Qu, J.; Jin, J.; Iglauer, S. Pore structure and fluid distribution of tight sandstone by the combined use of SEM, MICP and X-ray micro-CT. J. Pet. Sci. Eng. 2022, 208, 109241. [Google Scholar] [CrossRef]
  52. Wang, X.; Pan, J.; Wang, K.; Ge, T.; Wei, J.; Wu, W. Characterizing the shape, size, and distribution heterogeneity of pore-fractures in high rank coal based on X-ray CT image analysis and mercury intrusion porosimetry. Fuel 2020, 282, 118754. [Google Scholar] [CrossRef]
  53. Zhao, H.; Zhao, T.; Ning, Z.; Zhang, R.; Duan, T.; Wang, Q.; Lian, P.; Zhang, D.; Zhang, W. Petrophysical characterization of tight oil sandstones by microscale X-ray computed tomography. Mar. Pet. Geol. 2019, 102, 604–614. [Google Scholar] [CrossRef]
  54. Wu, Y.; Tahmasebi, P.; Lin, C.; Zahid, M.A.; Dong, C.; Golab, A.N.; Ren, L. A comprehensive study on geometric, topological and fractal characterizations of pore systems in low-permeability reservoirs based on SEM, MICP, NMR, and X-ray CT experiments. Mar. Pet. Geol. 2019, 103, 12–28. [Google Scholar] [CrossRef]
  55. Liu, K.; Mirzaei-Paiaman, A.; Liu, B.; Ostadhassan, M. A new model to estimate permeability using mercury injection capillary pressure data: Application to carbonate and shale samples. J. Nat. Gas Sci. Eng. 2020, 84, 103691. [Google Scholar] [CrossRef]
  56. Li, P.; Jia, C.; Jin, Z.; Liu, Q.; Bi, H.; Zheng, M.; Wu, S.; Huang, Z. Pore Size Distribution of a Tight Sandstone Reservoir and its Effect on Micro Pore-throat Structure: A Case Study of the Chang 7 Member of the Xin’anbian Block, Ordos Basin, China. Acta Geol. Sin.—Engl. Ed. 2020, 94, 219–232. [Google Scholar] [CrossRef]
  57. Zhang, F.; Jiang, Z.; Sun, W.; Li, Y.; Zhang, X.; Zhu, L.; Wen, M. A multiscale comprehensive study on pore structure of tight sandstone reservoir realized by nuclear magnetic resonance, high pressure mercury injection and constant-rate mercury injection penetration test. Mar. Pet. Geol. 2019, 109, 208–222. [Google Scholar] [CrossRef]
  58. Yang, Y.; Xiao, W.; Zheng, L.; Lei, Q.-H.; Qin, C.-Z.; He, Y.-A.; Liu, S.-S.; Li, M.; Li, Y.-M.; Zhao, J.-Z.; et al. Pore throat structure heterogeneity and its effect on gas-phase seepage capacity in tight sandstone reservoirs: A case study from the Triassic Yanchang Formation, Ordos Basin. Pet. Sci. 2023, 20, 2892–2907. [Google Scholar] [CrossRef]
  59. Wu, Y.; Liu, C.; Ouyang, S.; Luo, B.; Zhao, D.; Sun, W.; Awan, R.S.; Lu, Z.; Li, G.; Zang, Q. Investigation of pore-throat structure and fractal characteristics of tight sandstones using HPMI, CRMI, and NMR methods: A case study of the lower Shihezi Formation in the Sulige area, Ordos Basin. J. Pet. Sci. Eng. 2022, 210, 110053. [Google Scholar] [CrossRef]
  60. Qu, Y.; Sun, W.; Wu, H.; Huang, S.; Li, T.; Ren, D.; Chen, B. Impacts of pore-throat spaces on movable fluid: Implications for understanding the tight oil exploitation process. Mar. Pet. Geol. 2022, 137, 105509. [Google Scholar] [CrossRef]
  61. Gao, H.; Cao, J.; Wang, C.; He, M.; Dou, L.; Huang, X.; Li, T. Comprehensive characterization of pore and throat system for tight sandstone reservoirs and associated permeability determination method using SEM, rate-controlled mercury and high pressure mercury. J. Pet. Sci. Eng. 2019, 174, 514–524. [Google Scholar] [CrossRef]
  62. Yuan, Y.; Rezaee, R.; Zhou, M.-F.; Iglauer, S. A comprehensive review on shale studies with emphasis on nuclear magnetic resonance (NMR) technique. Gas Sci. Eng. 2023, 120, 205163. [Google Scholar] [CrossRef]
  63. Wang, J.; Tian, L.; Wang, Z.; Liu, Z.; Wang, H.; Yang, D.; Chai, X.; Huang, C.; Jiang, L. Performance evaluation of commingled production in a multilayer oil reservoir based on microscopic pore-throat structures. Fuel 2023, 348, 128482. [Google Scholar] [CrossRef]
  64. Yin, N.; Hu, Q.; Becker, S.J.; Jones, R.; Meng, M.; Zhu, X.; Liu, H. Development of an NMR workflow for determining nano-petrophysical properties of marine and lacustrine mudrocks. J. Pet. Sci. Eng. 2022, 214, 110491. [Google Scholar] [CrossRef]
  65. Liu, Z.; Liu, D.; Cai, Y.; Yao, Y.; Pan, Z.; Zhou, Y. Application of nuclear magnetic resonance (NMR) in coalbed methane and shale reservoirs: A review. Int. J. Coal Geol. 2020, 218, 103261. [Google Scholar] [CrossRef]
  66. Zhang, P.; Lu, S.; Li, J. Characterization of pore size distributions of shale oil reservoirs: A case study from Dongying sag, Bohai Bay basin, China. Mar. Pet. Geol. 2019, 100, 297–308. [Google Scholar] [CrossRef]
  67. Zhang, P.; Lu, S.; Li, J.; Chen, C.; Xue, H.; Zhang, J. Petrophysical characterization of oil-bearing shales by low-field nuclear magnetic resonance (NMR). Mar. Pet. Geol. 2018, 89, 775–785. [Google Scholar] [CrossRef]
  68. Li, J.; Huang, W.; Lu, S.; Wang, M.; Chen, G.; Tian, W.; Guo, Z. Nuclear Magnetic Resonance T1–T2 Map Division Method for Hydrogen-Bearing Components in Continental Shale. Energy Fuels 2018, 32, 9043–9054. [Google Scholar] [CrossRef]
  69. Xiao, D.; Jiang, S.; Thul, D.; Huang, W.; Lu, Z.; Lu, S. Combining rate-controlled porosimetry and NMR to probe full-range pore throat structures and their evolution features in tight sands: A case study in the Songliao Basin, China. Mar. Pet. Geol. 2017, 83, 111–123. [Google Scholar] [CrossRef]
  70. Sun, M.; Zhao, J.; Pan, Z.; Hu, Q.; Yu, B.; Tan, Y.; Sun, L.; Bai, L.; Wu, C.; Blach, T.P.; et al. Pore characterization of shales: A review of small angle scattering technique. J. Nat. Gas Sci. Eng. 2020, 78, 103294. [Google Scholar] [CrossRef]
  71. Sun, M.; Zhang, L.; Hu, Q.; Pan, Z.; Yu, B.; Sun, L.; Bai, L.; Connell, L.D.; Zhang, Y.; Cheng, G. Pore connectivity and water accessibility in Upper Permian transitional shales, southern China. Mar. Pet. Geol. 2019, 107, 407–422. [Google Scholar] [CrossRef]
  72. Sun, M.; Yu, B.; Hu, Q.; Yang, R.; Zhang, Y.; Li, B.; Melnichenko, Y.B.; Cheng, G. Pore structure characterization of organic-rich Niutitang shale from China: Small angle neutron scattering (SANS) study. Int. J. Coal Geol. 2018, 186, 115–125. [Google Scholar] [CrossRef]
  73. Clarkson, C.R.; Freeman, M.; He, L.; Agamalian, M.; Melnichenko, Y.; Mastalerz, M.; Bustin, R.; Radliński, A.; Blach, T. Characterization of tight gas reservoir pore structure using USANS/SANS and gas adsorption analysis. Fuel 2012, 95, 371–385. [Google Scholar] [CrossRef]
  74. Radlinski, A.P.; Mastalerz, M.; Hinde, A.; Hainbuchner, M.; Rauch, H.; Baron, M.; Lin, J.; Fan, L.; Thiyagarajan, P. Application of SAXS and SANS in evaluation of porosity, pore size distribution and surface area of coal. Int. J. Coal Geol. 2004, 59, 245–271. [Google Scholar] [CrossRef]
  75. Wang, X.; Geng, J.; Zhang, D.; Xiao, W.; Chen, Y.; Zhang, H. Influence of sub-supercritical CO2 on pore structure and fractal characteristics of anthracite: An experimental study. Energy 2022, 261, 125115. [Google Scholar] [CrossRef]
  76. Zhang, M.; Fu, X. Characterization of pore structure and its impact on methane adsorption capacity for semi-anthracite in Shizhuangnan Block, Qinshui Basin. J. Nat. Gas Sci. Eng. 2018, 60, 49–62. [Google Scholar] [CrossRef]
  77. Ge, X.; Fan, Y.; Liu, J.; Zhao, J.; Zeng, B.; Xing, D. Numerical investigating the low field NMR response of representative pores at different pulse sequence parameters. Comput. Geosci. 2021, 151, 104761. [Google Scholar] [CrossRef]
  78. Wu, F.; Li, Y.; Burnham, B.; Zhang, Z.; Yao, C.; Yuan, L.; Zhang, F.; Deng, H.; Xi, Y.; He, J. Fractal-based NMR permeability estimation in tight sandstone: A case study of the Jurassic rocks in the Sichuan Basin, China. J. Pet. Sci. Eng. 2022, 218, 110940. [Google Scholar] [CrossRef]
  79. Wang, F.; Wang, L. Pore structure analysis and permeability prediction of shale oil reservoirs with HPMI and NMR: A case study of the Permian Lucaogou Formation in the Jimsar Sag, Junggar Basin, NW China. J. Pet. Sci. Eng. 2022, 214, 110503. [Google Scholar] [CrossRef]
  80. Krzyzak, A.T.; Habina-Skrzyniarz, I.; Machowski, G.; Mazur, W. Overcoming the barriers to the exploration of nanoporous shales porosity. Microporous Mesoporous Mater. 2020, 298, 110003. [Google Scholar] [CrossRef]
  81. Qin, Y.; Yao, S.; Xiao, H.; Cao, J.; Hu, W.; Sun, L.; Tao, K.; Liu, X. Pore structure and connectivity of tight sandstone reservoirs in petroleum basins: A review and application of new methodologies to the Late Triassic Ordos Basin, China. Mar. Pet. Geol. 2021, 129, 105084. [Google Scholar] [CrossRef]
  82. Fleury, M.; Chevalier, T.; Jorand, R.; Jolivet, I.; Nicot, B. Oil-water pore occupancy in the Vaca Muerta source-rocks by NMR cryoporometry. Microporous Mesoporous Mater. 2021, 311, 110680. [Google Scholar] [CrossRef]
  83. Webber, J.B.W.; Strange, J.H.; Dore, J.C. An evaluation of NMR cryoporometry, density measurement and neutron scattering methods of pore characterisation. Magn. Reson. Imaging 2001, 19, 395–399. [Google Scholar] [CrossRef] [PubMed]
  84. Stokes, R.H.; Tomlins, R.P. Thermodynamic functions of melting for cyclohexane. J. Chem. Thermodyn. 1974, 6, 379–386. [Google Scholar] [CrossRef]
  85. Zhu, F.; Hu, W.; Cao, J.; Sun, F.; Liu, Y.; Sun, Z. Micro/nanoscale pore structure and fractal characteristics of tight gas sandstone: A case study from the Yuanba area, northeast Sichuan Basin, China. Mar. Pet. Geol. 2018, 98, 116–132. [Google Scholar] [CrossRef]
  86. Zhu, F.; Hu, W.; Cao, J.; Liu, B.; Liu, Y.; Chang, C. Probe material choice for nuclear magnetic resonance cryoporometry (NMRC) measurements of the nano-scale pore size distribution of unconventional reservoirs. Energy Explor. Exploit. 2018, 37, 412–428. [Google Scholar] [CrossRef]
  87. Wang, M.; Yu, Q. Pore structure characterization of Carboniferous shales from the eastern Qaidam Basin, China: Combining helium expansion with low-pressure adsorption and mercury intrusion. J. Pet. Sci. Eng. 2017, 152, 91–103. [Google Scholar] [CrossRef]
  88. Barrett, E.P.; Joyner, L.G.; Halenda, P.P. The determination of pore volume and area distributions in porous sub stances. I. Comput. Nitrogen Isotherms. J. Am. Chem. Soc. 1951, 73, 373–380. [Google Scholar] [CrossRef]
  89. Wang, Z.; Hao, C.; Wang, X.; Wang, G.; Ni, G.; Cheng, Y. Effects of micro-mesopore structure characteristics on methane adsorption capacity of medium rank coal. Fuel 2023, 351, 128910. [Google Scholar] [CrossRef]
  90. Guo, X.; Huang, Z.; Zhao, L.; Han, W.; Ding, C.; Sun, X.; Yan, R.; Zhang, T.; Yang, X.; Wang, R. Pore structure and multi-fractal analysis of tight sandstone using MIP, NMR and NMRC methods: A case study from the Kuqa depression, China. J. Pet. Sci. Eng. 2019, 178, 544–558. [Google Scholar] [CrossRef]
  91. Gao, Z.; Duan, L.; Jiang, Z.; Huang, L.; Chang, J.; Zheng, G.; Wang, Z.; An, F.; Wei, W. Using laser scanning confocal microscopy combined with saturated oil experiment to investigate the pseudo in-situ occurrence mechanism of light and heavy components of shale oil in sub-micron scale. J. Pet. Sci. Eng. 2023, 220, 111234. [Google Scholar] [CrossRef]
  92. Hackley, P.C.; Kus, J.; Filho, J.G.M.; Czaja, A.D.; Borrego, A.G.; Životić, D.; Valentine, B.J.; Hatcherian, J.J. Characterization of bituminite in Kimmeridge Clay by confocal laser scanning and atomic force microscopy. Int. J. Coal Geol. 2022, 251, 103927. [Google Scholar] [CrossRef]
  93. Liu, B.; Yan, M.; Sun, X.; Bai, Y.; Bai, L.; Fu, X. Microscopic and Fractal Characterization of Organic Matter within Lacustrine Shale Reservoirs in the First Member of Cretaceous Qingshankou Formation, Songliao Basin, Northeast China. J. Earth Sci. 2020, 31, 1241–1250. [Google Scholar] [CrossRef]
  94. Hassenkam, T.; Mitchell, A.; Pedersen, C.; Skovbjerg, L.; Bovet, N.; Stipp, S. The low salinity effect observed on sandstone model surfaces. Colloids Surf. A Physicochem. Eng. Asp. 2012, 403, 79–86. [Google Scholar] [CrossRef]
  95. Yang, G.; Chen, T.; Zhao, J.; Yu, D.; Liu, F.; Wang, D.; Fan, M.; Chen, W.; Zhang, J.; Yang, H.; et al. Desorption Mechanism of Asphaltenes in the Presence of Electrolyte and the Extended Derjaguin–Landau–Verwey–Overbeek Theory. Energy Fuels 2015, 29, 4272–4280. [Google Scholar] [CrossRef]
  96. Xie, L.; Cui, X.; Gong, L.; Chen, J.; Zeng, H. Recent Advances in the Quantification and Modulation of Hydrophobic Interactions for Interfacial Applications. Langmuir 2020, 36, 2985–3003. [Google Scholar] [CrossRef] [PubMed]
  97. Afekare, D.; Garno, J.; Rao, D. Enhancing oil recovery using silica nanoparticles: Nanoscale wettability alteration effects and implications for shale oil recovery. J. Pet. Sci. Eng. 2021, 203, 108897. [Google Scholar] [CrossRef]
  98. Tajikmansori, A.; Hosseini, M.; Dehaghani, A.H.S. Mechanistic study to investigate the injection of surfactant assisted smart water in carbonate rocks for enhanced oil recovery: An experimental approach. J. Mol. Liq. 2021, 325, 114648. [Google Scholar] [CrossRef]
  99. Al-Anssari, S.; Wang, S.; Barifcani, A.; Lebedev, M.; Iglauer, S. Effect of temperature and SiO2 nanoparticle size on wettability alteration of oil-wet calcite. Fuel 2017, 206, 34–42. [Google Scholar] [CrossRef]
  100. Afekare, D.; Gupta, I.; Rao, D. Nanoscale investigation of silicon dioxide nanofluids and implications for enhanced oil recovery—An atomic force microscope study. J. Pet. Sci. Eng. 2020, 191, 107165. [Google Scholar] [CrossRef]
  101. Lu, Y.; Liu, D.; Cai, Y.; Gao, C.; Jia, Q.; Zhou, Y. AFM measurement of roughness, adhesive force and wettability in various rank coal samples from Qinshui and Junggar basin, China. Fuel 2022, 317, 123556. [Google Scholar] [CrossRef]
  102. Xie, L.; Wang, J.; Shi, C.; Cui, X.; Huang, J.; Zhang, H.; Liu, Q.; Liu, Q.; Zeng, H. Mapping the Nanoscale Heterogeneity of Surface Hydrophobicity on the Sphalerite Mineral. J. Phys. Chem. C 2017, 121, 5620–5628. [Google Scholar] [CrossRef]
  103. Deng, Y.; Xu, L.; Lu, H.; Wang, H.; Shi, Y. Direct measurement of the contact angle of water droplet on quartz in a reservoir rock with atomic force microscopy. Chem. Eng. Sci. 2018, 177, 445–454. [Google Scholar] [CrossRef]
  104. Wang, F.; Zai, Y. Fractal and multifractal characteristics of shale nanopores. Results Phys. 2021, 25, 104277. [Google Scholar] [CrossRef]
  105. Meng, Z.; Sun, W.; Liu, Y.; Luo, B.; Zhao, M. Effect of pore networks on the properties of movable fluids in tight sandstones from the perspective of multi-techniques. J. Pet. Sci. Eng. 2021, 201, 108449. [Google Scholar] [CrossRef]
  106. Wang, Y.; Liu, L.; Zheng, S.; Luo, Z.; Sheng, Y.; Wang, X. Full-scale pore structure and its controlling factors of the Wufeng-Longmaxi shale, southern Sichuan Basin, China: Implications for pore evolution of highly overmature marine shale. J. Nat. Gas Sci. Eng. 2019, 67, 134–146. [Google Scholar] [CrossRef]
  107. Zhao, Y.; Peng, L.; Liu, S.; Cao, B.; Sun, Y.; Hou, B. Pore structure characterization of shales using synchrotron SAXS and NMR cryoporometry. Mar. Pet. Geol. 2019, 102, 116–125. [Google Scholar] [CrossRef]
  108. Liu, J.; Zhang, C.; Jiang, Y.; Hou, S. Investigation on pore structure characteristics of ultra-tight sandstone reservoirs in the upper Triassic Xujiahe Formation of the northern Sichuan Basin, China. Mar. Pet. Geol. 2022, 138, 105552. [Google Scholar] [CrossRef]
  109. Du, S. Pore characterization of unconventional reservoirs. Nat. Gas Ind. B 2022, 9, 365–375. [Google Scholar] [CrossRef]
  110. Yadav, D.; Awasthi, M.K.; Al-Siyabi, M.; Al-Nadhairi, S.; Al-Rahbi, A.; Al-Subhi, M.; Ragoju, R.; Bhattacharyya, K. Double diffusive convective motion in a reactive porous medium layer saturated by a non-Newtonian Kuvshiniski fluid. Phys. Fluids 2022, 34, 024104. [Google Scholar] [CrossRef]
  111. Kim, M.C.; Yadav, D. Linear and Nonlinear Analyses of the Onset of Buoyancy-Induced Instability in an Unbounded Porous Medium Saturated by Miscible Fluids. Transp. Porous Media 2014, 104, 407–433. [Google Scholar] [CrossRef]
  112. Yadav, D.; Nair, S.B.; Awasthi, M.K.; Ragoju, R.; Bhattacharyya, K. Linear and nonlinear investigations of the impact of chemical reaction on the thermohaline convection in a permeable layer saturated with Casson fluid. Phys. Fluids 2024, 36, 014106. [Google Scholar] [CrossRef]
  113. Yadav, D.; Awasthi, M.K.; Mohamad, A.M.; Ragoju, R.; Bhattacharyya, K.; Hassan, M. The onset of Casson fluid convection in a permeable medium layer produced by purely inner heating with magnetic field. J. Comput. Appl. Mech. 2024; in press. [Google Scholar] [CrossRef]
  114. Zhou, J.; Yang, K.; Xian, X.; Tian, S.; Cai, J. Fractal characteristics of pore structure and its impact on adsorption and flow behaviors in shale. In Modelling of Flow and Transport in Fractal Porous Media; Elsevier: Amsterdam, The Netherlands, 2021; pp. 37–77. [Google Scholar]
  115. Liu, X.; Nie, B. Fractal characteristics of coal samples utilizing image analysis and gas adsorption. Fuel 2016, 182, 314–322. [Google Scholar] [CrossRef]
  116. Li, A.; Ding, W.; He, J.; Dai, P.; Yin, S.; Xie, F. Investigation of pore structure and fractal characteristics of organic-rich shale reservoirs: A case study of Lower Cambrian Qiongzhusi formation in Malong block of eastern Yunnan Province, South China. Mar. Pet. Geol. 2016, 70, 46–57. [Google Scholar] [CrossRef]
  117. Zhou, Y.; Xu, J.; Lan, Y.; Zi, H.; Cui, Y.; Chen, Q.; You, L.; Fan, X.; Wang, G. New insights into pore fractal dimension from mercury injection capillary pressure in tight sandstone. Geoenergy Sci. Eng. 2023, 228, 212059. [Google Scholar] [CrossRef]
  118. Khurpade, P.D.; Kshirsagar, L.K.; Nandi, S. Characterization of heterogeneous petroleum reservoir of Indian Sub-continent: An integrated approach of hydraulic flow unit—Mercury intrusion capillary pressure—Fractal model. J. Pet. Sci. Eng. 2021, 205, 108788. [Google Scholar] [CrossRef]
  119. Qu, Y.; Sun, W.; Tao, R.; Luo, B.; Chen, L.; Ren, D. Pore–throat structure and fractal characteristics of tight sandstones in Yanchang Formation, Ordos Basin. Mar. Pet. Geol. 2020, 120, 104573. [Google Scholar] [CrossRef]
  120. Han, H.; Guo, C.; Zhong, N.-N.; Pang, P.; Gao, Y. A study on fractal characteristics of lacustrine shales of Qingshankou Formation in the Songliao Basin, northeast China using nitrogen adsorption and mercury injection methods. J. Pet. Sci. Eng. 2020, 193, 107378. [Google Scholar] [CrossRef]
  121. Wang, J.; Cao, Y.; Liu, K.; Gao, Y.; Qin, Z. Fractal characteristics of the pore structures of fine-grained, mixed sedimentary rocks from the Jimsar Sag, Junggar Basin: Implications for lacustrine tight oil accumulations. J. Pet. Sci. Eng. 2019, 182, 106363. [Google Scholar] [CrossRef]
  122. Wang, H.; Wu, W.; Chen, T.; Yu, J.; Pan, J. Pore structure and fractal analysis of shale oil reservoirs: A case study of the Paleogene Shahejie Formation in the Dongying Depression, Bohai Bay, China. J. Pet. Sci. Eng. 2019, 177, 711–723. [Google Scholar] [CrossRef]
  123. Zhang, K.; Pang, X.; Zhao, Z.; Shao, X.; Zhang, X.; Li, W.; Wang, K. Pore structure and fractal analysis of Lower Carboniferous carbonate reservoirs in the Marsel area, Chu-Sarysu basin. Mar. Pet. Geol. 2018, 93, 451–467. [Google Scholar] [CrossRef]
  124. Hao, L.; Tang, J.; Wang, Q.; Tao, H.; Ma, X.; Ma, D.; Ji, H. Fractal characteristics of tight sandstone reservoirs: A case from the Upper Triassic Yanchang Formation, Ordos Basin, China. J. Pet. Sci. Eng. 2017, 158, 243–252. [Google Scholar] [CrossRef]
  125. Liu, B.; Nakhaei-Kohani, R.; Bai, L.; Wen, Z.; Gao, Y.; Tian, W.; Yang, L.; Liu, K.; Hemmati-Sarapardeh, A.; Ostadhassan, M. Integrating advanced soft computing techniques with experimental studies for pore structure analysis of Qingshankou shale in Southern Songliao Basin, NE China. Int. J. Coal Geol. 2022, 257, 103998. [Google Scholar] [CrossRef]
  126. Wang, M.; Xue, H.; Tian, S.; Wilkins, R.W.; Wang, Z. Fractal characteristics of Upper Cretaceous lacustrine shale from the Songliao Basin, NE China. Mar. Pet. Geol. 2015, 67, 144–153. [Google Scholar] [CrossRef]
  127. Sun, W.; Feng, Y.; Jiang, C.; Chu, W. Fractal characterization and methane adsorption features of coal particles taken from shallow and deep coalmine layers. Fuel 2015, 155, 7–13. [Google Scholar] [CrossRef]
  128. Yao, Y.; Liu, D.; Tang, D.; Tang, S.; Huang, W. Fractal characterization of adsorption-pores of coals from North China: An investigation on CH4 adsorption capacity of coals. Int. J. Coal Geol. 2008, 73, 27–42. [Google Scholar] [CrossRef]
  129. Zang, Q.; Liu, C.; Awan, R.S.; Yang, X.; Li, G.; Wu, Y.; Lu, Z.; Feng, D. Occurrence characteristics of the movable fluid in heterogeneous sandstone reservoir based on fractal analysis of NMR data: A case study of the Chang 7 Member of Ansai Block, Ordos Basin, China. J. Pet. Sci. Eng. 2022, 214, 110499. [Google Scholar] [CrossRef]
  130. Liu, X.; Jin, Z.; Lai, J.; Fan, X.; Guan, M.; Shu, H.; Wang, G.; Liu, M.; Luo, Y. Fractal behaviors of NMR saturated and centrifugal T2 spectra in oil shale reservoirs: The Paleogene Funing formation in Subei basin, China. Mar. Pet. Geol. 2021, 129, 105069. [Google Scholar] [CrossRef]
  131. Dai, Q.; Wang, G.; Zhao, X.; Han, Z.; Lu, K.; Lai, J.; Wang, S.; Li, D.; Li, Y.; Wu, K. Fractal model for permeability estimation in low-permeable porous media with variable pore sizes and unevenly adsorbed water lay. Mar. Pet. Geol. 2021, 130, 105135. [Google Scholar] [CrossRef]
  132. Zhang, K.; Lai, J.; Bai, G.; Pang, X.; Ma, X.; Qin, Z.; Zhang, X.; Fan, X. Comparison of fractal models using NMR and CT analysis in low permeability sandstones. Mar. Pet. Geol. 2020, 112, 104069. [Google Scholar] [CrossRef]
  133. Wu, B.; Xie, R.; Wang, X.; Wang, T.; Yue, W. Characterization of pore structure of tight sandstone reservoirs based on fractal analysis of NMR echo data. J. Nat. Gas Sci. Eng. 2020, 81, 103483. [Google Scholar] [CrossRef]
  134. Li, A.; Ding, W.; Jiu, K.; Wang, Z.; Wang, R.; He, J. Investigation of the pore structures and fractal characteristics of marine shale reservoirs using NMR experiments and image analyses: A case study of the Lower Cambrian Niutitang Formation in northern Guizhou Province, South China. Mar. Pet. Geol. 2018, 89, 530–540. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of pores, pore throats and microcracks.
Figure 1. Schematic diagram of pores, pore throats and microcracks.
Jmse 12 00908 g001
Figure 2. Techniques for characterizing reservoir pores.
Figure 2. Techniques for characterizing reservoir pores.
Jmse 12 00908 g002
Figure 3. (A) Principle of calculating the T2cutoff; (B) 2D core-NMR model used for fluid identification.
Figure 3. (A) Principle of calculating the T2cutoff; (B) 2D core-NMR model used for fluid identification.
Jmse 12 00908 g003
Figure 4. T2 relaxation distribution of sandstone reservoir samples, and differences between Coates and SDR penetration models.
Figure 4. T2 relaxation distribution of sandstone reservoir samples, and differences between Coates and SDR penetration models.
Jmse 12 00908 g004
Figure 5. Schematic diagram of water saturation model for samples with tree–like pore–throat structure; NMR fluid type map for shale [60,62].
Figure 5. Schematic diagram of water saturation model for samples with tree–like pore–throat structure; NMR fluid type map for shale [60,62].
Jmse 12 00908 g005
Figure 6. Conversion method for NMR T2 spectrum [66]. (A) SEM image; (B) cumulative PSD; (C) pore morphology determined by N2GA; (D,E) conversion method for NMR T2 spectrum.
Figure 6. Conversion method for NMR T2 spectrum [66]. (A) SEM image; (B) cumulative PSD; (C) pore morphology determined by N2GA; (D,E) conversion method for NMR T2 spectrum.
Jmse 12 00908 g006
Figure 7. Schematic diagram of the mercury leading edge breaking through each pore structure with corresponding pressure changes, obtained throat and total pore-throat size distribution [60]. 1–4 represents different pore throats.
Figure 7. Schematic diagram of the mercury leading edge breaking through each pore structure with corresponding pressure changes, obtained throat and total pore-throat size distribution [60]. 1–4 represents different pore throats.
Jmse 12 00908 g007
Figure 8. Comparison of MICP and RCMI.
Figure 8. Comparison of MICP and RCMI.
Jmse 12 00908 g008
Figure 9. Types of pore shapes and N2 adsorption/desorption isotherms.
Figure 9. Types of pore shapes and N2 adsorption/desorption isotherms.
Jmse 12 00908 g009
Figure 10. Image of tight sandstone and shale samples by FESEM [2,90]. (A) intragranular dissolution micro-pores in feldspar minerals; (B) inter-crystalline pores in illite aggregates; (C) microfractures around the rock particles; (D) type II dissolution intraP pores and linear intraP pores in fractured sheets of clay minerals; (E) interP pores between pyrite particles; (F) micro-fractures along the OM particles and clay minerals.
Figure 10. Image of tight sandstone and shale samples by FESEM [2,90]. (A) intragranular dissolution micro-pores in feldspar minerals; (B) inter-crystalline pores in illite aggregates; (C) microfractures around the rock particles; (D) type II dissolution intraP pores and linear intraP pores in fractured sheets of clay minerals; (E) interP pores between pyrite particles; (F) micro-fractures along the OM particles and clay minerals.
Jmse 12 00908 g010
Figure 11. LSCM micrographs of the 3D pore structure of silty and clayed laminae [93]. Green: detrital mineral; red: light component; yellow: laminations; blue: heavy component. (a): detrital mineral; (b): light and heavy component.
Figure 11. LSCM micrographs of the 3D pore structure of silty and clayed laminae [93]. Green: detrital mineral; red: light component; yellow: laminations; blue: heavy component. (a): detrital mineral; (b): light and heavy component.
Jmse 12 00908 g011
Figure 12. FIB-SEM demonstrating the 2D pore fracture structure of a coal sample. (A) Image of sample FIB slice; (B) reconstructed 3D model and extracted pore network by FIB-SEM in coal formed by regional metamorphism. Gray: organic matter; purple: pores; blue: minerals [32,33].
Figure 12. FIB-SEM demonstrating the 2D pore fracture structure of a coal sample. (A) Image of sample FIB slice; (B) reconstructed 3D model and extracted pore network by FIB-SEM in coal formed by regional metamorphism. Gray: organic matter; purple: pores; blue: minerals [32,33].
Jmse 12 00908 g012
Figure 13. (A) TEM micrograph of compacted IDP; (B) segmented image of (A) [29]. (C) Clay-hosted pores; (D) 2D profile of pore topography by Image J in (C) [1].
Figure 13. (A) TEM micrograph of compacted IDP; (B) segmented image of (A) [29]. (C) Clay-hosted pores; (D) 2D profile of pore topography by Image J in (C) [1].
Jmse 12 00908 g013
Figure 15. Range of linear sizes of pores observable by various neutron and X-ray optical instruments. The solid line shows the typical radiation wavelength region used in the experiment [9].
Figure 15. Range of linear sizes of pores observable by various neutron and X-ray optical instruments. The solid line shows the typical radiation wavelength region used in the experiment [9].
Jmse 12 00908 g015
Figure 16. 3D visualizations of heterogeneity in coal samples at different scales: (ac). Typical 2D sections; (df). 3D visualizations [52]. Red shading: micrometer scale; blue shading: macro-scale; green shading: nanometer scale [52].
Figure 16. 3D visualizations of heterogeneity in coal samples at different scales: (ac). Typical 2D sections; (df). 3D visualizations [52]. Red shading: micrometer scale; blue shading: macro-scale; green shading: nanometer scale [52].
Jmse 12 00908 g016
Figure 17. Three-dimensional characteristics of pore network of the sample: (A) individual pores with volume rendering; (B) pore skeleton network; (C) interconnected pores with volume rendering; (D) pore network model of the interconnected pores; 3D pore-throat skeleton structure, throat volume and channel length distribution: (E,F) DS9 in nano-CT scanning; red ball: pores; green tubes: throats [47].
Figure 17. Three-dimensional characteristics of pore network of the sample: (A) individual pores with volume rendering; (B) pore skeleton network; (C) interconnected pores with volume rendering; (D) pore network model of the interconnected pores; 3D pore-throat skeleton structure, throat volume and channel length distribution: (E,F) DS9 in nano-CT scanning; red ball: pores; green tubes: throats [47].
Jmse 12 00908 g017
Figure 18. (A) PV distribution with pore size diameters combining the MICP and CO2 and N2GA methods [16]. (B) Comparing the pore volume proportion of NMRC and N2GA at different pore sizes [15]; (C) pore (throat) size distributions of the investigated shale samples by NMR, SEM, and MICP [66]; (D) PSD from SANS and (N2, CO2) gas adsorption as well as pore-throat sizes from MICP, for shale samples [72]; (E) the PSDs are derived from NMR, MICP, and RCMI of the outlier [105]; (F) PSD results by comparing NMRC and SAXS: (a) comparison of PSD calculated according to the Gaussian distribution model of SAXS and PSD calculated by NMRC; (b) comparison of PSD calcu–lated from SAXS maximum entropy distribution and PSD calculated from NMRC [107].
Figure 18. (A) PV distribution with pore size diameters combining the MICP and CO2 and N2GA methods [16]. (B) Comparing the pore volume proportion of NMRC and N2GA at different pore sizes [15]; (C) pore (throat) size distributions of the investigated shale samples by NMR, SEM, and MICP [66]; (D) PSD from SANS and (N2, CO2) gas adsorption as well as pore-throat sizes from MICP, for shale samples [72]; (E) the PSDs are derived from NMR, MICP, and RCMI of the outlier [105]; (F) PSD results by comparing NMRC and SAXS: (a) comparison of PSD calculated according to the Gaussian distribution model of SAXS and PSD calculated by NMRC; (b) comparison of PSD calcu–lated from SAXS maximum entropy distribution and PSD calculated from NMRC [107].
Jmse 12 00908 g018
Figure 19. The relationship between pore volume and SSA [76].
Figure 19. The relationship between pore volume and SSA [76].
Jmse 12 00908 g019
Figure 20. The relationship between pore type and PSD in Group A reservoirs [108]. (a) Type 3: pores between clay platelets; Type 4: intraplatelet pores within clay aggregates; (b) Type 1 and Type 2: intergranular pores produced by cement and debris dissolution; (c) relationship between pore size distribution and pore types.
Figure 20. The relationship between pore type and PSD in Group A reservoirs [108]. (a) Type 3: pores between clay platelets; Type 4: intraplatelet pores within clay aggregates; (b) Type 1 and Type 2: intergranular pores produced by cement and debris dissolution; (c) relationship between pore size distribution and pore types.
Jmse 12 00908 g020
Figure 24. Relationships between porosity and permeability and Df. Fractal dimensions Dg-s of (A) or (C) (with respect to small pore throats) are from MIP; Fractal dimensions of (B) and (D) Dg-b (>3.0, with respect to larger pore throats) are from NMR.
Figure 24. Relationships between porosity and permeability and Df. Fractal dimensions Dg-s of (A) or (C) (with respect to small pore throats) are from MIP; Fractal dimensions of (B) and (D) Dg-b (>3.0, with respect to larger pore throats) are from NMR.
Jmse 12 00908 g024
Table 1. Types of Petroleum Resources.
Table 1. Types of Petroleum Resources.
Resource TypeDistribution
Characteristics
Accumulation Type
conventional hydrocarbondiscrete-typestructural pool
clustered-typestratigraphic pool
lithologic pool
unconventional hydrocarboncontinuous hydrocarbon
accumulation
tight (sandstone) oil, gas
tight (carbonates) oil, gas
hydrate
shale oil, gas
coal bed methane
Table 2. Classification of Sediment Pore System.
Table 2. Classification of Sediment Pore System.
ClassificationMineral ComponentDevelopmental Genesis Spatial LocationDevelopment ScaleMorphological Characteristics
poresclay mineral poresintercrystalline poreslarge size (>1000 nm)


medium size (100–1000 nm)


small size (10–100 nm)


micro size (2–10 nm)
ultra-micro size (<2 nm)
round, slit, flat, and zigzag
aggregates (stacks)
circumferential pores
clay minerals or aggregates circumferential pores
brittle
mineral pores
circumferential pores
dissolution pores
intragranular pores (non-solviferous genesis)
intergranular pores
fracture pores
other mineral poresother mineral-related pores
microcrackmicrofractures, inter-mineral microfractures, etc., caused by sedimentary tectonic genesissub-micron, micronslit, flat, zigzag
Table 3. Throat Types.
Table 3. Throat Types.
Throat TypesExamplesFeatures
neckingJmse 12 00908 i001pore throats vary greatly in size, and this variability is common in rocks with grain-supported textures, point contacts and contact cementation.
laminarJmse 12 00908 i002pore-throat widths are generally less than 1 mm; pore-throat ratios are moderately high, common in contact cemented, line-contact, and concave-contact rocks.
curved laminarJmse 12 00908 i003
tubeJmse 12 00908 i004the diameter of tube is generally smaller than 0.5 μm;
the ratios of pore to throat is 1:1, which are very common in rocks with matrix support, suture contact and pore cementation.
Table 4. Properties of the Four Different Probe Materials.
Table 4. Properties of the Four Different Probe Materials.
Probe MaterialMelting Point (K)KGT (K·nm)Liquid Density (g/cm3)Enthalpy of Fusion (kJ/mol)Molecular Size (nm)
Water [83]273.157.30.9976.010.4
Cyclohexane [84]279.91480.7792.720.67
Octamethyl cyclotetrasiloxane [85]290.51600.95519.70.9–1.0
CaCl2·6H2O [86]303.15-1.35237.24-
Table 5. Comparison of quantitative characterization techniques.
Table 5. Comparison of quantitative characterization techniques.
MethodAdvantagesLimitations
Low-Field NMRNon-intrusive, does not destroy the sample May be interfered with by magnetic impurities in the sample
Fast measurement for a large number of samplesLimited resolution, may not detect very small pores
Provides pore size and distribution information
NMR Cryoporometry (NMRC)Directly reveals the relationship between melting point and pore volumeSpecial sample preparation and handling required
Suitable for complex pore structuresMay be interfered with by other components in the sample
Provides a relatively complete pore size distribution
Mercury Intrusion Capillary Pressure (MICP)Measures a wide range of pore size distributionsHigh pressure may alter the pore structure
Suitable for large pore measurementsThe pore limit values measured by this method are related to the maximum mercury pressure of the device
Provides pore volume and specific surface area informationCalculated PSD overestimates the volume of large pores to the detriment of tiny pores
Constant-Rate Mercury Injection (CRMI)Distinguishes between pores and throatsLonger experiment time
Provides pore and throat radii and quantitiesLimited maximum mercury intrusion pressure
Obtains three capillary pressure curves
Nitrogen Adsorption
(N2GA)
Suitable for micropore and mesopore measurementsUnderestimate the content of larger mesopores and macropores
Provides specific surface area and pore size distribution informationRequires a longer adsorption equilibration time
Non-intrusive method, does not destroy the sample
Table 6. Three Operation Modes of AFM.
Table 6. Three Operation Modes of AFM.
Operation ModeContact ModeNon-Contact ModeTapping Mode
advantagesyielding stable, high-resolution images.no force applied to the sample surface.
  • eliminates the effect of lateral forces;
  • reduces of forces caused by adsorbed liquid layers;
  • high image resolution;
  • suitable for soft, fragile or sticky samples without damaging its surface.
disadvantages
  • lateral forces may affect image quality;
  • the capillary effect of the adsorbed liquid layer on surface results in high adhesion forces between the needle tip and the sample;
  • the combined force of the lateral force and the adhesion force reduces the spatial resolution of the image, the tip of the needle crossing the sample can damage soft samples.
  • low scanning speed;
  • needle tip separates from the sample, resulting in low lateral resolution;
  • can only be used for hydrophobic samples; the adsorption layer must be very thin.
slower scanning speed than contact mode.
Table 7. Comparison of qualitative characterization methods.
Table 7. Comparison of qualitative characterization methods.
MethodAdvantagesLimitations
Atomic Force Microscope (AFM)Provides true three-dimensional surface imagesLimited field of view, only small areas can be observed
High resolution up to the atomic levelMay be affected by interactions between the probe and the sample
Suitable for conductors and non-conductors
Field Emission Scanning Electron Microscope
(FE-SEM)
High-resolution imagingSpecial sample preparation (e.g., metal coating) required
Direct observation of pore morphology and structurePotential damage to the sample by the electron beam
Suitable for various material types
Transmission Electron Microscope (TEM)High-resolution imaging, observes internal structuresSample needs to be sliced, which may alter the pore structure
Suitable for various material typesLimited field of view, only small areas can be observed
Direct observation of pore morphology and structure
Focused Ion Beam–Scanning Electron Microscope (FIB–SEM)Combines high resolution of SEM with precise cutting of FIBExpensive equipment and operational costs
Enables 3D reconstruction and quantitative analysisComplex sample preparation, may introduce artifacts
Suitable for various material types
Laser Scanning Confocal Microscope (LSCM)Enables 3D imaging and quantitative analysisLower resolution compared to electron microscopes
Suitable for fluorescently labeled samplesMay be affected by fluorescent dyes
Non-intrusive, does not destroy the sample
Table 9. Comparison of radiation methods for characterizing pores in reservoirs.
Table 9. Comparison of radiation methods for characterizing pores in reservoirs.
MethodAdvantagesLimitations
Small-Angle X-ray Scattering (SAXS)Measures a wide range of pore size distributionsComplex data analysis requiring specialized software support
Non-destructive to the sample, suitable for various materialsLimited sensitivity to small pores
Provides pore shape and structure information
X-ray Computed Tomography (XCT) Provides non-destructive testing, maintaining sample integrity.Limited detection capability for large-volume samples using XCT.
High resolution and contrast, clearly showing pore structure.Potential radiation safety concerns, requiring strict operational guidelines.
Applicable to a variety of materials and morphologies.Relatively high equipment and maintenance costs.
Table 10. Factors affecting the fractal dimension Df.
Table 10. Factors affecting the fractal dimension Df.
Influencing FactorFeatures of DfImpact on Reservoir Heterogeneity
Sedimentary environmentRange of Df valuesDifferent sedimentary environments may lead to varying fractal characteristics of the reservoir, such as river, delta, and lake environments, which affect the structure, pore distribution, and connectivity of the reservoir, thereby influencing its heterogeneity.
DiagenesisPattern of Df value distributionDiagenetic processes such as compaction, cementation, and dissolution can affect the petrophysical properties of the reservoir, such as porosity and permeability, thereby affecting the heterogeneity of the reservoir.
Tectonic activityAnomalous regions in Df valuesTectonic activities like folding and faulting can lead to complex fracture and fault systems within the reservoir, which often exhibit higher heterogeneity, manifesting as anomalous values of the Df.
Scale effectScale-dependence of Df valuesAt different observation scales, the heterogeneity of the reservoir may exhibit different characteristics. Larger scales may mask local heterogeneity, while smaller scales may more accurately reveal the heterogeneous structure of the reservoir.
Measurement methodAccuracy of Df value calculationDifferent measurement methods may yield varying fractal dimension values.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guan, W.; Cai, W.; Li, Z.; Lu, H. Microscopic Characterization and Fractal Analysis of Pore Systems for Unconventional Reservoirs. J. Mar. Sci. Eng. 2024, 12, 908. https://doi.org/10.3390/jmse12060908

AMA Style

Guan W, Cai W, Li Z, Lu H. Microscopic Characterization and Fractal Analysis of Pore Systems for Unconventional Reservoirs. Journal of Marine Science and Engineering. 2024; 12(6):908. https://doi.org/10.3390/jmse12060908

Chicago/Turabian Style

Guan, Wen, Wenjiu Cai, Zhenchao Li, and Hailong Lu. 2024. "Microscopic Characterization and Fractal Analysis of Pore Systems for Unconventional Reservoirs" Journal of Marine Science and Engineering 12, no. 6: 908. https://doi.org/10.3390/jmse12060908

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop