Next Article in Journal
Effect of Preparation Conditions of Fe@SiO2 Catalyst on Its Structure Using High-Pressure Activity Studies in a 3D-Printed SS Microreactor
Previous Article in Journal
Novel Intravenous Nanoemulsions Based on Cannabidiol-Enriched Hemp Oil—Development and Validation of an HPLC-DAD Method for Cannabidiol Determination
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational Modeling and Experimental Investigation of CO2-Hydrocarbon System Within Cross-Scale Porous Media

1
University of Chinese Academy of Sciences, Beijing 100049, China
2
Institute of Porous Flow and Fluid Mechanics, Chinese Academy of Sciences, Langfang 065007, China
3
Research Institute of Petroleum Exploration & Development, Beijing 100083, China
4
State Key Laboratory of Enhanced Oil & Gas Recovery, Beijing 100083, China
*
Author to whom correspondence should be addressed.
Submission received: 9 December 2024 / Revised: 8 January 2025 / Accepted: 9 January 2025 / Published: 12 January 2025

Abstract

:
CO2 flooding plays a crucial role in enhancing oil recovery and achieving carbon reduction targets, particularly in unconventional reservoirs with complex pore structures. The phase behavior of CO2 and hydrocarbons at different scales significantly affects oil recovery efficiency, yet its underlying mechanisms remain insufficiently understood. This study improves existing thermodynamic models by introducing Helmholtz free energy as a convergence criterion and incorporating adsorption effects in micro- and nano-scale pores. This study refines existing thermodynamic models by incorporating Helmholtz free energy as a convergence criterion, offering a more accurate representation of confined phase behavior. Unlike conventional Gibbs free energy-based models, this approach effectively accounts for confinement-induced deviations in phase equilibrium, ensuring improved predictive accuracy for nanoscale reservoirs. Additionally, adsorption effects in micro- and nano-scale pores are explicitly integrated to enhance model reliability. A multi-scale thermodynamic model for CO2-hydrocarbon systems is developed and validated through physical simulations. Key findings indicate that as the scale decreases from bulk to 10 nm, the bubble point pressure shows a deviation of 5% to 23%, while the density of confined fluids increases by approximately 2%. The results also reveal that smaller pores restrict gas expansion, leading to an enhanced CO2 solubility effect and stronger phase mixing behavior. Through phase diagram analysis, density expansion, multi-stage contact, and differential separation simulations, we further clarify how confinement influences CO2 injection efficiency. These findings provide new insights into phase behavior changes in confined porous media, improving the accuracy of CO2 flooding predictions. The proposed model offers a more precise framework for evaluating phase transitions in unconventional reservoirs, aiding in the optimization of CO2-based enhanced oil recovery strategies.

1. Introduction

Since the exploration and production of oil and gas, numerous industries have benefited from their development. However, the increasing CO2 emissions associated with the combustion of fossil fuels have exacerbated environmental degradation. In this context, CCS (Carbon Capture and Storage) technology has gradually emerged. However, the high cost of CCS projects under current technological conditions is difficult to bear [1,2]. Therefore, the CCUS (Carbon Capture, Utilization, and Storage) technology, which is based on CO2 reuse to reduce costs, is gradually gaining attention. As an important branch of tertiary oil recovery, the relatively low critical temperature and pressure of CO2 make it a good displacing medium with good miscibility with crude oil. The injected CO2 interacts with the crude oil, resulting in mass transfer diffusion, dispersion, and dissolution phenomena [3]. This process enhances sweeping efficiency and mobilizes crude oil more effectively [4,5]. In addition, CO2 can effectively occupy the pore space in the reservoir and remain in the formation during oil and gas field development, contributing to low-carbon and decarbonization goals. Although CO2 flooding has achieved certain results in improving the recovery rate in low-permeability and unconventional reservoirs, and has also played a role in CO2 sequestration to some extent, the micro-pore structure of reservoirs suitable for CCUS-EOR is complex, with a wide range of pore throat sizes and developed fractures/micro-fractures. The constraints caused by the reduction of pore sizes in porous media may lead to significant changes in physical properties and phase behavior, such as critical pressure and density, viscosity, surface tension, bubble point, dew point, interfacial properties at the fluid-wall interface, and oil–gas interfacial tension [5,6]. This can result in deviations in reservoir scheme design, affecting the actual oil and gas recovery rate and CO2 sequestration rate. For example, the bubble point pressure of the confined fluid is lower than that of the bulk oil, thus delaying the process of light hydrocarbon escaping from the oil phase to the gas phase during production. This can explain the positive discrepancy between actual production data from reservoirs rich in nanoscale storage spaces and conventional simulation results in terms of oil production and oil recovery ratio. Currently, there have been numerous studies on the macroscopic property changes and measures such as adjusting the injection profile through Water Alternating Gas (WAG), which provide a certain research basis [7,8,9]. However, the transformation of phase behavior due to the multi-scale effect still has a significant impact on CO2 displacement efficiency [10,11]. Recent studies have shown that CO2 diffusion in methane-dissolved pore fluids alters phase behavior at multiple scales, influencing CO2 transport and utilization in geological storage and enhanced oil recovery (EOR) processes. These effects need to be systematically analyzed to optimize CO2 injection strategies. Therefore, clarifying the changes in phase behavior of the CO2-reservoir oil system at multiple scales is of great engineering significance for enhancing oil recovery with CO2. Various methods have been employed to study the above-mentioned phenomena, including molecular simulations, thermodynamic models, and physical simulations. However, there is still significant room for improvement in existing methods (as shown in Table 1).
The phase behavior of confined fluids depends on the fluid composition in each phase and is evaluated through phase equilibrium calculations. The phase equilibrium calculation process includes phase stability analysis and flash calculations, both of which depend on the fugacity values of each component. In existing models, the calculation of fugacity is based on the equation of state. Therefore, modifications need to be made to the form or usage of the equation of state, such as adding capillary forces to the fugacity equation or modifying the critical parameters of the equation of state, in order to apply the current calculation framework to nano-reservoirs. Currently, there is no commercial simulator that can predict the phase behavior of confined hydrocarbons in porous media with sufficient accuracy and efficiency. Although many researchers have attempted to conduct large-scale simulations using critical displacement, capillary pressure theory, adsorption models, or their combinations [12,13,14], it is still difficult to explain the nature of the driving forces that lead to different phase behaviors. Recent studies utilizing three-dimensional core reconstruction and numerical simulation of CO2 displacement in tight oil reservoirs have significantly improved our understanding of phase behavior and transport mechanisms at the reservoir scale. Additionally, microscale stress sensitivity in CO2 foam fracturing has been found to influence fluid distribution and displacement efficiency, further impacting the overall phase behavior of CO2-oil systems in tight formations [15].
So far, the theoretical calculation methods for the phase behavior of confined fluids at the nanoscale are mostly based on the calculation of Gibbs free energy. However, under the influence of the high capillary pressure caused by the restriction effect, the assumptions of Gibbs free energy under constant temperature and pressure are no longer applicable, which affects the calculation results. Moreover, Gibbs free energy-based methods often require empirical corrections to account for confinement effects, limiting their applicability to nano-reservoirs. To address these limitations, the Helmholtz free energy framework provides a more fundamental thermodynamic approach that inherently incorporates adsorption effects, molecular interactions, and capillary constraints in confined systems. By directly incorporating these effects into the thermodynamic potential, Helmholtz free energy-based models improve the accuracy and stability of phase equilibrium calculations for CO₂-hydrocarbon systems. This approach eliminates the need for empirical corrections and ensures a more robust theoretical foundation for modeling confined phase behavior. Additionally, the current numerical simulation methods that consider scale effects mostly involve constructing fluid numerical simulation algorithms and do not consider the changes in the flow characteristics of confined fluids. In this study, by introducing the Helmholtz free energy as a convergence criterion, improving the existing thermodynamic model, and considering the influence of adsorption on fluid behavior in micro- and nano-scale pores, a multi-scale thermodynamic calculation model suitable for CO2-hydrocarbon systems is developed to investigate the influence of scale changes on the phase behavior of CO2-hydrocarbon systems. By performing calculations on different block oil data and pressures, the changes in phase behavior parameters during processes such as gas expansion and successive degassing at different scales are simulated, providing a basis for describing the phase behavior of CO2-hydrocarbon systems at multiple scale.
Table 1. Research methods of phase behavior in confined space.
Table 1. Research methods of phase behavior in confined space.
MethodSubtypeLimitationsFindingsReferences
Molecular simulationMolecular Dynamics Simulation (MD)The study of fluid behavior at the rock core scale remains computationally expensive due to the nanosecond-level time step and the presence of “Lyapunov instability.” Additionally, the effect of pore-throat geometry on fluid behavior is not fully capturedMethane exhibits different diffusion behaviors in nanopores with various geometric structures, with throat size determining its self-diffusion ability[16]
The number of adsorbed layers of methane in organic shale nanopores depends on pore size and temperature, where increased temperature weakens methane adsorption[17]
In rough nanopores of shale matrix, methane diffusion primarily occurs as planar diffusion, with significant sensitivity to temperature and pressure for rarefied gases[18]
The phase behavior and composition distribution of hydrocarbon binary mixtures in heterogeneous nanopores are strongly influenced by nanopore confinement, particularly in smaller pores[19]
Monte Carlo (MC)Monte Carlo methods face challenges in accurately capturing complex molecular interactions, especially near the critical region. Grand Canonical Monte Carlo (GCMC) simulations struggle with phase transition predictions and require correction methods for critical point estimations. Additionally, vapor–liquid equilibrium simulations in nanopores demand extensive computational resources and careful validation against experimental dataAdsorption of methane and ethane in organic shale nanopores is highly dependent on pore size, temperature, and pressure, with a preference for adsorption in smaller pores[20]
Competitive adsorption between methane and ethane is observed, with methane showing higher adsorption capacity due to its molecular size and interaction potential[21]
Confinement effects in shale reservoirs result in reduced bubble point pressures and increased dew point pressures compared to bulk conditions[22]
Thermodynamic modelDensity Functional Theory (DFT)DFT simulations often rely on simplified pore structures and assumptions about fluid-wall interactions, which may not fully capture real shale reservoir conditionsConfined fluids exhibit non-uniform density distributions, with higher densities near pore walls[23]
The phase behavior of hydrocarbons in nanopores deviates significantly from bulk conditions, with critical pressure and temperature shifting downward[24]
Competitive adsorption of hydrocarbons and CO₂ in calcite nanopores influences miscibility pressure and phase equilibrium[25]
Thermodynamic Model ModificationThe critical displacement equation correlates with the apparent deviations in nanopores but fails to consider the fluid-wall surface interactions separately. Mesoscopic corrections to macroscopic theories are needed to account for wetting effectsThe Helmholtz free energy of confined fluids is calculated using a van der Waals mean-field model. Phase behavior shifts depend significantly on pore geometry and wetting properties[26]
State equation modificationModified Equation of StateExisting models lack accuracy and efficiency in predicting confined hydrocarbons’ phase behavior. They fail to fully account for adsorption effects, capillary pressure, and fluid–wall interactions, limiting their applicability to shale reservoirsAdsorption alters the phase equilibrium of confined hydrocarbons, shifting critical temperature and pressure[27]
The equation of state is modified to include adsorption effects and capillary pressure in nanopores, improving phase behavior predictions[28]
A pressure correction parameter is defined from a microscopic perspective and correlated with the fluid–fluid potential well depth parameter[29]
A new mixing rule is proposed to extend the configurational energy to mixtures[30]
Physical simulationAdsorption–Desorption MethodTraditional methods often overlook the effect of water saturation on shale pore systems and assume uniform adsorption across all poresAdsorption behavior varies between organic and inorganic pores. Image recognition and simulation improve understanding of shale gas adsorption–desorption mechanisms. Hysteresis effects are observed, revealing new insights into adsorption dynamics[31,32,33]
Differential Scanning Calorimetry (DSC)Measurements are constrained by the complexity of multicomponent systems, potential errors in heat flow calibration, and the influence of pore size, geometry, and chemical compositionPhase transitions of confined fluids are significantly influenced by nanopore confinement, including shifts in bubble and dew points, and altered thermal behaviors[34,35,36,37]
Nanofluidic Control MethodDue to limitations in the observational conditions, a significant amount of experimental work is still required for comprehensive descriptionsThe deviation of the saturation point increases as the depth of the nanochannel decreases[38,39,40]

2. Model Establishment

In this section, a thermodynamic model based on the cubic three-parameter equation of state (PR) is established using the Helmholtz free energy, and relevant fluid phase equilibrium calculations are conducted. This includes solving nonlinear equations and stability analysis based on the Helmholtz free energy, in order to achieve fluid equilibrium calculations at a given temperature and pressure. The specific content includes flash calculation equations, mass conservation boundary conditions, and state equations.
The Helmholtz free energy formulation allows direct incorporation of capillary pressure, adsorption effects, and molecular interactions, which are critical for phase equilibrium calculations in confined porous media. Unlike conventional approaches that introduce capillary pressure corrections as external parameters, this method integrates these effects inherently into the thermodynamic framework. As a result, the phase equilibrium calculations are more robust and consistent across different scales. Additionally, this framework ensures numerical stability by formulating equilibrium conditions based on Helmholtz free energy minimization, avoiding convergence issues associated with traditional fugacity-based models in nano-scale systems.
According to thermodynamic specifications, the Helmholtz free energy of the system is in equilibrium between the gas and liquid phases at a specified temperature T, volume V, and mole numbers of Nc components Ni (i = 1, ……, Nc) in the system.
d A = d A V + d A L + d A V L = 0
In Equation (1), A, AV, AL, and AVL are the Helmholtz free energies of the entire system, gas phase, liquid phase, and gas–liquid transition zone, respectively.
The different parts of Equation (1) can be expanded based on the variation of the Helmholtz free energy:
d A V = S V d T V P V d V V + i = 1 N c G i V ¯ d N i V
d A L = S L d T L P L d V L + i = 1 N c G i L ¯ d N i L
d A V L = S V L d T V L P V L d V V L + σ d a + i = 1 N c G i V L ¯ d N i V L
In Equation (2), S is entropy, P is pressure, σ is interfacial tension, a is interfacial area, and G i ¯ is the partial molar Gibbs free energy of component i.
When the Helmholtz free energy of the system is minimized, material conservation should be satisfied:
d N L i + d N V i + d N V L i = 0 , i = 1 , , N c
In conclusion, the expression for d A can be obtained:
d A = ( P V P L ) d V V ( P V L P L ) d V V L + σ d a + i = 1 N C ( G ¯ i V G ¯ i L ) d N i V + i = 1 N C ( G ¯ i V L G ¯ i L ) d N i V L
Since the volume of the gas–liquid transition zone is very small, the variation of components can be neglected, and the above equation can be simplified as
d A = ( P V P L ) d V V + σ d a + i = 1 N C ( G ¯ i V G ¯ i L ) d N i V
when a system reaches equilibrium, the Helmholtz free energy reaches its minimum value, which is represented as d A = 0 . This can be transformed into the expression for fugacity:
f i L = f i V
x i φ i x i , P L , T P L = y i φ i y i , P V , T P V
where φ i represents the fugacity coefficient of component i, generally expressed as
ln φ i = 1 R T V P n i T , V , n j i j R T V d V ln Z
where φ i is the fugacity coefficient of the corresponding component, R is the gas volume constant, Z is the compressibility factor, T is the system temperature, P is the system pressure, and V is the volume occupied by the system.
Expanding Equation (10) can yield the calculation expression for fugacity coefficient:
φ i = e ln Z B A 2 2 B 2 j = 1 n c A i j c i j A B i B ln Z + 1 2 B Z + 1 + 2 B + B i B Z 1
where
A i j = 1 k i j A i A j 0.5 , A i = 0.45724 1 + m 1 T r 0.5 2 P r i L V T r i 2 , B i = 0.07778 P r i L V T r i
where kij is the binary interaction parameter, Tr is the reduced temperature, and Pr is the reduced pressure. For simplicity of calculation, it is assumed that the total molar quantity N = 1 mol for each calculation. Therefore, in equilibrium calculations, material balance must also be satisfied:
n L x i + n V y i = N i
x i and y i are the mole fractions of the liquid phase and gas phase, respectively, and n L and n V correspond to the molar quantities of the liquid phase and gas phase. When the system reaches equilibrium, the distribution of each component in the two phases should reach equilibrium and satisfy material conservation. Therefore, introducing equilibrium constants, satisfying
x i = c i 1 + n V K i 1
y i = c i K i 1 + n V K i 1 = K i x i
Organizing, we get
i = 1 n c y i x i = 0
f n V = i = 1 n c c i K i 1 1 + n V K i 1 = 0
This set of equations contains 2 N c + 2 unknowns ( N c liquid phase mole fractions x i , N c gas phase mole fractions y i , and the phase ratio of the two phases: n L and n V ). Correspondingly, solving the set of equations involves 2 N c + 2 equations for closure, and all the equations are solved nonlinearly.
In addition, the model considers restricted fluid calculations, so capillary forces and critical parameter corrections need to be added in the calculation process.
In the critical point transition, the critical point calculation method proposed by Zarragoicoechea et al. [41] is used in this study, and the critical point of the mixture is calculated through a mixing rule.
Δ T c = 0.9409 σ L J R p 0.2415 σ L J R p 2
Δ P c = 0.9409 σ L J R p 0.2415 σ L J R p 2
where σ and R p are Lennard–Jones size parameters and pore radius.
The model corrects the capillary force by calculating the Young–Laplace equation and then derives the changes in oil and gas phase pressure and fugacity coefficients. For adsorption phenomena, this paper refers to a description method proposed by Song et al. in 2020 [42], which calculates the adsorption layer thickness by establishing a columnar model using isothermal adsorption. In the calculation process, R p 2 γ is taken as the actual pore throat radius. The thickness of the adsorption layer is given by
γ = m ln R p σ L J + n σ L J R p
where m is the correction factor in the equation of state.
Therefore, the capillary force can be obtained as
P c a p = 2 σ cos θ R p 2 γ
Finally, by formulating appropriate iterative variables to solve the above nonlinear equations, all calculation steps are completed, and a convergence check is performed on the results.

3. Model Validation

To verify the accuracy of the model, the composition data of the B79, B131, and B18 block formations were taken into account, as shown in Table 2. The selected blocks, B79, B131, and B18, represent reservoirs with different lithological and pore structure characteristics. The B79 block primarily consists of tight sandstone with moderate porosity and permeability, while the B131 and B18 blocks contain a mix of carbonate and sandstone formations with varying degrees of heterogeneity. These blocks were chosen due to their representative geological conditions for studying gas injection and phase behavior, covering a range of pore structures, including nano- and micro-scale pores, which are typical in unconventional reservoirs. The optimal grouping method was determined through injection expansion experiments. On this basis, a contact experiment and multiple contact experiments were conducted to investigate the calculation performance of the model in terms of component, pressure, and volume parameters, and to assess the model’s accuracy in predicting phase behavior under varying reservoir conditions. The main gases selected include CO2 and natural gas, while the reservoir fluids consist of the formation oils from the aforementioned three blocks.

3.1. Grouping of Components

In model calculations, the required initial data include temperature, pressure, components, and basic phase parameters of components such as critical properties and binary interaction coefficients. The component parameters can be obtained through oil and gas chromatography data, but if all components are included in the calculation, it may lead to non-convergence or excessive computational burden. Therefore, the best grouping method is selected by comparing the consistency of the results obtained from different component splitting calculations with the injection experiment, which provides a basis for subsequent calculations. The specific experimental steps are as follows.
The full window high-pressure PVT analyzer is cleaned under reservoir temperature conditions and then evacuated. A certain amount of crude oil sample is transferred to the PVT analyzer to maintain a single-phase state at reservoir temperature for 8 h. First, the sample volume is tested under reservoir pressure, and then a certain amount of CO2 is injected into the reservoir crude oil at this pressure to increase the system pressure until all the CO2 is dissolved, at which point the system becomes single-phase. Parameters such as saturation pressure and volume expansion coefficient of the CO2-reservoir crude oil system are tested. Finally, the CO2-reservoir crude oil mixture sample in the PVT analyzer is transferred to the high-temperature and high-pressure electromagnetic viscometer to test the viscosity of the system in a single phase. This completes the first gas expansion experiment. The gas injection amount in the next level of the experiment is higher than that in the previous level (10% higher). This process is repeated, and a total of seven gas expansion experiments are conducted at reservoir temperature.
A large amount of data on different component hydrocarbon systems at different pressures were obtained through experiments, and a total of eight grouping methods were used to group the components (as shown in Table 3). The best grouping method was selected as the seventh group in Figure 1. Based on this optimal result, calculations were performed for the B131 and B18 reservoir oils and compared with experimental results, as shown in Figure 2 (the calculations were conducted using the phase equilibrium equations (Equations (11)–(14)) presented in Section 2).

3.2. Experimental Verification

3.2.1. Experimental Procedure for Single-Contact Mass Transfer Experiment

The specific experimental procedure is shown in Figure 3. By injecting reservoir oil and a sufficient amount of injection gas into the high-temperature and high-pressure PVT apparatus, the pressure is gradually increased above the reservoir oil saturation pressure. After reaching equilibrium, the gas and oil phase components are analyzed by gas chromatography to investigate the changes in the composition of the two phases. The experimental temperature is 90 °C, and the B131 reservoir oil is selected as the liquid phase component, while CO2 is used as the injection gas. The pressure is gradually increased (from 26 MPa to 38 MPa), and the equilibrium oil and gas phase compositions are separately analyzed.
Taking the B131 experiment as an example, experimental photos of the interaction between CO2 and reservoir oil were captured. The morphology of CO2 in the PVT experiment is shown in Figure 4. The gradient of gas phase components with pressure change is shown in Figure 5, with volatile components at 0.0975 mol%/MPa, intermediate components at +0.039 mol%/MPa, and C7+ components at −0.137 mol%/MPa. The gas phase volume gradually increases, the color deepens, and the phase boundary between the two phases becomes blurry. This indicates that the extraction effect is stronger than the dissolution effect in the process of single-contact mass transfer of CO2, which is related to the content of intermediate hydrocarbon components such as C2–C6 in the reservoir oil. Combining the component data in Table 2, it is not difficult to find that the B131 reservoir oil contains a large number of light components and intermediate hydrocarbons.

Calculation Results of Component Parameters

In order to verify the accuracy of the model’s component calculations, the oil and gas phase components in the experiment were calculated at the corresponding experimental pressure. The calculation was based on a gas–liquid substance ratio of 2:1 to 3:1 to ensure that the volume ratio of the oil and gas phases at the initial pressure was 1:1. The calculation results are shown in Figure 6 (the phase equilibrium calculations in this figure were conducted using the state equations (Equations (13)–(14)) and equilibrium equations (Equations (16)–(21)) from Section 2). The calculation results and experimental results show a good agreement, with a component error control within 5%, indicating that the model accurately calculates the oil and gas components. This error margin arises from multiple factors, including measurement precision in the PVT analysis, minor variations in fluid composition, and the simplifications inherent in the thermodynamic model, such as the grouping of hydrocarbon components and the use of binary interaction coefficients. Furthermore, similar studies on phase equilibrium calculations typically report deviations ranging from 3% to 6%, suggesting that our results are within a reasonable range for engineering applications. Although minor discrepancies exist, they do not significantly impact the overall phase behavior trends, which are the primary focus of this study in assessing CO₂ injection efficiency. This also demonstrates that the current grouping division of oil and gas phase components has a positive effect on the calculations, which helps improve the accuracy of the calculations.

Calculation Results of Volume Parameters

By calculating the area of Figure 4, the volume ratio of the oil and gas phases of the M131 formation oil-CO2 was estimated at different pressures. The comparison between the estimated results and the calculated results is shown in Figure 7 (the volume calculations were performed based on the state equations (Equations (13)–(14)) and volumetric balance equations (Equations (16)–(23)) from Section 2). The solid-colored region represents the area calculated from the photographs taken in the experiment, while the square-filled region represents the calculated volume ratio of the oil and gas phases. The error in volume calculation for the CO2-multiple hydrocarbon system is less than 4%, indicating strong consistency between the model predictions and experimental observations. This deviation is primarily due to uncertainties in experimental volume measurements, as well as the assumptions made in the equation of state modifications. Given that previous studies report volume prediction errors of approximately 3–6% in similar thermodynamic calculations, our model’s accuracy remains within an acceptable range. More importantly, the overall trends in phase behavior remain consistent, ensuring that the model can be effectively used for analyzing CO₂ injection in unconventional reservoirs This indicates that the computational model constructed in the second part effectively captures the influence of pressure changes on the phase transition of fluids at conventional scales.

3.2.2. Multistage Contact Experiments and Component Parameter Calculation Results

The mixing of CO2 and formation oil is divided into forward multi-stage contact mixing (evaporative mixing) and backward multi-stage contact mixing (condensation mixing). In this experiment, forward and backward multi-stage contact mixing experiments were conducted to simulate the process of gas enrichment or oil depletion during injection. Based on the accurate validation of a single contact experiment, the accuracy of the model calculations under fluid composition changes was further validated through multi-stage contact experiments.
The specific method is illustrated in Figure 8. In the forward contact experiment, fresh formation oil and injected gas were mixed, and then the resulting liquid phase was removed. This process was repeated to investigate the changes in gas phase composition. On the contrary, in the backward contact experiment, fresh injected gas was brought into contact with the formation oil, and the resulting gas phase was removed. Fresh injected gas was then introduced, and this process was repeated to investigate the changes in the oil phase. In this experiment, B131 block formation oil was selected as the liquid phase, and natural gas (95% C1 + 5% C2) was chosen as the gas phase. The experimental pressure was 33.1 MPa, and a total of four forward contact experiments and backward contact experiments were conducted.
The results of the multi-stage contact experiments and calculations for the B131 formation oil–gas are shown in Figure 9 and Figure 10 (the calculations used in this figure are based on the phase equilibrium equations (Equations (11)–(14)) and equilibrium expressions (Equations (16)–(21)) from Section 2) for forward multi-stage contact and in Figure 11 and Figure 12 for backward multi-stage contact. The errors in the gas and liquid phase component compositions for multiple contacts were controlled within 3%. The calculated gas and liquid phase volumes and molecular weights were compared with the results from multiple contact experiments, with errors controlled within 10%. These results demonstrate that the model exhibits good stability and accuracy in calculating the fugacity coefficient, volume, and viscosity for CO2-hydrocarbon systems. The model is capable of simulating the mass transfer process of component exchange in actual oil and gas multiple contact scenarios.

3.3. Literature Comparison and Verification

At the conventional scale, the model has been validated through gas injection expansion, single-stage contact mass transfer, and multi-stage contact experiments. Under restricted conditions, this study also compared the model predictions with recent research findings. Han et al. [43] investigated the phase behavior of oil-CO2-water systems under water-alternating-gas (CO2-WAG) injection and found that the saturation pressure of the oil-CO2-water system was lower than that of the oil-CO2 system due to partial CO2 dissolution in water, which led to a shift in the phase envelope. The model predictions in this study exhibit a similar trend, further confirming the accuracy of the proposed model. Additionally, this study selected data from the article published by Wu et al. in 2020 [42] for repeated calculation and verification, as shown in Figure 13. The calculated gas–liquid interfacial tension at 20 nm, 50 nm, 100 nm, and 200 nm was found to be consistent with the results reported in the literature, indicating that the thermodynamic calculation model developed in this study can meet the computational requirements for restricted fluids.
Through experimental validation and a literature comparison, the model has been thoroughly verified. By grouping the fluid components in the target reservoir, a foundation has been laid for subsequent phase equilibrium calculations at multiple scales. Validation has been carried out through single-stage contact experiments and multi-stage contact experiments to verify important parameters such as component composition, volume, and pressure. At the conventional scale, the model meets the computational requirements. However, it is important to note that experimental measurements have inherent uncertainties, and empirical parameters, such as binary interaction coefficients and isoenthalpic compressibility, for the pseudo-component fluid also introduce certain errors. These errors are in line with the characteristics of thermodynamic models, and it can be considered that the model has a certain level of accuracy. By comparing with the literature, the calculation of restricted fluids at the nanoscale has been validated. The model generally agrees with the literature and satisfies the computational requirements.

4. Results and Discussion

In order to investigate the phase behavior of oil and gas at multiple scales, this section presents the calculations and diagrams of phase diagrams for four different fluids, namely B131, B18, B79, and B18-CO2 (1:1), at sizes of 10 nm, 50 nm, 100 nm, 150 nm, and 200 nm, as well as at bulk phase. Taking B79 reservoir oil as an example, three aspects, including density expansion, differential separation, and multi-stage contact simulation, were explored to comprehensively analyze the possible phase behavior of reservoir fluids in restricted spaces.

4.1. Phase Diagram Analysis

By controlling the fugacity, the bubble and dew points at different pressures were determined, and a large amount of bubble and dew point data were obtained through iterative calculations. The fluid phase diagrams for B131, B18, B79, and B18-CO2 (1:1) at different scales are shown in Figure 14 and Figure 15. When the size is larger than 150 nm, the phase envelope diagram remains essentially unchanged. However, when the size is smaller than 150 nm, the envelope lines contract towards the central region. When the scale decreases to 100 nm, the fluid phase envelope diagram exhibits significant contraction, and when the scale decreases to 10 nm, the phase envelope diagram contracts by an average of around 28%, as shown in Figure 14. This contraction indicates a significant shift in bubble and dew point pressures, which suggests that phase equilibrium conditions in nanopores deviate substantially from those in bulk conditions. Notably, Figure 15 further highlights that this contraction effect is more pronounced for higher gas content systems (e.g., B18-CO₂), where gas–liquid interactions dominate phase behavior. Recent studies have also explored the phase behavior of CO2 in nanoporous media. Wan et al. [44] found that CO2 solubility decreases in smaller pores, and the phase envelope contraction effect becomes more pronounced. This aligns with our findings, as shown in Figure 14, where the phase envelope reduction reaches 28% in pores below 10 nm. These results indicate that traditional bulk-phase models may not accurately predict phase behavior in nano-scale reservoirs, necessitating scale-dependent thermodynamic models for improved predictions. Their study showed that the phase envelope shrinkage reached 20–30% for pores smaller than 10 nm, which aligns well with the simulation results of this study, further validating the model’s applicability. These findings have important implications for reservoir management and CO2 injection strategies. The observed phase envelope contraction at nanoscale indicates that phase separation and miscibility conditions in tight reservoirs may significantly differ from those predicted by conventional bulk-phase models. This suggests that for CO2-EOR processes in nano-scale reservoirs, injection pressures may need to be adjusted to ensure optimal miscibility and maximize oil displacement efficiency. Furthermore, the reduced bubble point pressure in confined systems implies that gas liberation and exsolution mechanisms will be delayed, which could influence production strategies in ultra-low permeability reservoirs. These insights can help refine reservoir simulation models and improve the design of CO2 injection schemes in unconventional formations. Furthermore, compared to B18 reservoir oil and B18 reservoir oil mixed with CO2 at a ratio of 1:1, the bubble point pressure increases relatively with the increase in gas phase. In terms of component composition, as the proportion of gas phase increases (B79 < B18 < B131), the influence of scale decreases, and the magnitude of the decrease in bubble point pressure at 10 nm increases (23%~31%), indicating a more significant impact of restricted fluids on the gas phase. Compared to conventional EOS-based models that introduce empirical capillary pressure corrections to account for nanoscale effects, the Helmholtz free energy-based model naturally incorporates these confinement effects into the thermodynamic potential. This allows for a more accurate prediction of bubble point pressure shifts, density variations, and interfacial tension changes in nanoporous media. Unlike conventional approaches, which require separate parameter tuning to fit experimental data, this model inherently captures the thermodynamic constraints imposed by nanoscale confinement, making it more reliable for confined phase behavior modeling. The simulation results demonstrate that the predicted phase envelope shrinkage aligns well with experimental trends, further validating the robustness of this approach.

4.2. Density Expansion Simulation

Reservoirs suitable for CO2-enhanced oil recovery (EOR) often have poor permeability and slow pressure transmission. As a result, the reservoir fluids in some pore throats undergo a pressure drop process. Before effective flow is established, the reservoir fluids are confined in closed spaces, and the phase behavior of the fluids at different scales varies with the pressure drop process. Therefore, it is necessary to conduct a density expansion simulation for this purpose. In this section, the B79-CO2 system is taken as an example to simulate the pressure drop and density expansion and to explore the phase behavior of the fluid at different scales. The simulation results are shown in Figure 16. As the scale decreases, the saturation pressure of the reservoir fluid gradually decreases, especially when it is below 100 nm, the deviation in saturation pressure can reach around 5% to 23%, as shown in Figure 16. This suggests that confined fluids in nanopores exhibit a delayed gas liberation effect, which has direct implications for CO2 injection strategies in tight reservoirs. Figure 16a further illustrates how saturation pressure changes across different scales, showing a more significant deviation below 50 nm. Additionally, Figure 16b presents the variation in fluid density, revealing that confined fluids exhibit an increased density of approximately 2% compared to bulk conditions. These results emphasize the importance of considering nanoscale effects when designing CO2 injection schemes, as failure to account for these shifts could lead to suboptimal reservoir performance predictions. This indicates that the fluids in smaller pores have more difficulty degassing. Compared to the BULK state, the fluid density increases by an average of 2%, and the volume decreases by an average of 2% for sizes below 10 nm. Density variation is a crucial phenomenon in nanopores. Pan et al. [45] demonstrated that, in CO₂-hydrocarbon systems, the confinement effect at the nanoscale results in a 1.5–2.8% increase in gas–liquid density compared to bulk conditions. The experimental results in this study indicate a density expansion ratio of approximately 2% within the 10 nm pore size range, further verifying the universality of this trend. In terms of lateral comparison, as the scale decreases below 50 nm, the compressibility factor and compressibility of the fluid also undergo significant changes. In restricted spaces, the ability of the fluid to degas and expand is weakened. This means that the fluid in the pore may exhibit gas–liquid two-phase behavior before passing through the throat, leading to the Jamin effect. Additionally, when passing through the throat, the gas phase may dissolve back into the liquid phase, thereby weakening or even eliminating the J effect caused by partial degassing.

4.3. Differential Separation Simulation

As the pressure depletion production progresses and gas–liquid separation occurs, the fluid composition within the near-well matrix undergoes changes, as shown in Figure 17 and Figure 18. At 50 bar, the proportion of C1 components in the liquid phase increases significantly at 10 nm and 50 nm, while the proportions of C1–C4 components in the gas phase decrease and the proportion of N2 increases. This indicates that during the degassing process, N2, which is more difficult to dissolve, is preferentially removed, while C1, as the main dissolved gas component, decreases in proportion, indicating that a considerable amount of C1 remains in the liquid phase. This also explains why the density of the liquid phase decreases as the scale decreases.

4.4. Multi-Stage Contact Simulation

The basis of multi-stage contact mixing is the exchange and mass transfer of components between the two-phase fluids. Therefore, understanding the variations in fluid composition at different scales is crucial in describing the phase behavior changes during multi-stage contact. As shown in Figure 19, Figure 20 and Figure 21, in the backward multi-stage contact simulations for the fourth and fifth stages, as the scale decreases, the CO2 content in the liquid phase gradually increases (from 69.01% to 72.56% and from 68.42% to 72.33%), while the content of C2-C5 components also increases, and the content of C7+ decreases. This indicates that as the scale decreases, the ability of CO2 to dissolve into the reservoir oil is enhanced, and the reservoir oil retains gas phase components more effectively. The mass transfer between the gas and liquid phases becomes more significant, and the properties of the liquid phase approach those of the CO2 phase. Consequently, phase mixing becomes easier, which is consistent with numerous recent research findings. In the forward multi-stage contact simulations, after one contact at 100nm and 50nm, the oil and gas phases form a single-phase state after the second contact. However, at 10nm, a single-phase state is achieved after the first contact. This indicates that phase mixing formed by gas extraction and stripping is easier compared to the phase mixing formed by CO2 dissolution.

5. Conclusions

(1)
Based on the traditional thermodynamic theory, the thermodynamic fluid phase equilibrium calculation model was established by introducing the Helmholtz free energy and considering capillary force, critical point transition, and adsorption. This model provides a new approach for calculating phase parameters of oil and gas at multiple scales. It clarifies the limits (around 200 nm) and extent (2.5 MPa pressure drop for oil–gas mixing at 50 nm) of changes in the phase properties of oil and gas under spatial variations (PVT cylinder and porous media). The theoretical analysis also reveals the phase characteristics changes in the CO2 displacement front, oil–gas mixing zone, diffusion zone, and pressure drop zone within the porous media. Unlike conventional Gibbs free energy models, which require empirical corrections for nanoscale effects, the Helmholtz free energy model inherently integrates capillary pressure and adsorption effects, leading to more accurate phase behavior predictions and improved computational stability.
(2)
Taking B131, B18, and B79 reservoir oil components as examples and combining them with gas injection expansion experiments for component splitting, the model was validated by comparing the results of single-stage contact experiments, multiple-stage contact experiments, and previous research findings. The results confirm that the model provides accurate calculations for both bulk and nanoscale systems.
(3)
Using a B79-CO2 displacement block in a specific oilfield as an example, the phase behavior of the fluid during the CO2 injection process was analyzed through phase diagram analysis, density expansion simulation, differential separation simulation, and multi-stage contact simulation. The results clarify the phase characteristics of the fluid under different scales. As the scale decreases (from 200 nm to 10 nm), the fluid experiences enhanced confinement effects, making gasification more difficult in the liquid phase and resulting in an increase in overall density (1.7% increase at 10 nm). The mass transfer and phase mixing abilities between CO2 and reservoir oil increase with decreasing scale. Under the condition of sufficient contact between oil and gas, the reduction in scale has a positive effect on improving the oil-washing efficiency of CO2. These findings indicate that scale-dependent phase behavior should be considered in CO2-EOR designs. For ultra-tight formations, adjusting CO2 injection rates and optimizing contact time based on the scale of nanopores can enhance oil recovery efficiency. Field data integration with the model can provide actionable insights for designing site-specific injection strategies.
(4)
Application to CO2-EOR optimization: The findings of this study provide practical guidelines for optimizing CO2 injection strategies in unconventional reservoirs. The observed phase behavior shifts suggest that higher injection pressures (e.g., above 15 MPa for nanoporous reservoirs) are critical to maintaining miscibility and preventing early gas liberation. Additionally, the enhanced CO2 solubility at nanoscale indicates that Continuous injection is recommended over cyclic methods like WAG for nanoporous formations due to enhanced CO2 solubility. For field implementation, the proposed model can be embedded in reservoir simulators to refine injection schedules and predict phase transitions under confined conditions. For example, simulations can identify optimal injection pressure ranges tailored to specific pore size distributions. This approach aids in designing injection schedules that optimize miscibility windows and minimize CO2 loss. Moreover, integrating this model with real-time monitoring data can enhance dynamic adjustment strategies for CO2 flooding efficiency in heterogeneous formations. This model can assist in fine-tuning injection pressures and predicting phase stability across multiple scales, contributing to improved oil recovery efficiency in complex porous media.

Author Contributions

Writing—original draft, writing—review and editing, conceptualization, F.C.; supervision, B.L., X.P., B.J., X.H., Z.Z. and C.F.; conceptualization, L.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the CNPC Major Project “Research on New Methods and Technologies for Enhanced Oil Recovery”, grant number 2023ZZ04, CNPC Major project “CCUS Oil Displacement Geological Body Fine Description and Reservoir Engineering Key Technology Research”, grant number 2021ZZ01-03, CNPC Science & Technology Research Institute Open Fund Project “Study on the Mechanism of Oil–Gas Interaction on Relative Permeability of CO2-Oil System”, grant number 2023-KFKT-23, and the CNPC Science & Technology Research Institute Open Fund Project “Study on Multi-Medium Flow-Solid Coupling Mechanisms in Ultra-Deep Carbonate Reservoirs”, grant number 2024-KFKT-21.

Data Availability Statement

Not appliable.

Acknowledgments

The support given by The State Key Laboratory of Enhanced Oil Recovery of Open Fund Funded Project, Major Special Projects of CNPC, is acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lau, H.C.; Ramakrishna, S.; Zhang, K.; Radhamani, A.V. The Role of Carbon Capture and Storage in the Energy Transition. Energy Fuels 2021, 35, 7364–7386. [Google Scholar] [CrossRef]
  2. Martin-Roberts, E.; Scott, V.; Flude, S.; Johnson, G.; Haszeldine, R.S.; Gilfillan, S. Carbon capture and storage at the end of a lost decade. One Earth 2021, 4, 1569–1584. [Google Scholar] [CrossRef]
  3. Hoteit, H. Proper Modeling of Diffusion in Fractured Reservoirs. In Proceedings of the SPE Reservoir Simulation Symposium, The Woodlands, TX, USA, 21–23 February 2011. [Google Scholar]
  4. Erfan, M.; Badrul Mohamed, J.; Amin, A.; Hossein, H.; Nur Hidayati Binti, O.; Aqilah, D.; Siti Nurliyana Binti Che Mohamed, H.; Rozana Azrina Binti, S. CO2-EOR/Sequestration: Current Trends and Future Horizons. In Enhanced Oil Recovery Processes; Ariffin, S., Ed.; IntechOpen: Rijeka, Croatia, 2019; Chapter 7. [Google Scholar]
  5. Song, Y.; Jun, S.; Na, Y.; Kim, K.; Jang, Y.; Wang, J. Geomechanical challenges during geological CO2 storage: A review. Chem. Eng. J. 2023, 456, 140968. [Google Scholar] [CrossRef]
  6. Kim, T.W.; Yoon, H.C.; Lee, J.Y. Review on carbon capture and storage (CCS) from source to sink; part 1: Essential aspects for CO2 pipeline transportation. Int. J. Greenh. Gas Control 2024, 137, 104208. [Google Scholar] [CrossRef]
  7. Wang, L.; He, Y.; Wang, Q.; Liu, M.; Jin, X. Multiphase flow characteristics and EOR mechanism of immiscible CO2 water-alternating-gas injection after continuous CO2 injection: A micro-scale visual investigation. Fuel 2020, 282, 118689. [Google Scholar] [CrossRef]
  8. Ren, D.; Wang, X.; Kou, Z.; Wang, S.; Wang, H.; Wang, X.; Tang, Y.; Jiao, Z.; Zhou, D.; Zhang, R. Feasibility evaluation of CO2 EOR and storage in tight oil reservoirs: A demonstration project in the Ordos Basin. Fuel 2023, 331, 125652. [Google Scholar] [CrossRef]
  9. Khan, M.Y.; Mandal, A. Analytical model of incremental oil recovery as a function of WAG ratio and tapered WAG ratio benefits over uniform WAG ratio for heterogeneous reservoir. J. Pet. Sci. Eng. 2022, 209, 109955. [Google Scholar] [CrossRef]
  10. Xu, H. Probing nanopore structure and confined fluid behavior in shale matrix: A review on small-angle neutron scattering studies. Int. J. Coal Geol. 2020, 217, 103325. [Google Scholar] [CrossRef]
  11. Li, S.Y.; Wang, L.; Su, L.N.; Li, Z.M.; Zhang, K.Q. Carbon dioxide diffusions in Methane-Dissolved pore Fluids: Implications for geological carbon storage and utilization in tight formations. Chem. Eng. J. 2022, 429, 132147. [Google Scholar] [CrossRef]
  12. Li, B.; Mehmani, A.; Chen, J.; Georgi, D.; Jin, G. The Condition of Capillary Condensation and Its Effects on Adsorption Isotherms of Unconventional Gas Condensate Reservoirs. In Proceedings of the SPE Annual Technical Conference and Exhibition, New Orleans, LA, USA, 30 September–2 October 2013. [Google Scholar]
  13. Dong, X.; Liu, H.; Hou, J.; Wu, K.; Chen, Z. Phase Equilibria of Confined Fluids in Nanopores of Tight and Shale Rocks Considering the Effect of Capillary Pressure and Adsorption Film. Ind. Eng. Chem. Res. 2016, 55, 798–811. [Google Scholar] [CrossRef]
  14. Li, D.C.; Zhang, Y.; Jiao, Z.S.; Saraji, S. Three-dimensional core reconstruction and performance evaluation of CO2 displacement in a tight oil reservoir. Fuel 2023, 349, 128622. [Google Scholar] [CrossRef]
  15. Zhang, H. Study on microscale stress sensitivity of CO2 foam fracturing in tight reservoirs. Energy 2024, 294, 130766. [Google Scholar] [CrossRef]
  16. Sun, R.; Xu, K.; Huang, T.; Zhang, D. Methane Diffusion Through Nanopore- Throat Geometry: A Molecular Dynamics Simulation Study. Spe J. 2023, 28, 819–830. [Google Scholar] [CrossRef]
  17. Deng, J.; Guo, S.; Wan, J.; Zhang, L.; Song, H. Molecular dynamics of CH4 adsorption and diffusion characteristics through different geometric shale kerogen nanopores. Chem. Eng. J. 2024, 500, 156784. [Google Scholar] [CrossRef]
  18. Zhao, Y.; Luo, M.; Liu, L.; Wu, J.; Chen, M.; Zhang, L. Molecular dynamics simulations of shale gas transport in rough nanopores. J. Pet. Sci. Eng. 2022, 217, 110884. [Google Scholar] [CrossRef]
  19. de Andrade, D.d.C.J.; Nojabaei, B. Phase Behavior and Composition Distribution of Multiphase Hydrocarbon Binary Mixtures in Heterogeneous Nanopores: A Molecular Dynamics Simulation Study. Nanomaterials 2021, 11, 2431. [Google Scholar] [CrossRef]
  20. Moradi, M.; Mahmoudi, J.; Sadeghzadeh, S. Grand Canonical Monte Carlo Simulation Experiences of Methane and Ethane Adsorption Behaviors on Simplified Organic Shale Formed by Graphene Layering. Energy Fuels 2023, 37, 18698–18712. [Google Scholar] [CrossRef]
  21. Chen, F.; Mehana, M.; Nasrabadi, H. Molecular Simulation of Hydrogen-Shale Gas System Phase Behavior under Multiscale Conditions: A Molecular-Level Analysis of Hydrogen Storage in Shale Gas Reservoirs. Energy Fuels 2023, 37, 2449–2456. [Google Scholar] [CrossRef]
  22. Xing, X.; Feng, Q.; Zhang, W.; Wang, S. Vapor-liquid equilibrium and criticality of CO2 and n-heptane in shale organic pores by the Monte Carlo simulation. Fuel 2021, 299, 120909. [Google Scholar] [CrossRef]
  23. Vaganova, M.; Nesterova, I.; Kanygin, Y.; Kazennov, A.; Khlyupin, A. Linking theoretical and simulation approaches to study fluids in nanoporous media: Molecular dynamics and classical density functional theory. Chem. Eng. Sci. 2022, 25, 0117383. [Google Scholar] [CrossRef]
  24. Wang, Y.; Shardt, N.; Lu, C.; Li, H.; Elliott, J.A.W.; Jin, Z. Validity of the Kelvin equation and the equation-of-state-with-capillary-pressure model for the phase behavior of a pure component under nanoconfinement. Chem. Eng. Sci. 2020, 226, 115839. [Google Scholar] [CrossRef]
  25. Wang, Y.; Lei, Z.; Sun, L.; Pan, X.; Liu, Y.; Xu, Z.; Zheng, X.; Wang, Y.; Liu, P. Study on the minimum miscibility pressure and phase behavior of CO2-shale oil in nanopores. Chem. Eng. J. 2024, 497, 154493. [Google Scholar] [CrossRef]
  26. Pospíšil, M.; Malijevský, A. Phase behavior of fluids in undulated nanopores. Phys. Rev. E 2022, 106, 024801. [Google Scholar] [CrossRef] [PubMed]
  27. Song, Z.; Song, Y.; Guo, J.; Zhang, Z.; Hou, J. Adsorption induced critical shifts of confined fluids in shale nanopores. Chem. Eng. J. 2020, 385, 123837. [Google Scholar] [CrossRef]
  28. Huang, J.; Yin, X.; Barrufet, M.; Killough, J. Lattice Boltzmann simulation of phase equilibrium of methane in nanopores under effects of adsorption. Chem. Eng. J. 2021, 419, 129625. [Google Scholar] [CrossRef]
  29. Travalloni, L.; Castier, M.; Tavares, F.W. Phase equilibrium of fluids confined in porous media from an extended Peng–Robinson equation of state. Fluid Phase Equilibria 2014, 362, 335–341. [Google Scholar] [CrossRef]
  30. Barbosa, G.D.; D’Lima, M.L.; Daghash, S.M.H.; Castier, M.; Tavares, F.W.; Travalloni, L. Cubic equations of state extended to confined fluids: New mixing rules and extension to spherical pores. Chem. Eng. Sci. 2018, 184, 52–61. [Google Scholar] [CrossRef]
  31. Alafnan, S. Adsorption-Desorption Hysteresis in Shale Formation: New Insights into the Underlying Mechanisms. Energy Fuels 2022, 36, 5307–5315. [Google Scholar] [CrossRef]
  32. Feng, D.; Chen, Z.; Zhao, W.; Wu, K.; Li, J.; Li, X.; Gao, Y.; Zhang, S.; Peng, F. Determination of Apparent Pore Size Distributions of Organic Matter and Inorganic Matter in Shale Rocks Based on Water and N2 Adsorption. Energy Fuels 2022, 36, 11787–11797. [Google Scholar] [CrossRef]
  33. Lin, K.; Huang, X.; Zhao, Y.-P. Combining Image Recognition and Simulation to Reproduce the Adsorption/Desorption Behaviors of Shale Gas. Energy Fuels 2020, 34, 258–269. [Google Scholar] [CrossRef]
  34. Luo, S.; Lutkenhaus, J.L.; Nasrabadi, H. Confinement-Induced Supercriticality and Phase Equilibria of Hydrocarbons in Nanopores. Langmuir 2016, 32, 11506–11513. [Google Scholar] [CrossRef] [PubMed]
  35. Luo, S.; Nasrabadi, H.; Lutkenhaus, J.L. Effect of confinement on the bubble points of hydrocarbons in nanoporous media. Aiche J. 2016, 62, 1772–1780. [Google Scholar] [CrossRef]
  36. Pipertzis, A.; Abdou, N.; Xu, J.; Asp, L.E.; Martinelli, A.; Swenson, J. Ion transport, mechanical properties and relaxation dynamics in structural battery electrolytes consisting of an imidazolium protic ionic liquid confined into a methacrylate polymer. Energy Mater. 2023, 3, 300050. [Google Scholar] [CrossRef]
  37. Qiu, X.; Tan, S.P.; Dejam, M.; Adidharma, H. Binary fluid mixtures confined in nanoporous media: Experimental evidence of no phase coexistence. Chem. Eng. J. 2021, 405, 127021. [Google Scholar] [CrossRef]
  38. Yang, Q.; Jin, B.; Banerjee, D.; Nasrabadi, H. Direct visualization and molecular simulation of dewpoint pressure of a confined fluid in sub-10 nm slit pores. Fuel 2019, 235, 1216–1223. [Google Scholar] [CrossRef]
  39. Alfi, M.; Nasrabadi, H.; Banerjee, D. Effect of Confinement on Bubble Point Temperature Shift of Hydrocarbon Mixtures: Experimental Investigation Using Nanofluidic Devices. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 9–11 October 2017. [Google Scholar]
  40. Xie, C.; Li, H. Multiscale simulations of nanofluidics: Recent progress and perspective. WIREs Comput. Mol. Sci. 2023, 13, e1661. [Google Scholar] [CrossRef]
  41. Zarragoicoechea, G.J.; Kuz, V.A. Critical shift of a confined fluid in a nanopore. Fluid Phase Equilibria 2004, 220, 7–9. [Google Scholar] [CrossRef]
  42. Wu, S.; Li, Z.; Sarma, H.K. Influence of confinement effect on recovery mechanisms of CO2-enhanced tight-oil recovery process considering critical properties shift, capillarity and adsorption. Fuel 2020, 262, 116569. [Google Scholar] [CrossRef]
  43. Han, X.; Song, Z.; Deng, S.; Li, B.; Li, P.; Lan, Y.; Song, Y.; Zhang, L.; Zhang, K.; Zhang, Y. Multiphase behavior and fluid flow of oil–CO2–water in shale oil reservoirs: Implication for CO2-water-alternating-gas huff-n-puff. Phys. Fluids 2024, 36, 063310. [Google Scholar] [CrossRef]
  44. Wan, T.; Ding, K.; Xiong, Q.; Guo, J. The phase behavior of CO2 injection in shale reservoirs with nano-pores. RSC Adv. 2024, 14, 27227–27240. [Google Scholar] [CrossRef]
  45. Pan, X.; Sun, L.; Liu, Q.; Huo, X.; Chen, F.; Wang, Y.; Feng, C.; Zhang, Z.; Ni, S. Mechanism of CO2 flooding in shale reservoirs—insights from nanofluids. Nanoscale, 2025; Advance Article. [Google Scholar] [CrossRef]
Figure 1. The average error rate of model calculation and experimental results under different grouping methods.
Figure 1. The average error rate of model calculation and experimental results under different grouping methods.
Molecules 30 00277 g001
Figure 2. Comparison of phase envelope predictions: Helmholtz-based model provides improved accuracy under confinement. (a) Saturation pressure (b) expansion coefficient.
Figure 2. Comparison of phase envelope predictions: Helmholtz-based model provides improved accuracy under confinement. (a) Saturation pressure (b) expansion coefficient.
Molecules 30 00277 g002
Figure 3. Phase envelope shrinkage of B131 formation oil at different pore sizes.
Figure 3. Phase envelope shrinkage of B131 formation oil at different pore sizes.
Molecules 30 00277 g003
Figure 4. Morphology of CO2 and B131 formation oil in PVT as pressure increases.
Figure 4. Morphology of CO2 and B131 formation oil in PVT as pressure increases.
Molecules 30 00277 g004
Figure 5. Gas phase component gradient with pressure.
Figure 5. Gas phase component gradient with pressure.
Molecules 30 00277 g005
Figure 6. B131 formation oil-CO₂ system: liquid and gas phase composition changes with pressure. (a) liquid phase (b) gas phase.
Figure 6. B131 formation oil-CO₂ system: liquid and gas phase composition changes with pressure. (a) liquid phase (b) gas phase.
Molecules 30 00277 g006
Figure 7. Bubble point pressure variation with decreasing pore size.
Figure 7. Bubble point pressure variation with decreasing pore size.
Molecules 30 00277 g007
Figure 8. Saturation pressure reduction in confined nanopores.
Figure 8. Saturation pressure reduction in confined nanopores.
Molecules 30 00277 g008
Figure 9. Comparison of composition changes in multi-stage contact experiments and model calculation results (a) liquid phase (b) gas phase. (The calculations were performed using the forward method.)
Figure 9. Comparison of composition changes in multi-stage contact experiments and model calculation results (a) liquid phase (b) gas phase. (The calculations were performed using the forward method.)
Molecules 30 00277 g009
Figure 10. Comparison of gas/liquid phase volume and molecular weight changes in multi-stage contact experiments with model calculation composition changes. (The calculations in this figure were performed using the forward method.).
Figure 10. Comparison of gas/liquid phase volume and molecular weight changes in multi-stage contact experiments with model calculation composition changes. (The calculations in this figure were performed using the forward method.).
Molecules 30 00277 g010
Figure 11. Comparison of liquid phase composition changes in multi-stage contact experiments and model calculation results (a) liquid phase (b) gas phase. (The calculations in this figure were performed using the backward method.)
Figure 11. Comparison of liquid phase composition changes in multi-stage contact experiments and model calculation results (a) liquid phase (b) gas phase. (The calculations in this figure were performed using the backward method.)
Molecules 30 00277 g011
Figure 12. Comparison of gas/liquid phase volume and molecular weight changes in multi-stage contact experiments with model calculation composition changes. (The calculations in this figure were performed using the backward method.)
Figure 12. Comparison of gas/liquid phase volume and molecular weight changes in multi-stage contact experiments with model calculation composition changes. (The calculations in this figure were performed using the backward method.)
Molecules 30 00277 g012
Figure 13. (a) Gas–liquid interfacial tension vs. scale variation [42]. (b) Model repeated calculation of gas–liquid interfacial tension versus scale change and the verification effect of Wu et al.
Figure 13. (a) Gas–liquid interfacial tension vs. scale variation [42]. (b) Model repeated calculation of gas–liquid interfacial tension versus scale change and the verification effect of Wu et al.
Molecules 30 00277 g013
Figure 14. The PT phase envelope of formation oil changes at different scales (a) B131 (b) B18 (c) B79 (d) B18 -CO2.
Figure 14. The PT phase envelope of formation oil changes at different scales (a) B131 (b) B18 (c) B79 (d) B18 -CO2.
Molecules 30 00277 g014
Figure 15. Enlarged results of PT phase envelopes at different scales of fluid near the formation temperature (a) B131 (b) B18 (c) B79 (d) B18-CO2.
Figure 15. Enlarged results of PT phase envelopes at different scales of fluid near the formation temperature (a) B131 (b) B18 (c) B79 (d) B18-CO2.
Molecules 30 00277 g015
Figure 16. (a) Variation law of fluid saturation pressure at different scales in the equilibrium mass expansion process. (b) Variation law of fluid density at different scales for equilibrium mass expansion.
Figure 16. (a) Variation law of fluid saturation pressure at different scales in the equilibrium mass expansion process. (b) Variation law of fluid density at different scales for equilibrium mass expansion.
Molecules 30 00277 g016
Figure 17. Scale-dependent variation of liquid phase composition at 50 bar increased retention of heavier hydrocarbons in smaller pores.
Figure 17. Scale-dependent variation of liquid phase composition at 50 bar increased retention of heavier hydrocarbons in smaller pores.
Molecules 30 00277 g017
Figure 18. Scale-dependent variation of gas phase composition at 50 bar enhanced CO2 retention and depletion of lighter hydrocarbons in smaller pores.
Figure 18. Scale-dependent variation of gas phase composition at 50 bar enhanced CO2 retention and depletion of lighter hydrocarbons in smaller pores.
Molecules 30 00277 g018
Figure 19. Variation of liquid phase components at different scales in the fourth stage of backward multi-stage contact process.
Figure 19. Variation of liquid phase components at different scales in the fourth stage of backward multi-stage contact process.
Molecules 30 00277 g019
Figure 20. Variation of liquid phase components at different scales in the fifth stage of backward multi-stage contact process.
Figure 20. Variation of liquid phase components at different scales in the fifth stage of backward multi-stage contact process.
Molecules 30 00277 g020
Figure 21. Variation of gas phase components at different scales in the first stage forword multi-stage contact process.
Figure 21. Variation of gas phase components at different scales in the first stage forword multi-stage contact process.
Molecules 30 00277 g021
Table 2. Formation oil components of BLOCK B131, B18, B79.
Table 2. Formation oil components of BLOCK B131, B18, B79.
ComponentB79B131B18
CO20.1530.2520.208
N22.8183.7083.06
C116.19353.52244.168
C23.9383.1122.568
C33.2240.7160.591
iC41.6750.2830.234
nC42.9780.2190.181
iC50.9040.1240.102
nC52.5941.3581.7
C62.4312.0662.625
C73.8353.5034.507
C85.1314.2035.404
C94.2253.7544.83
C103.8973.8254.921
C113.362.1192.727
C123.2562.9793.833
C133.2711.9762.543
C142.6971.7842.295
C152.7461.2421.599
C162.2081.5612.008
C172.2491.2911.661
C181.9990.8041.034
C191.9210.6370.819
C201.7640.4840.623
C211.6040.4830.622
C221.5540.4140.533
C231.4310.3670.472
C241.3850.3030.389
C251.2320.2860.368
C261.1530.2540.327
C271.0620.2530.326
C281.0240.2190.282
C290.9360.1910.246
C300.9050.1350.174
C310.70.1010.129
C320.7160.0740.095
C330.5480.0640.083
C340.5320.0670.086
C350.4760.0930.12
C36+5.2751.171.506
Table 3. Split grouping of components in calculation.
Table 3. Split grouping of components in calculation.
NumberGroup1Group2Group3Group4Group5Group6Group7Group8
G1CO2CO2CO2 + C2CO2CO2CO2CO2CO2
G2N2N2 + C1N2N2N2C1C1 + N2N2 + C1
G3C1C2–C4C1C1C1C2C2C2
G4C2C5–C6C3–C4C2–C4C2–C6C3C3–C4C3–C4
G5C3C7–C13C5–C6C5–C6C7–C13C4C5C5–C6
G6C4C14–C19C7–C13C7–C19C14–C19C5C6C7–C13
G7C5C20–C32C14–C19C20–C25C20–C25C6C7–C13C14–C19
G8C6C33+C20–C32C25–C32C25+C7–C19C14–C19C20+
G9C7+ C33+C33+ C20+C20–C32
G10 C33+
G11
G12
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

Chen, F.; Sun, L.; Li, B.; Pan, X.; Jiang, B.; Huo, X.; Zhang, Z.; Feng, C. Computational Modeling and Experimental Investigation of CO2-Hydrocarbon System Within Cross-Scale Porous Media. Molecules 2025, 30, 277. https://doi.org/10.3390/molecules30020277

AMA Style

Chen F, Sun L, Li B, Pan X, Jiang B, Huo X, Zhang Z, Feng C. Computational Modeling and Experimental Investigation of CO2-Hydrocarbon System Within Cross-Scale Porous Media. Molecules. 2025; 30(2):277. https://doi.org/10.3390/molecules30020277

Chicago/Turabian Style

Chen, Feiyu, Linghui Sun, Bowen Li, Xiuxiu Pan, Boyu Jiang, Xu Huo, Zhirong Zhang, and Chun Feng. 2025. "Computational Modeling and Experimental Investigation of CO2-Hydrocarbon System Within Cross-Scale Porous Media" Molecules 30, no. 2: 277. https://doi.org/10.3390/molecules30020277

APA Style

Chen, F., Sun, L., Li, B., Pan, X., Jiang, B., Huo, X., Zhang, Z., & Feng, C. (2025). Computational Modeling and Experimental Investigation of CO2-Hydrocarbon System Within Cross-Scale Porous Media. Molecules, 30(2), 277. https://doi.org/10.3390/molecules30020277

Article Metrics

Back to TopTop