Next Article in Journal
Stability Analysis of Isolated Roof in Overlapping Goaf Based on Strength Reduction
Next Article in Special Issue
Denudation Recovery and Prototype Basin Structural Pattern During the Faulting Period in the Southern Part of the Central Depression Area of the Songliao Basin
Previous Article in Journal
Verifiable Key-Aggregate Searchable Encryption Scheme for Fog-Based Internet of Autonomous Vehicles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Multiscale Compositional Numerical Study in Tight Oil Reservoir: Incorporating Capillary Forces in Phase Behavior Calculation

1
School of Energy Resources, China University of Geosciences (Beijing), Beijing 100083, China
2
Research Institute of Petroleum Exploration & Development, PetroChina Company Limited, Beijing 100083, China
3
Shandong Institute of Petroleum and Chemical Technology, Dongying 257061, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(6), 3082; https://doi.org/10.3390/app15063082
Submission received: 4 February 2025 / Revised: 6 March 2025 / Accepted: 6 March 2025 / Published: 12 March 2025

Abstract

:
Tight oil reservoirs offer significant development potential. Due to the pronounced capillary forces in their nanopores, phase behavior differs markedly from that in conventional reservoirs, challenging traditional equations of state and numerical simulation methods. This paper presents a multiscale compositional numerical simulation method that incorporates capillary forces, leveraging the parallel advantages of the multiscale finite volume method. The approach decouples the compositional model using a sequential format to derive pressure and transport equations, then solves the pressure equation iteratively in a multiscale format to enhance computational efficiency. Results show that the proposed method significantly improves simulation speed while maintaining accuracy. By considering capillary forces in phase equilibrium calculations, this model effectively characterizes phase behavior in tight oil reservoir development, making it highly relevant for Pressure Volume Temperature (PVT) simulation, development simulation, and forecasting development strategies.

1. Introduction

As oil and gas exploration transitions from conventional to unconventional fields, tight oil and gas have emerged as key resources for boosting China’s oil and gas production. China’s tight oil geological resources are estimated at approximately 200 × 10⁸ t, highlighting their abundance and potential, which have garnered increasing attention in recent years [1,2,3]. A defining characteristic of tight oil reservoirs is their extremely small pore size [4]. While conventional oil reservoirs typically have pore sizes exceeding 2 μm [5,6], tight oil reservoir pores range from 5 to 750 nm, with 20–40% measuring below 10 nm [7,8]. This nanometer-scale pore size amplifies capillary forces, significantly altering fluid phase behavior compared to conventional reservoirs [9,10,11].
Phase behavior in conventional reservoirs is typically described by traditional equations of state (EOS), such as Redlich-Kwong equation of state (RK-EOS), Soave-Redlich-Kwong equation of state (SRK-EOS), and Peng-Robinson equation of state (PR-EOS) [12,13,14]. These EOS assume that phase behavior is independent of pore size and provide sufficient accuracy for conventional oil and gas calculation [15]. However, in tight reservoirs, the nanometer-scale throats and strong interactions between pore walls and fluids significantly influence fluid phase behavior, deviating from the phase behavior observed in conventional reservoirs. Consequently, the phase state theory for fluids in large conventional pores becomes inadequate for tight reservoirs [16].
Numerous studies have confirmed that the phase behavior of fluids in nanometer pores cannot be accurately described by traditional equations of state [17,18]. Ribeiro [19] and Qiao [20] observed through adsorption-desorption experiments that the saturation pressure of fluids in nanometer pores is lower than in larger pores. Morishige [21], Russo [22], and Yun [23] further demonstrated that the critical temperature of fluids in nanometer pores is reduced compared to larger pores. Similarly, Zeigermann [24] found via diffusion methods that the critical temperature of n-pentane in 6 nm and 15 nm pores is lower than in larger pores. To directly observe and measure these phenomena, Ally [25] employed high-pressure nanofluidic chips and showed significant deviations in the dew point pressure of propane and carbon dioxide within 17–33 nm porous media compared to calculated values. Alfi [26] used microfluidic chips to simulate shale matrices and found that decreasing pore size significantly impacts the bubble point properties of fluids. Zhong [27], using precision-fabricated 8 nm silicon nanochannels, demonstrated that nanoscale pore and throat interactions influence fluid dew point and bubble point pressures. These experimental findings collectively indicate that pore size significantly affects the phase behavior of tight oil reservoirs, rendering traditional EOS inadequate for describing phase behavior in nanometer pores.
To accurately calculate the fluid phase behavior in tight reservoirs, improvements to traditional equations of state are essential. Brusilovsky [28] developed a new EOS for predicting the critical pressure of multi-component hydrocarbon systems and used numerical simulations to study the impact of capillary forces in porous media on dew and bubble points. Pang [29] analyzed the influence of capillary action on fluid saturation pressure in shale gas and oil reservoirs, while Teklu [30] incorporated critical temperature and pressure variations, along with capillary force effects, into the Pressure Volume Temperature (PVT) phase equilibrium of micropore fluids. Teklu also utilized an improved EOS and multi-component mixing algorithm (MMC) to calculate the minimum miscibility pressure for Bakken crude oil and carbon dioxide mixtures. Theoretical analyses by Qi [31], Nojabaei [32], and Zhang Yuan [33] concluded that bubble-point pressure decreases as pore size reduces, with the effect being more pronounced under conditions further from the critical point. Nojabaei [32] coupled capillary pressure with phase equilibrium equations, solving nonlinear fugacity equations and showing that micropores reduce fluid saturation pressure. Li et al. [34] proposed a phase equilibrium calculation method that incorporates pore throat-fluid molecule interactions and capillary forces, verifying through methane, butane, and octane mixtures that pore throat diameters below 10 nm significantly impact phase behavior. Zhang [35] further corrected the PR-EOS for Bakken shale oil by accounting for capillary effects on two-phase systems. These improvements to the traditional PR-EOS, considering capillary pressure and nanopore confinement effects, have been applied in numerical simulations of tight oil reservoirs [36,37,38,39]. However, the iterative computation required for component simulation and EOS adjustments increases computational complexity, limiting industrial adoption. As a result, these approaches are primarily used as alternatives when the accuracy of black oil simulations is insufficient.
To address convergence issues caused by the strong nonlinearity in compositional simulation, many researchers have employed multiscale techniques to optimize the method. These techniques construct multiscale basis functions for coarse grids using fine-scale detailed information, enabling the upscaling and downscaling of data across scales and achieving effective multiscale coupling. The mathematical foundation of the multiscale finite element method (MsFEM) originated from the pioneering work of Babuska and Osborn [40,41]. Hou and Wu [40] extended this concept to elliptic equation problems, introducing the innovative MsFEM. Building on this, Jenny [42] proposed the multiscale finite volume method (MsFVM), inspired by MsFEM, and applied it to incompressible flow in porous media. Moyner and Li [43], along with Zhou and Tchelepi [44], further extended the approach to black oil simulations. To address the challenges posed by reservoir models with complex geological structures, Lunati and Moyner et al. [45,46,47] incorporated the algebraic multigrid concept, developing multiscale formulations based on operator algebraic representations. These advancements have significantly enhanced the efficiency and applicability of multiscale methods in reservoir simulations.
A review of previous studies reveals that no existing multiscale compositional numerical simulation method for tight oil reservoirs incorporates the effects of capillary forces in phase behavior. Building on prior research, this study first modifies phase behavior calculations to account for capillary forces, proposing a phase calculation method that integrates these effects. Subsequently, the traditional multiscale finite volume method is enhanced to develop a multiscale compositional numerical model that considers capillary forces. Finally, the proposed model is validated, and a parameter sensitivity analysis is performed.

2. Materials and Methods

This section details the construction of a multiscale compositional numerical simulator incorporating capillary forces. First, it outlines the development of an equation of state that accounts for capillary forces. Next, it describes the sequential formulation and solution of the compositional numerical simulation, which is the foundation for multiscale construction. Following this, an improved iterative multiscale finite volume method is introduced. Finally, these components are integrated into a multiscale compositional numerical simulation method that fully considers capillary forces.

2.1. The Phase Equilibrium Calculation Theory Considers the Capillary Force Effect

The PR-EOS is modified to incorporate capillary pressure, which is influenced by pore geometry and the wettability of the rock surface. In the calculation of capillary pressure between oil and gas, the oil phase is regarded as the wetting phase relative to the gas phase. Under this assumption, the gas phase pressure can be expressed as:
p v = p l + p c
where p v is the gas phase pressure (Pa), p l is the liquid phase pressure (Pa), p c is the capillary pressure (Pa), which is determined by the geometry of the pores and the wettability of the rock surface.
Mercury intrusion experiments quantify the relationship between capillary pressure and mercury saturation (i.e., the pore volume occupied by mercury). Volume averaging of pore radius and interfacial tension allows derivation of this relationship from the Young-Laplace equation, a result validated in multiphase flow simulations [48]. This study investigates the impact of nanopore capillary pressure on phase equilibrium. Consequently, phase behavior calculations incorporate these effects as described by Zhang et al. [35] the calculation of capillary force in Equation (1) refers to the Young-Laplace equation
p c = 2 σ cos θ r
where σ represents the interfacial tension (N/m); θ represents the contact angle (For gas-liquid phase equilibrium in reservoirs, the gas phase is regarded as non-wetting, while the liquid phase is considered wetting, assuming a contact angle of 45° [49] in this study; r represents the pore radius (nm).
The method for calculating the interfacial tension σ in Equation (2) refers to the work of MacLeod and Pedersen [50,51]
σ = ( Ω l ρ ¯ l Ω v ρ ¯ v ) v Ω l = i = 1 N c x i Ω i Ω v = i = 1 N c y i Ω i
where v is the scaling exponent, v = 3.6; ρ ¯ l / v is the average molar densities of the liquid and gas phases in the pore (mol/cm3); x i and y i are the molar fractions of component i in the liquid and gas phases, respectively; Ω i is the isobaric molar volumes of the liquid and gas phases, respectively [50,51]
Ω i = ( 8.21307 + 1.97473 ω i ) T c i 1.03406 P c i 0.82636
where ω i is the acentric factor; T c i is the critical temperature of component i (K); P c i is the critical pressure of component i (bar).
Equations (1)–(3) provide a method for integrating microscopic mechanisms into phase behavior calculations. Since this study focuses on the impact of capillary pressure in nanopores on phase behavior, capillary pressure is computed using an average pore radius in subsequent calculations. However, the option to incorporate a pore size distribution remains. The proposed model can be extended to account for the actual pore radius distribution of a reservoir, based on geological pore size data.
In the PR-EOS, capillary pressure is not accounted for in the gas and liquid phases. Instead, Equations (1) and (2) compute capillary pressure and phase pressures for different pore sizes, subsequently affecting the compressibility factor and fugacity coefficient. As noted above, the original PR-EOS neglects capillary forces, resulting in no distinction between phase pressures. Consequently, the formulations for fugacity and compressibility factor must be modified to incorporate capillary effects and differentiate phase pressures. Referring to PR-EOS [12], the compressibility factor equation incorporating capillary forces is given as:
Z α 3 ( 1 B α ) Z α 2 + ( A α 3 B α 2 2 B α ) Z α ( A α B α B α 2 B α 3 ) = 0
where α denotes the phase, with l representing the gas phase and v the liquid phase. Z α is the phase compressibility factors; Aα and Bα are the PR-EOS parameters for the liquid and gas phases, with the specific expressions given as:
A α = a p α R 2 T 2       B α = b p α R T
According to Gibbs free energy principles [12], the smallest real root of Z l represents the liquid-phase compressibility factor, while the largest real root of Z v corresponds to the gas-phase compressibility factor. Using these factors, fugacity and fugacity coefficients can be determined. At phase equilibrium, the fugacity of the gas phase equals that of the liquid phase:
f l i ( T , P l , x i ) = f v i ( T , P v , y i )
where f l / v i are the fugacities of component i in the gas and liquid phases, respectively.
For a gas-liquid two-phase system, fugacity is defined as:
f = φ p
where φ is the fugacity coefficient, which is expressed as follows [12]:
l n φ = Z 1 ln Z + 1 R T V R T V p d V
Based on Chen’s research [52], the modified fugacity coefficients for the gas and liquid phases, incorporating capillary forces, are expressed as:
ln φ v i = b i b ( Z v 1 ) ln ( Z v B v ) Z v 2 2 B v 2 j = 1 N y j a i j a b i b ln Z v + ( 1 + 2 ) B v Z v + ( 1 2 ) B v ln φ l i = b i b ( Z l 1 ) ln ( Z l B l ) Z l 2 2 B l 2 j = 1 N x j a i j a b i b ln Z l + ( 1 + 2 ) B l Z l + ( 1 2 ) B l
where a and b represent the PR-EOS parameters for the mixture, while ai and bi denote those for pure component i. The calculation methods for these parameters are detailed in Reference [12].

2.2. Multiscale Compositional Model

2.2.1. Basis Model

Based on the compositional model, the following assumptions are made:
  • The reservoir consists of three phases: oil (denoted by subscript l), gas (subscript v), and water (subscript w).
  • The component model considers these three flowing phases. The water phase contains only the water component and does not achieve phase equilibrium with the other phases, whereas phase equilibrium is established among m components in the oil and gas phases.
  • Fluid flow follows Darcy’s law.
  • The reservoir is isothermal.
  • Phase equilibrium is achieved instantaneously.
Under these assumptions, the compositional control equations are formulated, beginning with the continuity equations for components in the oil and gas phases [52].
t ϕ ( ρ l S l X i + ρ v S v Y i ) + ( ρ l X i v l + ρ v Y i v v ) q i = 0
where ϕ represents porosity, ρ α and S α denote the density and saturation of phase α , respectively, l/v corresponds to the oil and gas phases, in accordance with phase equilibrium calculation theory.
The continuity equation for the water component is formulated as follows:
t ( ϕ ρ w S w ) + ( ρ w v w ) q w = 0
Phase velocities are determined according to multiphase Darcy’s law.
v α = K λ α ( p α ρ α g Δ z ) , λ α = k r α μ α
where K denotes the absolute positive-definite permeability tensor, while krα, μα, λα represent the relative permeability, viscosity, and mobility of phase α, respectively.
For each component, the source and sink terms in the continuity equation, along with the well control equations, follow the standard Peaceman model [53]. Additionally, the compositional model accounts for component transfer between the oil and gas phases, incorporating phase equilibrium equations to ensure accurate phase behavior representation
f l i f v i = 0
In addition, the constraints must also be considered:
S w + S v + S l = 1 , S α [ 0 , 1 ] α
i = 1 m x i = 1 , i = 1 m y i = 1 , x i , y i [ 0 , 1 ] i { 1 , , m }
By combining Equations (11) to (16), the governing equations are reformulated into residual forms. The Newton-Raphson iteration method is then applied to solve these equations efficiently.
J ε Δ ε = r ε , J ε = r ε ε , r ε = R ε W ε F ε C ε
where J ε represents the Jacobian matrix of the system, and r ε denotes the residual term, consisting of the residuals from the continuity equations ( R ε ), the well control equations ( W ε )—with W ε constructed based on the Peaceman model [53], fugacity equations ( F ε ), and constraints ( C ε ), as defined in Equations (15) and (16).
While the above equations can be solved directly in compositional numerical simulation, the computational efficiency remains low. To enhance efficiency, the Schur complement reduction method [54] is employed to decrease the degrees of freedom of the matrix. By partitioning the variables ( ε ) into primary variables (u) and eliminable variables (v), and reordering the Jacobian matrix accordingly, the system can be expressed as:
U u U v V u V v Δ u Δ v = r u r v
By reordering the system in Equation (17), the primary variable u is placed first, along with its corresponding continuity and well control equations U. Similarly, v represents the secondary variable, while V includes the fugacity equations and constraints associated with v. By applying block Gaussian elimination, a reduced system containing only the primary variables can be derived, effectively reducing the computational complexity
J ˜ u Δ u = r ˜ u , J ˜ u = U u U v V v 1 V u , r ˜ u = r u U v V v 1 r v
The increments of the secondary variables are updated using the increments of the primary variables, ensuring consistency between the two sets of variables.
Δ v = V v 1 r v V u Δ u

2.2.2. Sequential Formulation

To facilitate the use of multiscale formulations later, the sequential formulation is employed to split the entire system of equations into pressure equations and transport equations.
Various methods exist to construct the sequential formulation pressure equation based on reservoir numerical models. Using the linear algebraic approach by Coats [54], in Equation (19), considering the physical significance of different elements in the system, the term U corresponding to u represents the accumulation term containing the time derivative, while V corresponding to v includes the remaining constraint terms, excluding the accumulation term. Consequently, the Jacobian matrix J ε and its Schur complement reduced matrix J ˜ u can be reformulated to facilitate the explanation of subsequent transformations
J ε = A C u A C v ϑ u ϑ v , J ˜ u = A C u A C v ϑ v 1 ϑ u .
where A C represents the accumulation terms in the system of equations, ϑ represents the constraint equation block, and u and v represent the primary and secondary unknowns, respectively.
The purpose of the sequential formulation is to decouple the pressure variable from other variables. Referring to the derivation process of the sequential formulation in the black oil model, it achieves this by using a linear combination of the accumulated phases, which has zero derivatives with respect to non-pressure variables. Thus, the pressure variable can be decoupled from the non-pressure variables. Drawing on this concept, from a linear algebra perspective, we construct a linear combination coefficient vector—the weight factor w, which is formulated as follows:
( J ˜ u T ) 1 w = e p , ( e p ) i = 1 if   u i   is   a   pressure   variable 0 otherwise
Mathematically, the weight factor w aligns with the derivation of the sequential formulation in the black oil model, as both eliminate non-pressure variables through weighted summation. Thus, by applying w, a linear algebraic transformation can be performed on Equation (19) to derive the pressure equation. The pressure equation is derived based on Equations (19) and (22):
J ˜ p Δ u p = r ˜ p , J ˜ p R n c × n c , r ˜ p R n c .
The coefficients and the right-hand side of Equation (23) are constructed through the following expressions:
( r ˜ p ) j = i = 0 m 1 w j + i n c ( r ˜ u ) j + i n c , ( J ˜ p ) j , k = i = 0 m 1 w j + i n c ( J ˜ u ) j + i n c , k + w j + i n c p k ( r ˜ u ) j + i n c .
The pressure increment is obtained by solving the above equations. Subsequently, the transport equations are solved using the computed pressure increment.
Once the pressure Equation (23) has converged, the Darcy velocities for different phases are determined using the multiphase Darcy equation (Equation (13)). The total volumetric Darcy velocity is defined as:
v t = v w + v l + v v
Based on Equation (25), the integrated fractional flow is as follows [55]:
v α = λ α β { l , v , w } λ β v t + K β { l , v , w } λ β ( G α G β ) , G α = g Δ z ρ α + p c α .
where K represents the absolute positive-definite permeability tensor, while λα denotes the mobility of phase α, G accounting for the effects of gravity and capillary forces.
By substituting Equation (26) into the conservation Equations (11) and (12), the transport equations are derived, allowing the transport variables (components and saturations) to be solved.

2.2.3. Improved Multiscale Formulation

The multiscale formulation primarily targets the pressure Equation (23) and is written in the following form:
J ˜ p Δ p = r ˜ p
Since Equation (27) exhibits significant coefficient variations and strong coupling between adjacent cells, direct solution is inefficient. Instead, a multiscale expansion is employed to approximate pressure, enabling efficient computation on a coarser grid.
Following the algebraic multigrid approach, prolongation operator P and restriction operator R are defined to relate the fine-scale and coarse-scale pressure increments [44]. Fine-scale information is mapped to the coarse scale using a restriction operator. Solving on the coarse scale reduces the number of iterations due to low-frequency errors on the fine scale. The coarse-scale pressure increment is then mapped back to the fine-scale using a prolongation operator, approximating the fine-scale exact pressure increment:
Δ p Δ p f = P Δ p c
Using Equations (28) and (27), the solution of the fine-scale pressure increment is reformulated as solving the coarse-scale pressure increment, enhancing computational efficiency.
( R J ˜ p P ) Δ p c = J ˜ c Δ p c = R r ˜ p = r ˜ c
The construction of the operator format is key to achieving a multiscale solution. The restriction operator maps physical quantities from the fine-scale grid to the coarse-scale grid, with a common approach being the control volume restriction operator.
R c ν j i = 1 ,   if   x i Ω j b 0 ,   otherwise
Zhou and Lunati et al. [44,45] employed this restriction operator in the multiscale finite volume method, demonstrating its ability to ensure mass conservation. Wang et al. [56] further proposed a Galerkin restriction operator to improve convergence rates. Wang et al. have combined these approaches, applying the Galerkin restriction operator in the early stages for faster convergence and the control volume restriction operator later to maintain mass conservation [56]. This hybrid strategy is also applicable to the multiscale solution method proposed in this study. Møyner and Li et al. conducted numerical experiments on various restriction operators in multiscale formulations using different test models. Their findings indicate that in the original multiscale formulation, the choice of restriction operator influences the convergence of the multiscale method. However, when an iterative scheme is used to construct the prolongation operator, both types of restriction operators demonstrate identical convergence performance [57]. Therefore, this study adopts the control volume restriction operator.
The prolongation operator extends physical quantities from the coarse-scale grid to the fine-scale grid. A standard approach for constructing the multiscale prolongation operator P involves solving a series of local flow problems numerically. Unlike the classical MSFV method, this study follows Møyner and Li [44,46] by constructing P using an iterative scheme, eliminating the need for complex boundary conditions when setting up flow problems. This approach addresses the challenge of constructing high-quality basis functions on unstructured grids. The iterative format is as follows:
P n + 1 = P n M ( ω D s 1 J P n )
In Equation (31) D s represents the constructed smoother, a diagonal matrix composed of the diagonal elements of the matrix J ˜ p . The smoother is used to compute increments and progressively refine them. M serves as a restrictor, confining the smoothing of basis functions to a support region centered on the coarse-scale grid. The boundary of this region is defined by the eight surrounding grid center points, ensuring the accuracy of the prolongation operator.

2.3. Implementation of the Multiscale Compositional Simulator

The model consists of Equations (11)–(16), with phase equilibrium calculations governed by Equations (1), (5), (7), and (10), and the multiscale formulation described by Equation (29). The implementation is carried out using the MATLAB Reservoir Simulation Toolbox (MRST) [58]. This study refines the MRST-based compositional model by modifying the phase equation model class functions in accordance with the equations in Section 2.1. Additionally, the tool functions in the sequential formulation module are enhanced. A compositional pressure equation class and a transport equation class, accounting for capillary effects, are developed based on the sequential model base class. Finally, in line with the multiscale formulation, the local solution of the prolongation operator in the multiscale volume method is revised to an iterative solution. Within each nonlinear iteration, flash calculations incorporating capillary effects are first performed. The resulting component compositions are then integrated into the pressure and transport equations. The pressure equation is solved using a multiscale formulation for linear iterations. Upon convergence of the iteration error, pressure increments and secondary variables are transferred to the transport equation for subsequent calculations. The implementation process is illustrated in Figure 1.

3. Results and Discussion

3.1. Simulation Validation

Common commercial simulators still employ the traditional PR-EOS or SRK-EOS, without considering the effects of nanopore constraints on phase behavior [49]. For example, the CMG-GEM simulator (Version 2022) uses a hysteretic relative permeability approximation to simulate changes in flow caused by compositional variations. Therefore, this section adopts a model degradation approach, comparing the multiscale compositional model with the ECLIPSE E300 simulator (Version 2022) to validate the accuracy of the multiscale calculations in the current model. The validation employed a one-dimensional, two-phase model (1000 × 1 × 0.1 m), with its 3D view depicted in Figure 2. The parameters for the validation reservoir model are presented in Table 1.
Carbon dioxide was injected into a mixture of CO2, methane, and decane (10% CO2 + 30% C1 + 60% C10). The model comprises 1000 grid cells, with an initial reservoir temperature of 150 °C and an initial pressure of 7.5 MPa. The initial component properties are summarized in Table 2. The relative permeability curves are shown in Figure 3.
To ensure consistent time step selection across simulators, the problem is discretized into a large number of time steps. Given the relatively low pressure, phase behavior exhibits strong pressure dependence, with all components present in both phases. Since the well operates under bottom-hole pressure control, the injected fluid volume relies on precise calculations of fluid properties and densities within the reservoir. To validate the simulator (MS)’s accuracy, the predicted cumulative oil production curve over 20 years (Figure 4) and the total molar fractions of different components at various stages were compared with results from ECLIPSE E300 (Figure 5). The minimal discrepancies between the two sets of results confirm the reliability of the proposed simulator.
Additionally, due to the large number of time steps in the model calculations, the runtime logs for the first 50-time steps were extracted to compare the computational efficiency of the traditional fully implicit (FI) method with the multiscale method proposed in this study. As shown in Figure 6, the multiscale method (MS) demonstrates a clear advantage over the traditional fully implicit method (FI) in terms of iterative solution speed, significantly improving the model’s computational efficiency. The total runtime for FI and MS was measured, revealing that the runtime for MS was approximately 254 s, compared to about 408 s for FI. This indicates a 40% improvement in the computational efficiency of MS.

3.2. Sensitivity Analysis

The accuracy and efficiency of the multiscale compositional numerical simulator were validated in the previous section. This section focuses on the impact of capillary forces, emphasizing the proposed simulator’s importance in tight oil reservoir development. It is divided into two main parts: the first part examines the influence of capillary forces on phase equilibrium calculations, while the second part utilizes the previously established multiscale numerical simulator, incorporating capillary forces, to analyze a typical tight oil reservoir model. By referencing actual tight oil reservoir parameters, this part explores the effect of capillary forces on oil recovery.

3.2.1. The Influence of Capillary Forces on Phase Equilibrium Calculations

Referring to the mixture component properties of the Fifth SPE Comparative Solution Project [59], the PR-EOS was modified to incorporate capillary forces. The study compared the modified PR-EOS with capillary constraints (Equation (5) for pore radii of 15 nm, 30 nm, and 50 nm against the conventional PR-EOS without capillary constraints. Figure 7 and Figure 8 present the phase diagram and the iteration count, respectively.
As shown in Figure 7, both the bubble point and dew point lines decrease with the reduction of pore radius. The smaller the pore radius, the more pronounced the downward trend of both lines. During the pressure drop process, the capillary effect becomes more significant with decreasing pore radius, leading to a reduction in both the bubble point and dew point. As a result, the liquid phase fraction increases while the gas phase fraction decreases. Near the critical point, the calculation results for different pore radii converge, which is attributed to the interfacial tension approaching zero as the gas-liquid interface disappears, thus nullifying the capillary force. Therefore, capillary forces do not affect fluid phase behavior near the critical point. For larger pore radii, such as 50 nm and above, the capillary effect is negligible, and the phase diagram closely resembles that without considering capillary forces.
As shown in Figure 8, the number of iterations for phase equilibrium calculations increases with decreasing pore radius, particularly near the dew point line. This indicates that the PR-EOS incorporating capillary forces introduces greater computational overhead, highlighting the need for a multiscale fast solution. As described in Section 2.1, the conventional PR-EOS does not consider capillary pressure effects, leading to consistent phase pressures across different phases. However, when capillary pressure effects are incorporated, phase pressures differ, with smaller pores showing more significant disparities. This physical phenomenon increases the number of iterations required during the phase equilibrium solution process.

3.2.2. Key Findings on the Impact of Capillary Forces on Tight Oil Production

This section uses fluid data from B. Mallison et al. [60], including mixture composition, basic properties, and binary interaction parameters. A simple closed-external-boundary cubic reservoir model, incorporating 5 vertical wells and a hydraulic fracture (as detailed in Table 3), was created with 5000 grid cells. The grid size settings are based on Ma’s research [49].
The initial component properties are summarized in Table 4.
As shown in Figure 9, the five vertical wells form a five-spot pattern, with four injection wells at the reservoir edges and one production well at the center. A black line indicates the hydraulic fracture, and the target grid is marked with a red star. The production well operates at a constant pressure of 10 MPa, while the injection wells inject CO2 at a constant rate of 300 m3/day. The injection and production period lasts 10 years. The relative permeability curves used in the simulation refer to the core displacement test report of a certain block in the Changqing Oilfield, China. The specific data can be found in Figure 10.
This section presents four scenarios:
Base Case: In gas-phase equilibrium calculations, existing numerical simulators neglect the impact of confinement on fluid phase behavior in nanopores. Consequently, capillary pressure effects and shifts in critical properties are excluded.
Case 1: Capillary forces are calculated using the Young-Laplace equation with a pore radius of 20 nm.
Case 2: Capillary forces are calculated using the Young-Laplace equation with a pore radius of 35 nm.
Case 3: Capillary forces are calculated using the Young-Laplace equation with a pore radius of 50 nm.
The proposed multiscale compositional simulator is used to analyze the influence of capillary forces on cumulative production, component distribution, and reservoir saturation.
Based on the results of the four scenarios, the impact of capillary forces on cumulative production was analyzed. As shown in Figure 11, including capillary forces (Case 1–3) generally leads to higher cumulative oil production compared to the scenario without capillary effects (Base Case). The smaller the pore radius, the more pronounced the capillary effect, resulting in a greater increase in cumulative oil production. This underscores the importance of considering capillary forces in tight oil reservoir development. It is important to note that this section primarily addresses how capillary pressure influences phase behavior and, consequently, oil recovery efficiency. However, in the actual development of tight oil reservoirs, various complex mechanisms and operating conditions must be considered. For instance, capillary trapping effects play a role during multiphase flow in tight oil reservoirs. Therefore, capillary pressure should not be viewed as the sole key mechanism for improving oil recovery.
Figure 12 shows that capillary forces (Case 1–3) inhibit the increase in gas-phase components. This is due to a decrease in the bubble point and dew point lines caused by capillary effects. Smaller pore radii result in a more significant decrease, which leads to an increase in the liquid phase and a relative reduction in the gas phase.
The impact of capillary forces on component distribution was analyzed, focusing on the cell marked by a red star in Figure 9. The molar fractions of components in this cell under the Base Case and Case 1 were compared. As shown in Figure 13, the model categorizes components into light (N2/CH4, CO2) and heavy (C2–5, C6–13, C14–24, C25–80) groups. Compared to the model incorporating capillary forces (Case 1), the base-case model (excluding capillary forces) yields a higher proportion of light components and retains more heavy components in the reservoir. These findings suggest that accounting for capillary forces in nanopores may enhance recovery efficiency by modifying phase behavior.
The oil saturation at different time steps for various scenarios is compared in Figure 14. The bubble point pressure is higher in the scenario without capillary effects, leading to greater gas stripping. As a result, this scenario (Base Case) exhibits lower oil saturation and higher gas saturation compared to the scenario with capillary forces (Case 1–3).
Additionally, oil saturation varies with pore radius and capillary force magnitude. The analysis shifts to the characteristic scale to further investigate, focusing on the cell marked by a red star in Figure 9. Figure 15 shows that gas saturation is higher in the scenario without capillary effects (Base Case), and this trend becomes more pronounced with a stronger capillary effect. Eventually, saturation levels converge as CO2 breakthrough occurs.

4. Conclusions

This paper presents a multiscale compositional numerical simulation method that incorporates capillary forces in phase equilibrium calculations. A one-dimensional model is used for validation, demonstrating the method’s accuracy and advancement in analyzing the impact of capillary forces on phase equilibrium and tight oil recovery.
  • Phase behavior in tight oil reservoirs is influenced by pore size and capillary forces, which traditional equations of state fail to capture in nanopores. This paper introduces capillary forces into traditional equations of state and compositional numerical simulations, improving phase equilibrium calculations using MRST.
  • To enhance iterative solution efficiency, a sequential solution format is employed, leveraging the parallel advantages of the improved multiscale finite volume method for pressure equation resolution. Numerical validation with a one-dimensional model confirms that this method enhances solution efficiency without compromising accuracy.
  • Using the proposed multiscale compositional simulation method, the impact of capillary forces on phase equilibrium and tight oil recovery is analyzed. Sensitivity analysis shows that smaller pore radii lead to more pronounced capillary effects, with a decrease in bubble point and dew point lines, and an increase in calculation iterations. For tight oil recovery, capillary forces generally result in higher cumulative oil production and improved recovery rates for heavy components. Capillary effects lower bubble point pressure, suppress gas stripping and increase oil saturation while decreasing gas saturation.
  • Compared to traditional numerical simulators, the proposed multiscale compositional method, which accounts for capillary forces, better characterizes phase behavior in tight oil reservoirs. This approach is significant for PVT simulation, development of numerical simulations, and forecasting development strategies in tight oil reservoirs.
In summary, the proposed model advances numerical simulation research for tight oil reservoirs, though several challenges remain. First, this study examines the reservoir-scale impact of capillary forces on phase behavior under an instantaneous equilibrium assumption, without addressing nanopore kinetic limitations and complex flow mechanisms. Second, large-scale physical experiments could help overcome the validation constraints of commercial simulators. Finally, although MATLAB Reservoir Simulation Toolbox (MRST 2024b) offer significant community and ecosystem potential, incorporating additional open-source high-performance components may further promote the development of large-scale 3D tight oil reservoir simulators.

Author Contributions

Conceptualization, J.W. and S.W.; methodology, J.W.; software, W.C.; validation, L.W., J.W., and S.W.; formal analysis, R.Z.; investigation, W.C. and H.Y.; resources, Q.S.; data curation, L.W. and H.Y.; writing—original draft preparation, J.W.; writing—review and editing, Q.S.; visualization, R.Z.; supervision, S.W.; project administration, S.W.; funding acquisition, S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 5237405 and U24B2037), the Science and Technology Project of China National Petroleum Corporation (2021ZG12, 2023DJ8405).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to sincerely thank those who have contributed to this research.

Conflicts of Interest

Li Wu was employed by PetroChina Company Limited. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Zhu, W.; Yue, M.; Liu, Y.; Liu, K.; Song, Z. Research Progress on the Development Theory of Tight Oil Reservoirs in China. Chin. J. Eng. 2019, 41, 1103–1114. [Google Scholar] [CrossRef]
  2. Zhu, R.; Zou, C.; Wu, S.; Yang, Z.; Mao, Z.; Yang, H.; Fan, C.; Hui, X.; Cui, J.; Su, L.; et al. Formation Mechanism and Enrichment Patterns of Continental Tight Oil in China. Oil Gas Geol. 2019, 40, 1168–1184. [Google Scholar]
  3. Sun, L.; Zou, C.; Jia, A.; Wei, Y.; Zhu, R.; Wu, S.; Guo, Z. Characteristics and Development Trends of China’s Tight Oil and Gas. Pet. Explor. Dev. 2019, 46, 1015–1026. [Google Scholar] [CrossRef]
  4. Lee, E.Y.; Fathy, D.; Xiang, X.; Spahić, D.; Ahmed, M.S.; Fathi, E.; Sami, M. Middle Miocene syn-rift sequence on the central Gulf of Suez, Egypt: Depositional environment, diagenesis, and their roles in reservoir quality. Mar. Pet. Geol. 2025, 174, 107305. [Google Scholar] [CrossRef]
  5. Jia, C.; Zheng, M.; Zhang, Y. Unconventional hydrocarbon resources in China and the prospect of exploration and development. Pet. Explor. Dev. 2012, 39, 139–146. [Google Scholar] [CrossRef]
  6. Li, J.; Zheng, M.; Zhang, G.; Yang, T.; Wang, S.; Dong, D.; Wu, X.; Qu, H.; Chen, X. Potential and prospects of conventional and unconventional natural gas resource in China. Acta Pet. Sin. 2012, 33, 89–98. [Google Scholar]
  7. Loucks, R.G.; Reed, R.M.; Ruppel, S.C.; Jarvie, D.M. Morphology, Genesis, and Distribution of Nanometer-Scale Pores in Siliceous Mudstones of the Mississippian Barnett Shale. J. Sediment. Res. 2009, 79, 848–861. [Google Scholar] [CrossRef]
  8. Sigal, R.F. Pore-Size Distributions for Organic-Shale-Reservoir Rocks From Nuclear-Magnetic-Resonance Spectra Combined With Adsorption Measurements. SPE J. 2015, 20, 824–830. [Google Scholar] [CrossRef]
  9. Shi, W.; Liu, X.; Gao, M.; Tao, L.; Bai, J.; Zhu, Q. Pressuredrop Response Characteristics for Multi-Injection Well Interfered Vertical Well in Heterogeneous Fractured Anticline Reservoirs. J. Energy Resour. Technol. 2023, 145, 092902. [Google Scholar] [CrossRef]
  10. Jing, W.; Zhang, L.; Zhang, Y.; Memon, B.S.; Li, A.; Zhong, J.; Sun, H.; Yang, Y.; Cheng, Y.; Yao, J. Phase behavior of gas condensate in fractured-vuggy porous media based on microfluidic technology and real-time computed tomography scanning. Phys. Fluids 2023, 35, 122002. [Google Scholar] [CrossRef]
  11. Zuo, M.; Chen, H.; Qi, X.; Liu, X.; Xu, C.; Yu, H.; Brahim, M.S.; Wu, Y.C.; Liu, H. Effects of CO2 injection volume and formation of in-situ new phase on oil phase behavior during CO2 injection for enhanced oil recovery (EOR) in tight oil reservoirs. Chem. Eng. J. 2022, 452, 139454. [Google Scholar] [CrossRef]
  12. Peng, D.-Y.; Robinson, D.B. A New Two-Constant Equation of State. Ind. Eng. Chem. Fundam. 1976, 15, 59–64. [Google Scholar] [CrossRef]
  13. Melhem, G.A.; Saini, R.; Goodwin, B.M. A modified Peng-Robinson equation of state. Fluid Phase Equilibria 1989, 47, 189–237. [Google Scholar] [CrossRef]
  14. Stryjek, R.; Vera, J.H. PRSV—An improved Peng-Robinson equation of state with new mixing rules for strongly nonideal mixtures. Can. J. Chem. Eng. 1986, 64, 334–340. [Google Scholar] [CrossRef]
  15. Hu, J.; Du, C. Phase Behavior Study in Petroleum Systems. Inn. Mong. Petrochem. 2005, 87–88. [Google Scholar]
  16. Barsotti, E.; Tan, S.P.; Saraji, S.; Piri, M.; Chen, J.-H. A review on capillary condensation in nanoporous media: Implications for hydrocarbon recovery from tight reservoirs. Fuel 2016, 184, 344–361. [Google Scholar] [CrossRef]
  17. Jin, Z.; Firoozabadi, A. Thermodynamic Modeling of Phase Behavior in Shale Media. SPE J. 2016, 21, 190–207. [Google Scholar] [CrossRef]
  18. Sun, H.; Li, H.A. A New Three-Phase Flash Algorithm Considering Capillary Pressure in a Confined Space. Chem. Eng. Sci. 2019, 193, 346–363. [Google Scholar] [CrossRef]
  19. Ribeiro Carrott, M.M.L.; Candeias, A.J.E.; Carrott, P.J.M.; Ravikovitch, P.I.; Neimark, A.V.; Sequeira, A.D. Adsorption of nitrogen, neopentane, n-hexane, benzene and methanol for the evaluation of pore sizes in silica grades of MCM-41. Microporous Mesoporous Mater. 2001, 47, 323–337. [Google Scholar] [CrossRef]
  20. Qiao, S.Z.; Bhatia, S.K.; Nicholson, D. Study of Hexane Adsorption in Nanoporous MCM-41 Silica. Langmuir 2004, 20, 389–395. [Google Scholar] [CrossRef]
  21. Morishige, K.; Kawano, K.; Hayashigi, T. Adsorption Isotherm and Freezing of Kr in a Single Cylindrical Pore. J. Phys. Chem. B 2000, 104, 10298–10303. [Google Scholar] [CrossRef]
  22. Russo, P.A.; Carrott, M.M.L.R.; Carrott, P.J.M. Hydrocarbons adsorption on templated mesoporous materials: Effect of the pore size, geometry and surface chemistry. New J. Chem. 2011, 35, 407–416. [Google Scholar] [CrossRef]
  23. Yun, J.-H.; Düren, T.; Keil, F.J.; Seaton, N.A. Adsorption of Methane, Ethane, and Their Binary Mixtures on MCM-41:  Experimental Evaluation of Methods for the Prediction of Adsorption Equilibrium. Langmuir 2002, 18, 2693–2701. [Google Scholar] [CrossRef]
  24. Zeigermann, P.; Dvoyashkin, M.; Valiullin, R.; Kärger, J. Assessing the pore critical point of the confined fluid by diffusion measurement. Diffus. Fundam. 2009, 41, 1–2. [Google Scholar] [CrossRef]
  25. Ally, J.; Molla, S.; Mostowfi, F. Condensation in Nanoporous Packed Beds. Langmuir 2016, 32, 4494–4499. [Google Scholar] [CrossRef]
  26. Alfi, M.; Nasrabadi, H.; Banerjee, D. Experimental investigation of confinement effect on phase behavior of hexane, heptane and octane using lab-on-a-chip technology. Fluid Phase Equilibria 2016, 423, 25–33. [Google Scholar] [CrossRef]
  27. Zhong, J.; Riordon, J.; Zandavi, S.H.; Xu, Y.; Persad, A.H.; Mostowfi, F.; Sinton, D. Capillary Condensation in 8 nm Deep Channels. J. Phys. Chem. Lett. 2018, 9, 497–503. [Google Scholar] [CrossRef]
  28. Brusilovsky, A.I. Mathematical Simulation of Phase Behavior of Natural Multicomponent Systems at High Pressures With an Equation of State. SPE Reserv. Eng. 1992, 7, 117–122. [Google Scholar] [CrossRef]
  29. Pang, J.; Zuo, J.Y.; Zhang, D.; Du, L. Impact of Porous Media on Saturation Pressures of Gas and Oil in Tight Reservoirs. In Proceedings of the SPE Canadian Unconventional Resources Conference, Calgary, AB, Canada, 30 October–1 November 2012. [Google Scholar]
  30. Teklu, T.W.; Alharthy, N.; Kazemi, H.; Yin, X.; Graves, R.M.; AlSumaiti, A.M. Phase Behavior and Minimum Miscibility Pressure in Nanopores. Spe Reserv. Eval. Eng. 2014, 17, 396–403. [Google Scholar] [CrossRef]
  31. Qi, Z.; Liang, B.; Deng, R.; Du, Z.; Wang, S.; Zhao, W. Phase Behavior Study in the Deep Gas-Condensate Reservoir with Low Permeability. In Proceedings of the EUROPEC/EAGE Conference and Exhibition, London, UK, 11–14 June 2007. [Google Scholar]
  32. Nojabaei, B.; Johns, R.T.; Chu, L. Effect of Capillary Pressure on Phase Behavior in Tight Rocks and Shales. SPE Reserv. Eval. Eng. 2013, 16, 281–289. [Google Scholar] [CrossRef]
  33. Yuan, Z.; Yuan, D.; Yun, Z.; Dongli, Z. The Impact of Capillary Force Effects in Nanopores on the Productivity of Tight Oil Reservoirs. J. Zhejiang Univ. Sci. Technol. 2017, 29, 328–333. [Google Scholar]
  34. Li, Y.; Li, X.; Teng, S.; Xu, D. Equilibrium Calculation Considering the Interaction between Pore-Throats and Fluid Molecules, as well as Capillary Forces. Acta Pet. Sin. 2015, 36, 511–515. [Google Scholar]
  35. Zhang, Y.; Di, Y.; Liu, P.; Li, W. Simulation of tight fluid flow with the consideration of capillarity and stress-change effect. Sci. Rep. 2019, 9, 5324. [Google Scholar] [CrossRef]
  36. Wang, Y.; Yan, B.; Killough, J. Compositional Modeling of Tight Oil Using Dynamic Nanopore Properties. In Proceedings of the SPE Annual Technical Conference and Exhibition, New Orleans, LA, USA, 30 September–2 October 2013. [Google Scholar]
  37. Khoshghadam, M.; Khanal, A.; Rabinejadganji, N.; Lee, W.J. How to Model and Improve Our Understanding of Liquid-Rich Shale Reservoirs with Complex Organic/Inorganic Pore Network. In Proceedings of the SPE/AAPG/SEG Unconventional Resources Technology Conference, San Antonio, TX, USA, 1–3 August 2016. [Google Scholar]
  38. Yan, B.; Wang, Y.; Killough, J.E. A fully compositional model considering the effect of nanopores in tight oil reservoirs. J. Pet. Sci. Eng. 2017, 152, 675–682. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Di, Y.; Yu, W.; Sepehrnoori, K. A Comprehensive Model for Investigation of Carbon Dioxide Enhanced Oil Recovery With Nanopore Confinement in the Bakken Tight Oil Reservoir. SPE Reserv. Eval. Eng. 2018, 22, 122–136. [Google Scholar] [CrossRef]
  40. Hou, T.Y.; Wu, X.-h. A Multiscale Finite Element Method for Elliptic Problems in Composite Materials and Porous Media. J. Comput. Phys. 1997, 134, 169–189. [Google Scholar] [CrossRef]
  41. Babuska, I.; Osborn, J.E. Generalized Finite Element Methods: Their Performance and Their Relation to Mixed Methods. SIAM J. Numer. Anal. 1983, 20, 510–536. [Google Scholar] [CrossRef]
  42. Jenny, P.; Lee, S.H.; Tchelepi, H.A. Multi-scale finite-volume method for elliptic problems in subsurface flow simulation. J. Comput. Phys. 2003, 187, 47–67. [Google Scholar] [CrossRef]
  43. Møyner, O.; Lie, K.-A. A Multiscale Restriction-Smoothed Basis Method for Compressible Black-Oil Models. SPE J. 2016, 21, 2079–2096. [Google Scholar] [CrossRef]
  44. Zhou, H.; Tchelepi, H.A.A. Two-Stage Algebraic Multiscale Linear Solver for Highly Heterogeneous Reservoir Models. SPE J. 2012, 17, 523–539. [Google Scholar] [CrossRef]
  45. Lunati, I.; Tyagi, M.; Lee, S.H. An iterative multiscale finite volume algorithm converging to the exact solution. J. Comput. Phys. 2011, 230, 1849–1864. [Google Scholar] [CrossRef]
  46. Møyner, O.; Lie, K.-A. A multiscale restriction-smoothed basis method for high contrast porous media represented on unstructured grids. J. Comput. Phys. 2016, 304, 46–71. [Google Scholar] [CrossRef]
  47. Wu, L.; Wang, J.; Jia, D.; Zhang, R.; Zhang, J.; Yan, Y.; Wang, S. A Multi-Scale Numerical Simulation Method Considering Anisotropic Relative Permeability. Processes 2024, 12, 2058. [Google Scholar] [CrossRef]
  48. Hu, X.; Hu, S.; Jin, F.; Huang, S. Physics of Petroleum Reservoirs; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar] [CrossRef]
  49. Ma, Y. Study on the Phase Behavior of Nano-Pore Fluid and Numerical Simulation Method in Tight Reservoir. Ph.D. Thesis, China University of Geosciences, Beijing, China, 2023. [Google Scholar]
  50. Ferguson, A. On a relation between surface tension and density. Trans. Faraday Soc. 1923, 19, 407–412. [Google Scholar] [CrossRef]
  51. Pedersen, K.S.; Christensen, P.L.; Shaikh, J.A. Phase Behavior of Petroleum Reservoir Fluids; CRC Press: Boca Raton, FL, USA, 2006. [Google Scholar] [CrossRef]
  52. Riviere, B. Computational methods for multiphase flows in porous media. Math. Comput. 2007, 76, 2253–2255. [Google Scholar] [CrossRef]
  53. Peaceman, D.W. Interpretation of Well-Block Pressures in Numerical Reservoir Simulation(includes associated paper 6988). Soc. Pet. Eng. J. 1978, 18, 183–194. [Google Scholar] [CrossRef]
  54. Coats, K.H. A Note on IMPES and Some IMPES-Based Simulation Models. SPE J. 2000, 5, 245–251. [Google Scholar] [CrossRef]
  55. Møyner, O.; Tchelepi, H.A. A Mass-Conservative Sequential Implicit Multiscale Method for Isothermal Equation-of-State Compositional Problems. SPE J. 2018, 23, 2376–2393. [Google Scholar] [CrossRef]
  56. Wang, Y.; Hajibeygi, H.; Tchelepi, H.A. Algebraic multiscale solver for flow in heterogeneous porous media. J. Comput. Phys. 2014, 259, 284–303. [Google Scholar] [CrossRef]
  57. Møyner, O. Construction of Multiscale Preconditioners on Stratigraphic Grids. In Proceedings of the CMOR XIV—14th European Conference on the Mathematics of Oil Recovery Catania, Sicily, Italy, 8–11 September 2014. [Google Scholar]
  58. Lie, K.-A. An Introduction to Reservoir Simulation Using MATLAB/GNU Octave: User Guide for the MATLAB Reservoir Simulation Toolbox (MRST); Cambridge University Press: Cambridge, UK, 2019. [Google Scholar]
  59. Killough, J.E.; Kossack, C.A. Fifth Comparative Solution Project: Evaluation of Miscible Flood Simulators. In Proceedings of the SPE Symposium on Reservoir Simulation, San Antonio, TX, USA, 1–4 February 1987. [Google Scholar]
  60. Mallison, B.T.; Gerritsen, M.G.; Jessen, K.; Orr, F.M. High-Order Upwind Schemes for Two-Phase, Multicomponent Flow. SPE J. 2005, 10, 297–311. [Google Scholar] [CrossRef]
Figure 1. Flowchart of the simulation process.
Figure 1. Flowchart of the simulation process.
Applsci 15 03082 g001
Figure 2. Schematic diagram of the 3D model.
Figure 2. Schematic diagram of the 3D model.
Applsci 15 03082 g002
Figure 3. Relative permeability curves.
Figure 3. Relative permeability curves.
Applsci 15 03082 g003
Figure 4. Comparison of Cumulative Oil Production between Different Simulators for the One-Dimensional Validation Model.
Figure 4. Comparison of Cumulative Oil Production between Different Simulators for the One-Dimensional Validation Model.
Applsci 15 03082 g004
Figure 5. Comparison of Component Distribution between Different Simulators at Various Production Times (In the legend, MS denotes the multiscale model presented in this paper).
Figure 5. Comparison of Component Distribution between Different Simulators at Various Production Times (In the legend, MS denotes the multiscale model presented in this paper).
Applsci 15 03082 g005
Figure 6. Comparison of Runtime for the First 50 Computational Steps Using Different Solution Methods.
Figure 6. Comparison of Runtime for the First 50 Computational Steps Using Different Solution Methods.
Applsci 15 03082 g006
Figure 7. Phase diagrams corresponding to different pore radii.
Figure 7. Phase diagrams corresponding to different pore radii.
Applsci 15 03082 g007
Figure 8. Contour map of iteration counts corresponding to different pore radii.
Figure 8. Contour map of iteration counts corresponding to different pore radii.
Applsci 15 03082 g008
Figure 9. Hydraulic fracturing five-spot well pattern model (The red bars indicate well locations. I1–I4 represent four injection wells, and P1 is the production well. Black lines show hydraulic fractures, and the red star marks the target grid).
Figure 9. Hydraulic fracturing five-spot well pattern model (The red bars indicate well locations. I1–I4 represent four injection wells, and P1 is the production well. Black lines show hydraulic fractures, and the red star marks the target grid).
Applsci 15 03082 g009
Figure 10. Relative permeability curves (The data comes from a core displacement test in the Changqing Oilfield).
Figure 10. Relative permeability curves (The data comes from a core displacement test in the Changqing Oilfield).
Applsci 15 03082 g010
Figure 11. Comparison of cumulative oil production across different scenarios.
Figure 11. Comparison of cumulative oil production across different scenarios.
Applsci 15 03082 g011
Figure 12. Comparison of gas-oil ratio (GOR) across different scenarios.
Figure 12. Comparison of gas-oil ratio (GOR) across different scenarios.
Applsci 15 03082 g012
Figure 13. Comparison of components between the base case and Case 1.
Figure 13. Comparison of components between the base case and Case 1.
Applsci 15 03082 g013
Figure 14. Oil saturation distribution at different stages across various scenarios.
Figure 14. Oil saturation distribution at different stages across various scenarios.
Applsci 15 03082 g014
Figure 15. Oil saturation distribution focusing on the cell marked by a red star in Figure 9 at different stages across various scenarios.
Figure 15. Oil saturation distribution focusing on the cell marked by a red star in Figure 9 at different stages across various scenarios.
Applsci 15 03082 g015
Table 1. Validation reservoir model parameters.
Table 1. Validation reservoir model parameters.
Model GridGrid Size
(m)
Initial Reservoir Temperature (°C)Initial Reservoir Pressure
(MPa)
Porosity
(%)
Permeability
(mD)
Injection Pressure
(MPa)
Production Pressure (MPa)Simulation Time
(Year)
1000 × 1 × 1 (coarse grid 100 × 1 × 1)1 × 1 × 0.11507.5251010520
Table 2. Component properties of the fluid mixture in the one-dimensional validation model.
Table 2. Component properties of the fluid mixture in the one-dimensional validation model.
ComponentPc (Bar)Tc (K)Vc (L/mol)Mol. WeightAcentric FactInitial Total Composition Values
C145.922190.6540.0986316.040.011420.3
C1021.03617.70.6098142.280.48840.6
CO273.733304.1280.09412440.223940.1
Table 3. Reservoir model parameters.
Table 3. Reservoir model parameters.
Model GridGrid Size
(m)
Initial Reservoir Temperature (°C)Initial Reservoir Pressure
(MPa)
Porosity
(%)
Matrix Permeability
(mD)
Fracture Permeability
(mD)
Initial Saturation (Water, Oil, and Gas)
(f)
The Bottom-Hole Pressure of the Producing Well (MPa)The Daily Gas Injection Volume of an Injection Well (m3/day)
25 × 25 × 840 × 40 × 51141581010000.38,
0.62, 0
10300
Table 4. Component properties of the fluid mixture.
Table 4. Component properties of the fluid mixture.
ComponentPc (Bar)Tc (K)Vc (L/mol)Mol. WeightAcentric FactInitial Total Composition Values
N2/CH445.8189.50.099716.15940.011420.463
CO273.733304.1280.09412440.223940.0164
C2–541.0387.60.0217145.57250.1670.2052
C6–1333.5597.50.0381211.7740.3860.19108
C14–2417.7698.50.0721424.88270.8080.08113
C25–8011.7875.00.0113648.1251.2310.04319
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

Wang, J.; Wu, L.; Sun, Q.; Zhang, R.; Chen, W.; Yang, H.; Wang, S. A Multiscale Compositional Numerical Study in Tight Oil Reservoir: Incorporating Capillary Forces in Phase Behavior Calculation. Appl. Sci. 2025, 15, 3082. https://doi.org/10.3390/app15063082

AMA Style

Wang J, Wu L, Sun Q, Zhang R, Chen W, Yang H, Wang S. A Multiscale Compositional Numerical Study in Tight Oil Reservoir: Incorporating Capillary Forces in Phase Behavior Calculation. Applied Sciences. 2025; 15(6):3082. https://doi.org/10.3390/app15063082

Chicago/Turabian Style

Wang, Junqiang, Li Wu, Qian Sun, Ruichao Zhang, Wenbin Chen, Haitong Yang, and Shuoliang Wang. 2025. "A Multiscale Compositional Numerical Study in Tight Oil Reservoir: Incorporating Capillary Forces in Phase Behavior Calculation" Applied Sciences 15, no. 6: 3082. https://doi.org/10.3390/app15063082

APA Style

Wang, J., Wu, L., Sun, Q., Zhang, R., Chen, W., Yang, H., & Wang, S. (2025). A Multiscale Compositional Numerical Study in Tight Oil Reservoir: Incorporating Capillary Forces in Phase Behavior Calculation. Applied Sciences, 15(6), 3082. https://doi.org/10.3390/app15063082

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