Next Article in Journal
Marbling and Meat Quality of Kazakh Finewool Purebred and Suffolk × Finewool Crossbred Sheep on an Intensive Fattening Diet
Previous Article in Journal
Performance-Oriented Conceptual Design of Fastener Joint Configurations for Aerospace Equipment
Previous Article in Special Issue
Microfluidic Investigation on the Diffusion Law of Nano Displacement Agent in Porous Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Simulation of CO2 Injection and Production in Shale Oil Reservoirs with Radial Borehole Fracturing

1
BGP Inc., China National Petroleum Corporation (CNPC), Zhuozhou 072750, China
2
Xinjiang Oilfield Company, PetroChina, Karamay 834000, China
3
Karamay Branch of State Key Laboratory of Petroleum Resources and Engineering, China University of Petroleum-Beiing at Karamay, Karamay 834000, China
*
Author to whom correspondence should be addressed.
Processes 2025, 13(9), 2873; https://doi.org/10.3390/pr13092873
Submission received: 20 June 2025 / Revised: 8 July 2025 / Accepted: 14 July 2025 / Published: 8 September 2025
(This article belongs to the Special Issue Advanced Strategies in Enhanced Oil Recovery: Theory and Technology)

Abstract

Shale oil is a vital strategic resource in China. Developing shale oil using CO2 not only enhances oil recovery but also contributes to achieving Chinese “dual carbon” goals. Given the challenges of insufficient number of fractures, inadequate vertical stimulation volume, and poor reservoir mobility associated with horizontal well fracturing, this study proposes a method for CO2 flooding based on radial borehole fracturing in a single well to achieve long-term carbon sequestration. To this end, a multi-component numerical model is built to analyze the production capacity of radial borehole fracturing. This study analyzed the impacts of non-Darcy flow, diffusion, and adsorption mechanisms on CO2 migration and sequestration. It also compared the applicability of continuous CO2 flooding and CO2 huff-and-puff under different matrix permeabilities. The results indicate that (1) CO2 flooding using radial borehole fracturing can achieve long-term oil production and carbon sequestration. (2) Under low permeability conditions, the liquid non-Darcy effect retards the flow of oil and CO2, while diffusion and adsorption facilitate CO2 sequestration in the reservoir. The impact on carbon sequestration is ranked as follows: non-Darcy effect > adsorption > diffusion. (3) High-permeability reservoirs are more suitable for carbon sequestration and should utilize continuous CO2 flooding. For low-permeability reservoirs (<0.001 mD), huff-and-puff should be employed to mobilize the reservoir around fractures and achieve carbon sequestration. The findings of this study are expected to provide new methods and a theoretical basis for efficient and economical carbon sequestration in shale oil reservoirs.

1. Introduction

The goal proposed by China to achieve peak carbon dioxide emissions by 2030 and strive for carbon neutrality by 2060 is a key measure in promoting ecological civilization construction. Carbon capture, utilization, and storage (CCUS) technology is widely regarded as a key solution to achieving large-scale greenhouse gas emission reductions and meeting long-term carbon neutrality targets. Over the past two decades, CCUS has been successfully demonstrated in multiple deep geological formations, including saline aquifers, depleted oil and gas reservoirs, and coal seams. Notably, the Mount Simon Sandstone Formation in the Illinois Basin (USA) has been a major site for CO2 injection at depths exceeding 2100–2200 m, where the thick, laterally continuous saline aquifer and overlying shale caprock provide favorable conditions for long-term storage [1]. In China, field-scale CO2 sequestration pilot projects have been conducted in the Ordos Basin and Junggar Basin, with injection depths reaching 2500 to over 3000 m [2]. These sites feature low-permeability formations and multi-layered caprock systems, making them suitable analogs for evaluating CO2 behavior in deeply buried, heterogeneous reservoirs.
These examples highlight the geological feasibility and practical success of deep CO2 sequestration under varying lithological and structural conditions, and they serve as important references for the reservoir setting and injection strategy adopted in this study [3,4].
With advances in geological understanding and technological progress, the focus of oil and gas exploration and development has gradually shifted to unconventional resources such as tight oil and shale oil and gas [5]. China possesses shale oil resources with significant development potential [6]. The widespread distribution of shale oil reservoirs provides vast storage space for CO2 geological sequestration [7]. However, shale oil resources vary significantly across different regions and types, posing substantial challenges [8]. Under reservoir conditions, CO2 exhibits unique properties such as high density, low viscosity, and high diffusivity [9], enabling it to penetrate shale micro- and nanopores and mobilize the oil stored within. Therefore, CO2-enhanced shale oil recovery has the potential to improve shale oil recovery rates while achieving CO2 geological sequestration, addressing the critical needs of Chinese “dual carbon” goals.
The tight pores in shale oil reservoirs not only lead to rapid reservoir depletion but also cause fluid flow behavior to deviate from conventional reservoirs, transitioning from Darcy to non-Darcy flow [10]. In shale reservoirs with pore throat radii ranging from 1 to 200 nm, the continuum theory of fluid flow fails, and gas molecules move along random paths while still maintaining the general flow direction controlled by the pressure gradient. Molecules collide with pore walls and tend to slide along them [11]. Consequently, the measured gas permeability was observed by Klinkenberg in 1941 to be higher than the values predicted by Darcy’s law [12]. To correct for non-Darcy flow in different flow regimes within nanopore spaces and to fully account for slip flow and Knudsen diffusion in gas non-Darcy flow, previous studies have proposed two types of gas apparent permeability models: the dusty gas model and the Knudsen number-based model. The dusty gas model incorporates Knudsen diffusion and slip flow effects to account for the interaction between gas molecules and solid walls, integrating the three primary transport mechanisms of viscous flow, Knudsen diffusion, and molecular diffusion. This enables the description of gas flow behavior in porous media, including gas non-Darcy effects [13,14,15]. The Knudsen number-based model primarily focuses on gas flow at micro- and nano-scales, especially when flow is influenced by the Knudsen effect (i.e., when the mean free path of gas molecules is comparable to the pore size). This is achieved by adjusting the slip factor or apparent permeability based on a function of the Knudsen number [16,17,18,19,20], effectively multiplying the permeability by a gas non-Darcy coefficient greater than 1.
Horizontal wells with multi-stage hydraulic fracturing represent the primary application scenario for CO2-enhanced oil recovery (EOR) in shale oil reservoirs [21,22,23,24]. CO2 injection in shale oil reservoirs can be implemented in two main modes: continuous injection and huff-and-puff. Generally, huff-and-puff is considered more suitable for low-permeability reservoirs [25]. The fracture network created by hydraulic fracturing in horizontal wells effectively reduces flow resistance and creates pathways for shale oil and CO2 flow [26]. In 2023, Chinese exploration and development of shale oil advanced steadily [27,28,29]. However, Chinese lacustrine shale oil deposits are characterized by thick sedimentary layers, rapid changes in lithology and oil saturation both vertically and laterally. Shale oil reservoirs exhibit well-developed bedding fractures and poor vertical penetration capability of hydraulic fracturing, resulting in limited reservoir modification in the vertical direction after horizontal well fracturing [30]. This makes it difficult to form complex fracture networks, and vertical flowing is limited. Consequently, vertical CO2 migration, oil displacement, and sequestration effects are limited when based on horizontal well fracturing [31]. In summary, horizontal well hydraulic fracturing technology cannot meet all the development needs of CO2 EOR in shale oil reservoirs [32,33,34], and there is a need to explore a new technology that can enable long-term, efficient CO2 injection for shale oil reservoir development.
This study proposes a method for CO2 flooding in fractured shale oil reservoirs based on radial borehole fracturing. Focusing on the three-dimensional fracture network of radial boreholes, a multi-component compositional model was employed to predict the productivity and carbon sequestration capacity. The impacts of non-Darcy flow, diffusion, and adsorption on reservoir fluid flow, oil production, and carbon sequestration were analyzed. Additionally, the adaptability of continuous CO2 injection and CO2 huff-n-puff production modes was compared. The research findings are expected to provide a novel method and theoretical foundation for the efficient development of shale oil and carbon sequestration.

2. Radial Borehole Fracturing in Shale Oil Reservoirs

Increasing stimulated reservoir volume is crucial for shale oil reservoirs. The United States has achieved a revolution in shale oil and gas by employing three-dimensional development technologies, transitioning from an energy importer to an exporter [35]. In 2014, Li et al. first proposed the method of radial borehole fracturing. This technique involves drilling one or more radial horizontal holes along the vertical wellbore using hydraulic jetting, followed by fracturing [36,37]. Radial boreholes guide hydraulic fracturing, allowing for flexible deployment of radial boreholes and hydraulic fractures underground to form multi-layer three-dimensional fractures in shale oil reservoirs [38]. However, due to the tight nature of shale oil reservoirs, depletion development after fracturing fails to maintain stable production. Nevertheless, due to the inherently tight and low-permeability nature of shale formations, depletion development after fracturing often fails to sustain long-term production. Furthermore, in deep shale formations, elevated horizontal stresses pose additional challenges for borehole stability. Stress anisotropy may induce localized casing deformation, shear failure, or collapse-especially near radial outlets where stress concentrations are high [39]. The coupling between in situ stress fields and natural fracture systems must therefore be carefully considered in both fracture design and operational planning to ensure mechanical integrity and field applicability.
In addition to these mechanical challenges, induced geohazards such as microseismicity, fault reactivation, and uncontrolled fracture propagation are also more likely to occur in deep unconventional reservoirs [40]. These phenomena are linked to higher energy release potential and complex stress redistribution patterns, especially when CO2 injection disturbs pre-existing fracture networks [41]. Field-scale monitoring has documented microseismic events associated with both CO2 sequestration and hydraulic stimulation, raising concerns over caprock disturbance or fluid migration through compromised sealing zones [42]. Therefore, it is essential to incorporate these depth-related dynamic risks into reservoir modeling and injection planning to ensure long-term storage safety and operational reliability.
Therefore, CO2 flooding based on multi-layer radial borehole fracturing is proposed to economically and efficiently achieve oil displacement and carbon sequestration [43]. One continuous flooding implementation is shown in Figure 1. After hydraulic fracturing forms fractures in a vertical well with radial boreholes deployed at two depths in a shale oil reservoir, CO2 is injected into the upper radial boreholes through the annulus, driving shale oil to flow into the lower radial boreholes, from which oil is produced to the surface through the tubing. Another implementation involves CO2 huff-and-puff, where the upper and lower radial boreholes serve simultaneously as injection and production wells, alternating between production, injection, and soaking operations. In horizontal well fracturing, it is generally believed that CO2 continuous flooding and CO2 huff-and-puff methods should be selectively applied according to reservoir conditions in well-pattern development, with CO2 huff-and-puff being preferred for low-permeability reservoirs [44]. Therefore, it is necessary to analyze the reservoir adaptability of this method in the context of radial borehole fracturing. The novelty of this method lies in the vertical movement of CO2, which enables long-term carbon sequestration in a single vertical well.
In previous studies, reservoir dynamics and production predictions of CO2 flooding after radial borehole fracturing are analyzed under different fractures and radial borehole parameters. However, the black-oil model used did not account for mechanisms such as adsorption and diffusion. Therefore, further research is needed based on a compositional model.

3. Numerical Model and Solution

3.1. Compositional Model

During CO2-enhanced oil recovery (EOR) in shale oil reservoirs, the adsorption and diffusion of CO2, the miscible/non-miscible displacement between CO2 and oil, and non-Darcy flow are key mechanisms that cannot be ignored. Compared to the black-oil model, the compositional model can more accurately describe the multi-phase, multi-component kinetics during CO2 flooding, especially the diffusion and adsorption of various components. Therefore, this study employs a compositional model to simulate the CO2 flooding process. To optimize simulation convergence, the reservoir temperature is assumed to remain isothermal throughout the process, unaffected by the injection of CO2; the non-stimulated reservoir volume (non-SRV) region forms an effectively closed reservoir boundary due to extremely low permeability; the diffusion and solubility of CO2 in water are neglected, as are capillary pressure, relative permeability hysteresis, threshold pressure effects, and the influence of asphaltene and hydrate formation. For the oil and gas phases, the mass balance equation considering diffusion and adsorption can be expressed as [46]
t [ ϕ α = o , g ρ α S α X α i + δ s 1 ϕ ρ s i ] + α = l , v ( ρ α X α i v α + J α i ) = α = o , g ρ α X α i q α / V
where the superscript i represents the component; the subscript α represents the phase (liquid or vapor); ρ represents density (kg/m3); S represents saturation (dimensionless); X represents mass fraction (dimensionless); V is the Darcy velocity (m/s); δs is a discrimination coefficient, equal to 1 for shale matrix elements and 0 for non-matrix elements; J is the diffusion intensity (kg/s), expressed by Fick’s diffusion law [47] as
J α i = φ S α τ D α i ( ρ α X α i )
where φ is the matrix porosity (dimensionless); D represents the diffusion coefficient (m2/s). The diffusion coefficient in this study is calculated using the empirical formula proposed by Wilke and Chang [48]:
D i k = 7.4 × 10 8 M i k 1 / 2 T μ k V b i 0.6
where the subscript k represents the solvent component into which component i diffuses; M represents the solvent molecular weight; T represents temperature (K); μ represents the solute viscosity (cp); Vb represents the molar volume of the solute at boiling point.
The term involving ρs represents the adsorbed component mass (kg), where ρs is the adsorption density at pressure P (kg/m3). When multiple components are present, this can be expressed using the extended Langmuir equation [49]:
ρ s i = ρ s L i y i P P L i 1 + j = 1 n y i P P L j
where ρSL represents the maximum adsorption capacity (kg/m3), PL represents the minimum pressure required to reach maximum adsorption (Pa), and y represents the molar fraction of each component in the gas phase. It should be noted that using the extended Langmuir equation to describe adsorption phenomena in shale oil is a relatively simplified approach. Currently, the theory of shale oil adsorption is still incomplete, lacking appropriate models to accurately describe the competitive adsorption interactions between CO2 and other components within shale oil. Future research will further investigate component-based competitive adsorption theories for shale oil.
For the liquid phase non-Darcy law, this study employs the exponential non-Darcy equation proposed by Wang and Sheng [50], based on coreflood waterflooding experimental data [51]:
v α = λ α ρ α ( 1 1 + a α e b α | P α | ) Φ
where Φ represents the potential energy, and the dimensionless non-Darcy coefficients aα and bα in the non-Darcy term can be expressed as functions of phase mobility λα:
a α = 0.6095 λ α 3 + 2.5821 λ α 2 3.4594 λ α + 1.5836
a α = e 1.176 λ α
b α = 0.3603 λ α 2 0.1049 λ α + 1.0935
In the equation, the value may become negative when the mobility is excessively high. Therefore, when a negative value occurs, the value of aα is calculated using the equation.
For gases in shale reservoirs, gas transport mechanisms vary with the Knudsen number. When the Knudsen number is below 0.001, intermolecular collisions are prevalent, and the continuum flow mechanism dominates. When the Knudsen number is between 0.001 and 0.1, collisions between molecules and walls increase in frequency, leading to slip flow phenomena where slip effects become significant. When the Knudsen number is between 0.1 and 10, the frequency of molecular-wall collisions and intermolecular collisions is comparable, forming a transition flow state. When the Knudsen number exceeds 10, molecular-wall collisions dominate, and gas transport is primarily driven by Knudsen diffusion. This reflects the transition from continuum flow to Knudsen diffusion, highlighting the dynamic changes in gas transport mechanisms. According to the Beskok and Karniadakis (BK) model [52], the non-Darcy law for gases can be expressed as
v α = λ α ρ α 1 + 4 K n 1 + 4 K n 1 + 128 15 π 2 K n tan 1 4 K n 0.4 Φ
where Kn is the Knudsen number, defined as
K n = λ d
where d represents the pore diameter (m); λ represents the mean free path of molecules (m), defined as [53]
λ = μ v P v π R T 2 M
where μv is the dynamic viscosity of the gas phase (Pa·s). R is the universal gas constant, i.e., 8.314 J/(mol·K). M is the molecular molar mass (kg/mol).
The relationship between pore diameter, permeability, and porosity can be expressed as [54]
d = 2 2 τ K φ
where τ is the tortuosity.
Flash vaporization is solved using the Rachford–Rice method to determine gas–liquid equilibrium, and the physical properties of each phase (including the supercritical phase) are determined based on the Peng–Robinson equation of state. The hydraulic fractures and natural fractures in this study are coupled through the embedded discrete fracture method. This study is based on the MATLAB Reservoir Simulation Toolkit (MRST) platform (version: 2024b) provided by SINTEF (Trondheim, Norway), which runs on the basis of MATLAB R2024b (Beijing, China).
The Newton–Raphson method is employed to solve the discretized equation system [55]. Diffusion and adsorption are implemented through secondary modifications of the mass conservation equation, and the non-Darcy law is realized by multiplying the mobility with a coefficient. The detailed modifications to the formation code can be found in reference [56].

3.2. Model Description

The fluid properties and relative permeability characteristics employed in this study are detailed in reference [57]. The reservoir fluid is composed of six components: CO2, N2, C1, C2-C5, C6-C10, and C11+. A three-dimensional reservoir model was constructed with dimensions of 250 m × 250 m × 90 m. The model domain reflects a conceptual structure in which radial fracturing generates a high-permeability stimulated reservoir volume (SRV) surrounded by tight shale matrix that serves as a hydraulic barrier. This configuration supports the use of quasi-closed boundaries and aligns with existing practices in shale CO2 injection modeling [45].
Previous studies have shown that when the domain radius exceeds twice the fracture half-length (e.g., >200 m), boundary effects are negligible under low-permeability conditions [48]. With a fracture half-length of 50 m, the 250 m lateral extent used in this study ensures containment of pressure and saturation changes within the active region.
As shown in Figure 2, the reservoir features two layers of radial boreholes, located at depths of 15 m and 75 m, respectively. In the CO2 continuous injection process, the upper radial boreholes serve as the CO2 injection well, while the lower radial boreholes function as the production well. Physical experiments and numerical simulations have demonstrated that radial boreholes can induce fracture propagation. When radial boreholes are symmetrically distributed along the direction of the maximum horizontal principal stress, the fractures in the horizontal radial boreholes propagate a certain distance before gradually turning in the direction of the maximum horizontal principal stress under its influence [37]. In addition to governing fracture propagation, horizontal in situ stresses exert a direct influence on the mechanical stability of radial boreholes [41]. High differential stress between the maximum and minimum horizontal principal stresses may cause asymmetric deformation of the borehole wall, leading to localized ovalization, shear failure, or even collapse at radial junctions [39]. These risks are particularly pronounced in deep shale formations, where the energy stored in the stress field is higher and the presence of bedding planes or natural fractures further weakens structural integrity.
To address these challenges, the numerical model incorporated stress-related constraints during borehole layout and fracture simulation. The orientation of the radial boreholes was aligned with the σH direction to minimize shear concentration, and boundary conditions were applied to mimic the realistic stress envelope surrounding the wellbore [58]. These settings ensure that both the predicted fracture paths and borehole integrity remain geomechanically feasible throughout injection and production stages.
The fracture geometry is assumed as depicted in Figure 2, based on the previously established understanding of the influence of stress changes on fracture morphology. The solid black lines in the figure indicate the positions of the simulated fractures, while the dashed red lines represent the direction of hydraulic fracture propagation. Vertical fractures were selected as the primary configuration, reflecting the prevailing fracture orientation in shale reservoirs, where the minimum principal stress is horizontal, favoring vertical propagation. While horizontal fractures may form under special stratigraphic or structural conditions, they are uncommon in most shale gas reservoirs [37].
The fracture half-length of 50 m (total length 100 m) was chosen to match practical constraints in radial borehole fracturing, including limited injection energy, restricted well spacing, and reservoir model dimensions (250 m × 250 m). This design is consistent with prior radial fracturing simulations and field data [59], where fracture extensions typically range from 40 to 100 m.
Moreover, the focus of this study is on vertical CO2 migration between injection and production layers, and horizontal fracture growth plays a secondary role in transport efficiency under these conditions. Although fracture geometry can affect fluid behavior, dynamic fracture propagation and longer-length sensitivity were not included in the current model and are identified in Section 5 as important directions for future work. As the fracture length increases, the fracture will continue to extend along the direction of the dashed red lines, i.e., parallel to the maximum horizontal principal stress.
The radial boreholes are arranged around a vertical main well at the center of the reservoir, with each layer comprising four mutually perpendicular radial boreholes, having a phase angle of 0°. CO2 is injected, and the produced fluids are transported through the main well via the radial boreholes. However, since the main well is cased and cemented and does not directly contact the reservoir, it is not considered in the model. Table 1 lists the basic properties of the reservoir. The specific formulas describing the geometry of the radial borehole fractures can be found in reference [38]. To simplify the model, Jia et al. [59,60] adopted a model with constant fracture conductivity in the drainage area. Therefore, the width, height, and conductivity of the hydraulic fractures are assumed to be constant and independent of their positions. The specific parameters are listed in Table 2. The model is validated in our previous study [61,62,63]. In addition, Table 3 presents the initial and boundary conditions applied in the simulation, including pressure, temperature, saturation, and boundary type definitions for each reservoir face.

3.3. Model Validation

The simulation model was validated using production data from a hydraulically fractured well in the Eagle Ford shale condensate reservoir [45], covering the period from January 2010 to July 2012. The validation focused on the sixth stage of a nine-stage horizontal well, with the reservoir model dimensions set at 67 m (length) × 137 m (width) × 27.4 m (height) and discretized into 2210 grid cells. Key reservoir parameters included a top depth of 3048 m, initial pressure of 48.3 MPa, temperature of 175 °C, matrix porosity of 10%, and permeability of 0.64 mD, with initial fluid saturations of 65% gas and 35% water. The hydraulic fractures were characterized by a half-length of 41.1 m, height of 15.2 m, and initial conductivity of 140 mD·ft (~42.7 mD·m). A constant bottomhole pressure of 20.7 MPa was maintained during production. As shown in Figure 3, the simulated production closely matches field data, confirming the model’s reliability for subsequent analysis.

4. Results and Discussion

4.1. Influence of Non-Darcy Flow

When the matrix permeability is 0.1 mD, non-Darcy flow is a key factor influencing low-permeability oil reservoirs. The bottomhole pressure of the injection radial borehole and the production radial borehole are maintained at 49.26 MPa and 47.26 MPa, respectively, as shown in Figure 4.
As shown in Figure 5, the non-Darcy flow characteristics of liquids and gases in the matrix are distinct, manifesting as a retarding effect with a value less than 1 and a slip effect with a value greater than 1, respectively. According to Equation (5), the smaller the pressure gradient, the stronger the liquid non-Darcy coefficient, which is why the liquid non-Darcy coefficient is only 0.53 near the boundary. Conversely, the higher the mobility, the closer the value of a is to 0, indicating a weaker non-Darcy effect. Therefore, around the interface between CO2 and oil, the viscosity of oil decreases, resulting in a liquid non-Darcy coefficient of approximately 0.92. The low non-Darcy coefficient near the injection radial borehole is due to low liquid saturation, which leads to low relative permeability and, consequently, low mobility. However, since this region does not contain oil, it does not affect oil flow. Compared to the liquid non-Darcy effect, the gas non-Darcy effect is influenced only by viscosity under a given permeability condition. Thus, in regions with low viscosity, the gas non-Darcy coefficient is relatively low.
As shown in Figure 6, when the matrix permeability is 0.001 mD, CO2 injection into the reservoir is impeded, and the viscosity of the reservoir fluid remains essentially unchanged. Under the same bottomhole pressure conditions, the matrix non-Darcy coefficient is below 0.39, while the gas non-Darcy coefficient exceeds 1.1. Therefore, the lower the permeability, the more difficult it is for oil to flow.
Although the gas non-Darcy effect increases the apparent permeability of CO2, it is insufficient to offset the decline in carbon sequestration capacity due to low permeability, as shown in Figure 7 and Figure 8.

4.2. Influence of Diffusion

Diffusion is an important fluid driving force in addition to the pressure gradient, and its intensity is influenced by both the concentration gradient of the substance and its inherent diffusivity. As shown in Figure 9. Diffusion has a relatively high intensity only around the interface between oil and CO2, which is in line with the understanding that diffusion is driven by the concentration difference in substances. The area where the diffusion intensity alternates between positive and negative is the oil–CO2 interface, representing that CO2 diffuses from the inside out and C1 diffuses from the outside in. The strong dispersion intensity of CO2 is approximately ten times that of C1, indicating the strong migration ability of CO2 molecules.
Diffusion is relatively weaker compared to the driving force of pressure gradients. As shown in Figure 10, the rate of CO2 sequestration is relatively high at the beginning of production, but it rapidly decreases due to the release of elastic energy (0–1 h). Once the reservoir pressure gradient stabilizes, the oil production rate gradually increases (1 h–2700 days). Subsequently, CO2 begins to gas-breakthrough in the production well, eventually leading to a gradual decline in both the oil production rate and the rate of carbon sequestration (>2700 days). The overall weak influence of diffusion observed in Figure 7 can be attributed to two key factors. First, the matrix permeability is extremely low (0.001 mD), resulting in fracture-dominated flow where pressure-driven advection overwhelms molecular diffusion. Second, the diffusion coefficient used in this simulation is 10−10 m2/s, which, as confirmed in previous studies (Dai et al., 2025 [45]), lies below the threshold (~10−8 m2/s) required for diffusion to exert a significant impact. Parameter sensitivity analysis in that study showed that only when the diffusion coefficient increases by at least two orders of magnitude does it substantially redistribute CO2, delay breakthrough, or enhance oil production.
Therefore, the similarity between the diffusion and non-diffusion scenarios in Figure 7 is consistent with both physical expectations and prior simulation findings under low-permeability, fracture-dominated conditions.
Furthermore, the impact of diffusion on productivity and carbon sequestration capacity is relatively minor, whether diffusion is considered or not. However, when diffusion is taken into account, CO2 is not only driven downward by the pressure gradient but also migrates laterally into the surrounding reservoir. As a result, less CO2 enters the production well compared to when diffusion is neglected, meaning that diffusion is more favorable for carbon sequestration. Future research will further investigate the impact of diffusion on CO2 sequestration over longer timescales following the cessation of gas injection.
To amplify the effect of CO2 displacement, multiply it by a multiplier before the CO2 diffusion coefficient. The cumulative oil production and cumulative CO2 sequestration were plotted against the diffusion coefficient, as shown in Figure 11. Both the cumulative oil production and CO2 sequestration increased slightly with the increase in the diffusion coefficient multiplier (from 1 to 5 times) but then decreased significantly (from 5 to 40 times). As the diffusion coefficient multiplier increased, the mobility of CO2 within the reservoir was enhanced. Due to its higher diffusion capacity, gas breakthrough occurred earlier in the production well, ultimately leading to a gradual decline in both oil production and CO2 sequestration.

4.3. The Impact of Adsorption

Shale oil matrices are characterized by well-developed micro- and nanopores. The adsorption of multiple components is a significant factor influencing both shale oil production and CO2 sequestration. As shown in Figure 12, when the matrix permeability is 0.1 mD, with the upper radial boreholes injecting CO2 at a constant rate of 5 m3/day and the lower radial borehole maintaining a bottomhole pressure of 38.26 MPa, the distribution of CO2 and C1 adsorption reveals that CO2 adsorption is higher in the regions through which CO2 flows, and it displaces some of the adsorbed C1. Therefore, under the same CO2 injection rate, when the adsorption of CO2 is considered, a portion of the injected CO2 is adsorbed by the matrix, reducing the amount of CO2 flowing towards the production well. This results in a greater amount of CO2 being sequestered when adsorption is taken into account. The adsorption of CO2 displaces light hydrocarbons, and the decrease in reservoir pressure also causes some of the adsorbed oil to desorb, thereby increasing oil production. For example, as shown in Figure 13, the breakthrough time of CO2, i.e., the time when the CO2 production rate significantly increases, is nearly 500 days earlier when adsorption is not considered compared to when it is considered. On Day 4000, the cumulative oil production and carbon sequestration with adsorption considered are 1.047 and 1.312 times those without considering adsorption, respectively. This indicates that carbon sequestration capacity is more sensitive to adsorption.
To amplify the effect of adsorption, multiply it by a multiplier before the adsorption density. As shown in Figure 14, the cumulative oil production increases continuously with the increase in the adsorption density multiplier, while the carbon sequestration capacity decreases with the increase in the adsorption density multiplier. This is because as the adsorption density multiplier increases, the adsorption capacity of shale oil is enhanced. The adsorption of CO2 can displace more light hydrocarbons, but it becomes more difficult to displace the adsorbed oil. As a result, the sequestration capacity of CO2 in the reservoir is slightly reduced. This trend is consistent with previous findings [64,65,66], which showed that fracture connectivity and injection duration significantly influence both hydrocarbon mobilization and gas trapping behavior. A highly connected fracture network enhances early-stage production by creating dominant flow channels, but it can also accelerate CO2 breakthrough, reducing long-term storage effectiveness. Conversely, limited connectivity or shorter injection duration may suppress CO2 migration but compromise oil displacement efficiency. These trade-offs underscore the need for balanced injection strategies that match the physical structure of the reservoir.
Geological heterogeneity has also been shown to increase CO2 storage uncertainty [67,68,69], particularly when small-scale variations in permeability and capillarity influence CO2 phase behavior. Such heterogeneity can lead to uneven plume distribution, bypassed oil zones, and inconsistent trapping efficiency across the reservoir. Therefore, the spatial variability of formation properties must be carefully integrated into simulation workflows and injection design. Our results reaffirm these observations, indicating that adsorption-enhanced trapping in some regions may limit dynamic sweep efficiency elsewhere, and that tailoring injection parameters to local rock characteristics is essential for optimizing both recovery and sequestration outcomes.

4.4. Comparison of Continuous Flooding and Huff-n-Puff

Both CO2 continuous injection and CO2 huff-n-puff are commonly used in shale oil development and carbon sequestration. However, the productivity and carbon sequestration capacity of these two methods vary under different matrix permeabilities. Therefore, three matrix permeabilities—0.001 mD, 0.01 mD, and 0.1 mD—were set. When radial boreholes serve as injection and production wells, the bottomhole pressures are maintained at 38.26 MPa and 58.26 MPa, respectively. Within one cycle of huff-n-puff, oil production, CO2 injection, and shut-in are carried out sequentially, with durations of 36 days, 7.2 days, and 1.8 days, respectively.
The lower the permeability, the poorer the injection capacity of CO2 and the mobility of shale oil. Figure 15 and Figure 16 show the distribution of CO2 mole fraction under continuous injection and huff-n-puff at different permeabilities, respectively. When the permeability is 0.1 mD, CO2 can flow through the majority of the area between the upper and lower radial boreholes under continuous injection, while under huff-n-puff, CO2 affects a spindle-shaped region symmetrically above and below, with a significantly smaller high-concentration CO2 area compared to continuous injection. The results reflect the disadvantages of huff-n-puff at higher permeabilities, namely, discontinuous production and partial early breakthrough of injected CO2. As shown in Figure 17, both the oil production rate and CO2 production rate of huff-n-puff exhibit a saw-tooth pattern, with cumulative oil production and CO2 sequestration at Day 4000 being only 2.23 and 2.55 times that of continuous injection, respectively.
When the permeability is 0.01 mD, CO2 can sweep the upper half of the reservoir under continuous injection, while under huff-n-puff, CO2 affects the upper and lower 1/4 regions. Therefore, the productivity and carbon sequestration capacity of the two methods are relatively close. As shown in Figure 18, at Day 4000, the cumulative oil production of huff-n-puff is 1.55 times that of continuous injection. Although huff-n-puff has a higher initial carbon sequestration capacity than continuous injection, the significant CO2 breakthrough in the later stage reduces the CO2 sequestration efficiency, resulting in lower cumulative carbon sequestration than continuous injection.
When the permeability is 0.001 mD, CO2 under both continuous and huff-n-puff can only sweep a small area around the hydraulic fractures. Therefore, the advantage of huff-n-puff is demonstrated, as it can mobilize the reservoir in multiple layers simultaneously and leverage the diffusion of CO2. For example, as shown in Figure 19, the cumulative oil production and carbon sequestration of huff-n-puff are 4.72 and 2.32 times that of continuous injection, respectively.

4.5. Influence of Model Boundary and Fracture Length on Simulation Results

To evaluate whether simulation outcomes are sensitive to model scale and fracture length assumptions, we conducted a comparative analysis using extended model boundaries and longer fractures. Specifically, we doubled the lateral domain size from 250 m × 250 m to 500 m × 500 m and simultaneously increased the fracture half-length from 50 m to 100 m. Figure 20 presents the cumulative oil production results under four configurations.
The results indicate that while the absolute oil production and CO2 storage values are affected by boundary and fracture size-particularly under ultra-low permeability conditions—the relative performance trends between injection modes remain consistent. For instance, huff-n-puff consistently outperforms continuous injection in 0.001 mD permeability settings, regardless of fracture size or model boundary. The cumulative production of continuous gas injection conditions and huff-and-puff steam injection conditions increased to about 2.18 and 2.20 times, respectively, after the boundaries and fractures extension.
This suggests that although domain and fracture dimensions influence flow volume and injectivity as evidenced by the observation that scaling both the fracture network and domain boundary dimensions by a factor of two results in a proportional doubling of production volume, they do not alter the fundamental comparative insights derived from the simulation. This supports the robustness of the conclusions and confirms that the simplified base case is sufficient for identifying dominant flow patterns and injection strategy suitability.

5. Conclusions

To address the limitations of vertical reservoir stimulation and the rapid decline in shale oil production in fractured horizontal wells, and in conjunction with the concept of CO2-driven shale oil recovery for carbon sequestration, an innovative method of CO2-enhanced shale oil recovery and carbon sequestration using radial borehole multistage fracturing networks based on single-well foundations is proposed. A compositional model was established to analyze the impacts of different mechanisms on reservoir fluid flow, oil production, and carbon sequestration. The adaptability of continuous CO2 injection and CO2 huff-n-puff production modes was compared. The main conclusions of this study are as follows:
(1)
CO2 injection using radial borehole multistage fracturing requires only a single vertical well to achieve long-term oil recovery and carbon sequestration, providing a reference method for efficient carbon sequestration.
(2)
Non-Darcy flow, diffusion, and adsorption are important factors influencing the CO2 injection process in radial borehole fracturing. The liquid-phase non-Darcy effect not only retards oil flow but also counteracts the acceleration of CO2 flow by the gas-phase non-Darcy effect. Both diffusion and adsorption facilitate the retention of CO2 in the reservoir, with the impact order being non-Darcy > adsorption > diffusion.
(3)
For reservoirs with higher permeability (>0.01 mD), continuous CO2 injection can effectively mobilize and sequester carbon in the majority of the reservoir. For reservoirs with lower permeability (<0.001 mD), huff-n-puff development should be employed to mobilize and sequester carbon in the vicinity of the fractures.
(4)
The simulated CO2 injection pressure and rate are within the safe operating limits of conventional injection systems. While technically feasible, the method may still pose risks such as casing corrosion, wellbore instability, and near-well mechanical failure. Future studies should include wellbore integrity modeling and pilot testing to ensure safe field implementation.
(5)
While the compositional model captures key fluid flow and retention mechanisms, it does not explicitly represent fracture-scale leakage, caprock integrity, or heterogeneity in fracture connectivity. In complex field scenarios, these factors can influence both storage security and injectivity. Future improvements should include geomechanical coupling, sealing capacity evaluation, and site-specific calibration to enhance the model’s reliability under field conditions.

Author Contributions

Methodology, X.L.; Formal analysis, Y.C.; Investigation, X.W.; Resources, Q.W.; Data curation, J.D.; Writing—original draft, D.Z.; Writing—review and editing, H.D.; Supervision, W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study has been supported by the CNPC Major Science and Technology Project: Key Technologies for Site Selection Evaluation and Monitoring of CCUS/CCS Storage Geology (2021ZZ01-05), Company Project: Geological and Reservoir Engineering Research and Field Testing of CO2 Flooding in the Junggar Basin (2024XJZX0801) and Science Foundation of China University of Petroleum, Beijing (No. 2462023BJRC025).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Authors Dongyan Zhou, Xiaohui Wang, Wen Zhang, Xiaotian Li, Yang Cao and Qun Wang were employed by UBGP Inc., China National Petroleum Corporation (CNPC). Author Haihai Dong was employed by Xinjiang Oilfield Company, PetroChina. The remaining author declares 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. Bagci, A.S.; Jonnalagadda, A.; Hernandez-Diaz, D. CO2 Injection Well Design Considerations: Completion Design to Injectivity Analysis. In Proceedings of the SPE Annual Technical Conference and Exhibition, New Orleans, LA, USA, 23–25 September 2024; p. D021S016R005. [Google Scholar]
  2. Terrén-Serrano, G.; Martínez-Ramón, M. Kernel learning for intra-hour solar forecasting with infrared sky images and cloud dynamic feature extraction. Renew. Sustain. Energy Rev. 2023, 175, 113125. [Google Scholar] [CrossRef]
  3. Li, Q.; Li, Q.; Wang, F.; Xu, N.; Wang, Y.; Bai, B. Settling behavior and mechanism analysis of kaolinite as a fracture proppant of hydrocarbon reservoirs in CO2 fracturing fluid. Colloids Surf. A Physicochem. Eng. Asp. 2025, 724, 137463. [Google Scholar] [CrossRef]
  4. Li, Q.; Zhao, D.; Yin, J.; Zhou, X.; Li, Y.; Chi, P.; Han, Y.; Ansari, U.; Cheng, Y. Sediment instability caused by gas production from hydrate-bearing sediment in Northern South China Sea by horizontal wellbore: Evolution and mechanism. Nat. Resour. Res. 2023, 32, 1595–1620. [Google Scholar] [CrossRef]
  5. Guo, X.; Ma, X.; Li, M.; Qian, M.; Hu, Z. Mechanisms for lacustrine shale oil enrichment in Chinese sedimentary basins. Oil Gas. Geol. 2023, 44, 1333–1349. [Google Scholar]
  6. Zou, C.; Pan, S.; Jing, Z.; Gao, J.; Yang, Z.; Wu, S.; Zhao, Q. Shale oil and gas revolution and its impact. Acta Pet. Sin. 2020, 41, 1–12. [Google Scholar]
  7. Yan, L.; Chen, F.; Wang, Z.; Yan, Y.; Cao, J.; Wang, K. Challenges and Technical Countermeasures for Effective Development of Shale Oil in China. Pet. Drill. Tech. 2020, 48, 63–69. [Google Scholar]
  8. Sun, L.; Liu, H.; Zhu, R.; Cui, B.; Lei, Z.; Meng, S.; Tang, J. Ten noteworthy issues on shale oil revolution in China. Acta Pet. Sin. 2023, 44, 2007–2019. [Google Scholar]
  9. Li, Y.; Zhao, Q.; Qi, L.; Xue, Z.; Cao, X.; Liu, Z. Evaluation technology and practice of continental shale oil development in China. Pet. Explor. Dev. 2022, 49, 1098–1109. [Google Scholar] [CrossRef]
  10. Swift, G.; Kiel, O. The prediction of gas-well performance including the effect of non-Darcy flow. J. Pet. Technol. 1962, 14, 791–798. [Google Scholar] [CrossRef]
  11. Sherman, F.S. The transition from continuum to molecular flow. Annu. Rev. Fluid. Mech. 1969, 1, 317–340. [Google Scholar] [CrossRef]
  12. Klinkenberg, L.J. The permeability of porous media to liquids and gases: American Petroleum Institute. Drill. Prod. Pract. 1941, 200–213. [Google Scholar]
  13. Javadpour, F. Nanopores and apparent permeability of gas flow in mudrocks (shales and siltstone). J. Can. Pet. Technol. 2009, 48, 16–21. [Google Scholar] [CrossRef]
  14. Javadpour, F.; Fisher, D.; Unsworth, M. Nanoscale gas flow in shale gas sediments. J. Can. Pet. Technol. 2007, 46, PETSOC-07-10-06. [Google Scholar] [CrossRef]
  15. Darabi, H.; Ettehad, A.; Javadpour, F.; Sepehrnoori, K. Gas flow in ultra-tight shale strata. J. Fluid. Mech. 2012, 710, 641–658. [Google Scholar] [CrossRef]
  16. Florence, F.A.; Rushing, J.A.; Newsham, K.E.; Blasingame, T.A. Improved permeability prediction relations for low-permeability sands. In Proceedings of the SPE Rocky Mountain Petroleum Technology Conference/Low-Permeability Reservoirs Symposium, Denver, CO, USA, 16–18 April 2007; p. SPE-107954. [Google Scholar]
  17. Fathi, E.; Tinni, A.; Akkutlu, I.Y. Shale gas correction to Klinkenberg slip theory. In Proceedings of the SPE Unconventional Resources Conference/Gas Technology Symposium, Pittsburgh, PA, USA, 5–7 June 2012; p. SPE-154977. [Google Scholar]
  18. Singh, H.; Javadpour, F.; Ettehadtavakkol, A.; Darabi, H. Nonempirical apparent permeability of shale. SPE Reserv. Eval. Eng. 2014, 17, 414–424. [Google Scholar] [CrossRef]
  19. Shi, K.; Zhang, S.; Chen, J.; Pang, X.; Hui, S.; Chen, D.; Wang, L.; Jin, Y.; Li, C.; Wang, Y.; et al. Influence of pore heterogeneity of shale with different lithofacies on CO2 storage: Experiments and molecular dynamics simulation. Eng. Geol. 2025, 353, 108107. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Hu, J.; Zhang, Q. Simulation study of CO2 Huff-n-Puff in tight oil reservoirs considering molecular diffusion and adsorption. Energies 2019, 12, 2136. [Google Scholar] [CrossRef]
  21. Zhao, M.; Yang, W.; Wang, Y.; Lu, J.; Xu, L.; Li, L.; Li, X.; Yao, L. Distribution and genetic mechanisms of connected pore systems in continental shale reservoirs: A case study of Xujiahe Formation of Upper Triassic, Western Sichuan Depression. Pet. Geol. Exp. 2022, 44, 170–179. [Google Scholar]
  22. He, X.; Li, J.; Wang, J.; Cheng, L.; Chen, X.; Lv, Z. Theoretical understanding, key technologies, and field practice of high-efficiency reservoir stimulation for shale oil in Jimsar Sag. Xinjiang Oil Gas. 2021, 17, 28–35. [Google Scholar]
  23. Zou, C.; Zhu, R.; Bai, B.; Yang, Z.; Hou, L.; Ziia, M.; Fu, J.; Shao, Y.; Liu, K.; Cao, H.; et al. Significance, geologic characteristics, resource potential and future challenges of tight oil and shale oil. Bull. Mineral. Petrol. Geochem. 2015, 34, 3–17. [Google Scholar]
  24. Boak, J.; Kleinberg, R. 4—Shale Gas, Tight Oil, Shale Oil and Hydraulic Fracturing. In Future Energy, 3rd ed.; Trevor, M., Ed.; Elsevier: Letcher, SD, USA, 2020; pp. 67–95. [Google Scholar]
  25. Ma, X.; Li, X.; Liang, F.; Wan, Y.; Shi, Q.; Wang, Y.; Zhang, X.; Che, M.; Guo, W.; Guo, W. Dominating factors on well productivity and development strategies optimization in Weiyuan shale gas play, Sichuan Basin, SW China. Pet. Explor. Dev. 2020, 47, 594–602. [Google Scholar] [CrossRef]
  26. Christensen, J.; Stenby, E.H.; Skauge, A. Review of WAG field experience. SPE Reserv. Eval. Eng. 2001, 4, 97–106. [Google Scholar] [CrossRef]
  27. Sun, L.; Cui, B.; Zhu, R.; Wang, R.; Feng, Z.; Li, B.; Zhang, J.; Gao, B.; Wang, Q.; Zeng, H.; et al. Shale oil enrichment evaluation and production law in Gulong Sag, Songliao Basin, NE China. Pet. Explor. Dev. 2023, 50, 505–519. [Google Scholar] [CrossRef]
  28. Yang, Z.; Hou, L.; Lin, S.; Luo, X.; Zhang, L.; Wu, S.; Cui, J. Geologic characteristics and exploration potential of tight oil and shale oil in Lucaogou Formation in Jimsar sag. China Pet. Explor. 2018, 23, 76. [Google Scholar]
  29. Nguyen, N.T.; Dang, C.T.; Chen, Z.; Nghiem, L.X. Optimization of hydraulic fracturing design with future EOR considerations in shale oil reservoirs. In Proceedings of the SPE Europec Featured at EAGE Conference and Exhibition, Madrid, Spain, 1–4 June 2015. [Google Scholar]
  30. Sun, H. Exploration practice and cognitions of shale oil in Jiyang depression. China Pet. Explor. 2017, 22, 1. [Google Scholar]
  31. Sun, L.; Liu, H.; He, W.; Li, G.; Zhang, S.; Zhu, R.; Jin, X.; Meng, S.; Jiang, H. An analysis of major scientific problems and research paths of Gulong shale oil in Daqing Oilfield, NE China. Pet. Explor. Dev. 2021, 48, 527–540. [Google Scholar] [CrossRef]
  32. Dai, J.; Wang, T.; Tian, K.; Weng, J.; Li, J.; Li, G. CO2 huff-n-puff combined with radial borehole fracturing to enhance oil recovery and store CO2 in a shale oil reservoir. Geoenergy Sci. Eng. 2023, 228, 212012. [Google Scholar] [CrossRef]
  33. Pang, Y.; Wang, Y.; Wang, R. Production test analysis and productivity prediction of horizontal wells in Gulong shale oil reservoirs, Songliao Basin. Pet. Geol. Oilfield Dev. Daqing 2020, 39, 137–146. [Google Scholar]
  34. Sun, H.; Cai, X.; Hu, D.; Lu, Z.; Zhao, P.; Zheng, A.; Li, J.; Wang, H. Theory, technology and practice of shale gas three-dimensional development: A case study of Fuling shale gas field in Sichuan Basin, SW China. Pet. Explor. Dev. 2023, 50, 651–664. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Sheng, M.; Wang, B.; Li, J.; Tian, S.; Zhang, Z.; Li, G. Numerical Simulation for Controlling Fracture Propagation of an Infill Well with Radial Multilateral Wells. Xinjiang Oil Gas. 2022, 18, 31–37. [Google Scholar]
  36. Wang, T.; Guo, Z.; Li, G.; Ma, Z.; Yong, Y.; Chang, X.; Tian, S. Numerical simulation of three-dimensional fracturing fracture propagation in radial boreholes. Pet. Explor. Dev. 2023, 50, 699–711. [Google Scholar] [CrossRef]
  37. Guo, Z.; Tian, S.; Liu, Q.; Ma, L.; Yong, Y.; Yang, R. Experimental investigation on the breakdown pressure and fracture propagation of radial borehole fracturing. J. Pet. Sci. Eng. 2022, 208, 109169. [Google Scholar] [CrossRef]
  38. Dai, J.; Wang, T.; Tian, K.; Li, J.; Tian, S.; Li, G. Research into a productivity prediction model of radial-borehole fracturing in a shale oil reservoir. Pet. Sci. Bull. 2023, 8, 588–599. [Google Scholar]
  39. Liu, X.; Zhang, J.; Jia, H.; Zhang, T.; Du, X.; Wang, Z. The influence of temperature, seepage and stress on the area and category of wellbore instability. Sci. Rep. 2025, 15, 4711. [Google Scholar] [CrossRef] [PubMed]
  40. Li, S.; Zhou, P.; Lan, B. Study of wellbore instability in shale formation considering the effect of hydration on strength weakening. Front. Earth Sci. 2024, 12, 1403902. [Google Scholar] [CrossRef]
  41. Li, S.; Liang, K.; Wang, C.; Jiao, Y.; Liu, H.; Wang, C. Study on the mechanism of anisotropic wellbore instability in continental shale in Songliao Basin. J. Pet. Explor. Prod. Technol. 2022, 12, 2551–2563. [Google Scholar] [CrossRef]
  42. Verliac, M.; Le Calvez, J. Microseismic monitoring for reliable CO2 injection and storage—Geophysical modeling challenges and opportunities. Lead. Edge 2021, 40, 418–423. [Google Scholar] [CrossRef]
  43. Zheng, F.; Jha, B.; Jafarpour, B. Mitigating caprock failure and leakage risks through controlled CO2 injection and coupled flow-geomechanics-fracturing simulation. Int. J. Greenh. Gas. Control 2025, 144, 104387. [Google Scholar] [CrossRef]
  44. Jia, B.; Tsau, J.-S.; Barati, R. A review of the current progress of CO2 injection EOR and carbon storage in shale oil reservoirs. Energy Fuel 2019, 236, 404–427. [Google Scholar] [CrossRef]
  45. Dai, J.; Wang, T.; Zhang, Y.; Zhang, Y.; Zhang, Z.; Zeng, C.; Tian, K.; Li, J.; Tian, S.; Li, G. Enhancing shale oil recovery with water-alternating-CO2 injection through radial borehole fracturing. Pet. Sci. 2025; in press. [Google Scholar]
  46. Zhao, X.; Chen, Z.; Wang, B.; Liao, X.; Li, D.; Zhou, B. A Multi-medium and Multi-mechanism model for CO2 injection and storage in fractured shale gas reservoirs. Fuel 2023, 345, 128167. [Google Scholar] [CrossRef]
  47. Jaynes, D.B.; Rogowski, A.S. Applicability of Fick’s law to gas diffusion. Soil. Sci. Soc. Am. J. 1983, 47, 425–430. [Google Scholar] [CrossRef]
  48. Wilke, C.R.; Chang, P. Correlation of diffusion coefficients in dilute solutions. AIChE J. 1955, 1, 264–270. [Google Scholar] [CrossRef]
  49. Kapoor, A.; Ritter, J.A.; Yang, R.T. An extended Langmuir model for adsorption of gas mixtures on heterogeneous surfaces. Langmuir 1990, 6, 660–664. [Google Scholar] [CrossRef]
  50. Wang, X.; Sheng, J.J. Effect of low-velocity non-Darcy flow on well production performance in shale and tight oil reservoirs. Fuel 2017, 190, 41–46. [Google Scholar] [CrossRef]
  51. Wang, X.; Yang, Z.; Sun, Y.; Liu, X. Experimental and theoretical investigation of nonlinear flow in low permeability reservoir. Procedia Environ. Sci. 2011, 11, 1392–1399. [Google Scholar] [CrossRef]
  52. Beskok, A.; Karniadakis, G.E. A model for flows in channels, pipes, and ducts at micro and nano scales. Microscale Thermophys. Eng. 1999, 3, 43–78. [Google Scholar] [CrossRef]
  53. Civan, F. Effective correlation of apparent gas permeability in tight porous media. Transp. Porous Media 2010, 82, 375–384. [Google Scholar] [CrossRef]
  54. Civan, F. Porous Media Transport Phenomena; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  55. Lie, K.; Krogstad, S.; Ligaarden, I.S.; Natvig, J.R.; Nilsen, H.M.; Skaflestad, B. Open-source MATLAB implementation of consistent discretisations on complex grids. Comput. Geosci. 2012, 16, 297–322. [Google Scholar] [CrossRef]
  56. Yu, W.; Zhang, Y.; Varavei, A.; Sepehrnoori, K.; Zhang, T.; Wu, K.; Miao, J. Compositional simulation of CO2 huff’n’puff in Eagle Ford tight oil reservoirs with CO2 molecular diffusion, nanopore confinement, and complex natural fractures. SPE Reserv. Eval. Eng. 2019, 22, 492–508. [Google Scholar] [CrossRef]
  57. Li, L.; Su, Y.; Sheng, J.; Hao, Y.; Wang, W.; Lv, Y.; Zhao, Q.; Wang, H. Experimental and numerical study on CO2 sweep volume during CO2 huff-n-puff enhanced oil recovery process in shale oil reservoirs. Energy Fuels 2019, 33, 4017–4032. [Google Scholar] [CrossRef]
  58. Zhou, R.; Huang, L.; Rutledge, J. Microseismic event location for monitoring CO2 injection using double-difference tomography. Lead. Edge 2010, 29, 208–214. [Google Scholar] [CrossRef]
  59. Dai, J.C.; Wang, T.Y.; Weng, J.T.; Tian, K.J.; Zhu, L.Y.; Li, G.S. CO2 flooding in shale oil reservoir with radial borehole fracturing for CO2 storage and enhanced oil recovery. Pet. Sci. 2024, 21, 519–534. [Google Scholar] [CrossRef]
  60. Jia, A.; Wei, Y.; Liu, C.; Wang, J.; Qi, Y.; Jia, C.; Li, B. A dynamic prediction model of pressure-control production performance of shale gas fractured horizontal wells and its application. Nat. Gas. Ind. B 2020, 7, 71–81. [Google Scholar] [CrossRef]
  61. Dai, J.; Tian, K.; Xue, Z.; Ren, S.; Wang, T.; Li, J.; Tian, S. CO2-Enhanced Radial Borehole Development of Shale Oil: Production Simulation and Parameter Analysis. Processes 2024, 12, 116. [Google Scholar] [CrossRef]
  62. Guo, T.; Qu, Z.; Gong, D.; Lei, X.; Liu, M. Numerical simulation of directional propagation of hydraulic fracture guided by vertical multi-radial boreholes. J. Nat. Gas. Sci. Eng. 2016, 35, 175–188. [Google Scholar] [CrossRef]
  63. Li, G.; Huang, Z.; Li, J. Study of the key techniques in radial jet drilling. Pet. Drill. Tech. 2017, 45, 1–9. [Google Scholar]
  64. Lie, K.; Møyner, O. Advanced Modelling with the MATLAB Reservoir Simulation Toolbox; Cambridge University Press: Cambridge, UK, 2021. [Google Scholar]
  65. Luo, Z.; Xie, Y.; Zhao, L.; Cheng, L.; Wen, G. A two-phase flow extended finite element technology modeling CO2 in fractured porous media. Greenh. Gases Sci. Technol. 2022, 12, 712–728. [Google Scholar] [CrossRef]
  66. Nguyen, V.; Olayiwola, O.; Guo, B.; Liu, N. Well cement degradation and wellbore integrity in geological CO2 storages: A literature review. arXiv 2023, arXiv:2301.02357. [Google Scholar] [CrossRef]
  67. Vilarrasa, V. Impact of CO2 injection through horizontal and vertical wells on the caprock mechanical stability. Int. J. Rock. Mech. Min. Sci. 2014, 66, 151–159. [Google Scholar] [CrossRef]
  68. Wen, G.; Li, Z.; Long, Q.; Azizzadenesheli, K.; Anandkumar, A.; Benson, S. Real-time high-resolution CO2 geological storage prediction using nested Fourier neural operators. Energy Environ. Sci. 2023, 16, 1732–1741. [Google Scholar] [CrossRef]
  69. Zakirov, T.R.; Khramchenkov, M.G. Effect of pore space heterogeneity on the adsorption dynamics in porous media at various convection-diffusion and reaction conditions: A lattice Boltzmann study. J. Pet. Sci. Eng. 2022, 212, 110300. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of enhanced shale oil recovery using radial borehole fracturing combined with CO2 injection [45].
Figure 1. Schematic illustration of enhanced shale oil recovery using radial borehole fracturing combined with CO2 injection [45].
Processes 13 02873 g001
Figure 2. Distribution of fractures in a dual-layer radial borehole fracturing configuration (a) Three-dimensional schematic view. (b) Top-down plan view.
Figure 2. Distribution of fractures in a dual-layer radial borehole fracturing configuration (a) Three-dimensional schematic view. (b) Top-down plan view.
Processes 13 02873 g002
Figure 3. History matching of oil and gas production rates: comparison between simulation and field data over 1000 days.
Figure 3. History matching of oil and gas production rates: comparison between simulation and field data over 1000 days.
Processes 13 02873 g003
Figure 4. Reservoir characteristic distribution on day 2000 of CO2 continuous injection with radial borehole fracturing.
Figure 4. Reservoir characteristic distribution on day 2000 of CO2 continuous injection with radial borehole fracturing.
Processes 13 02873 g004
Figure 5. Distribution of non-Darcy coefficient on day 2000 of CO2 continuous injection development with matrix permeability of 0.1 mD.
Figure 5. Distribution of non-Darcy coefficient on day 2000 of CO2 continuous injection development with matrix permeability of 0.1 mD.
Processes 13 02873 g005
Figure 6. Distribution of non-Darcy coefficient on day 2000 of CO2 continuous injection development with matrix permeability of 0.001 mD.
Figure 6. Distribution of non-Darcy coefficient on day 2000 of CO2 continuous injection development with matrix permeability of 0.001 mD.
Processes 13 02873 g006
Figure 7. Impact of non-Darcy flow on productivity and carbon sequestration capacity with matrix permeability of 0.1 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 7. Impact of non-Darcy flow on productivity and carbon sequestration capacity with matrix permeability of 0.1 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g007
Figure 8. Impact of non-Darcy flow on productivity and carbon sequestration capacity with matrix permeability of 0.001 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 8. Impact of non-Darcy flow on productivity and carbon sequestration capacity with matrix permeability of 0.001 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g008
Figure 9. Distribution of CO2 and C1 diffusion intensity on day 4000 of CO2 continuous injection development with radial borehole Multistage Fracturing.
Figure 9. Distribution of CO2 and C1 diffusion intensity on day 4000 of CO2 continuous injection development with radial borehole Multistage Fracturing.
Processes 13 02873 g009
Figure 10. Impact of diffusion on productivity and carbon sequestration capacity in CO2 continuous injection development with radial borehole multistage fracturing. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 10. Impact of diffusion on productivity and carbon sequestration capacity in CO2 continuous injection development with radial borehole multistage fracturing. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g010
Figure 11. Impact of diffusion coefficient multiplier on productivity and carbon sequestration capacity after 10,000 days of development.
Figure 11. Impact of diffusion coefficient multiplier on productivity and carbon sequestration capacity after 10,000 days of development.
Processes 13 02873 g011
Figure 12. Distribution of CO2 and C1 adsorption on day 4000 of CO2 continuous injection development with radial borehole multistage fracturing (localized high adsorption values are caused by heterogeneous grid cell volumes).
Figure 12. Distribution of CO2 and C1 adsorption on day 4000 of CO2 continuous injection development with radial borehole multistage fracturing (localized high adsorption values are caused by heterogeneous grid cell volumes).
Processes 13 02873 g012
Figure 13. Impact of adsorption on productivity and carbon sequestration capacity in CO2 continuous injection development with radial borehole multistage fracturing. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 13. Impact of adsorption on productivity and carbon sequestration capacity in CO2 continuous injection development with radial borehole multistage fracturing. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g013
Figure 14. Impact of adsorption density multiplier on productivity and carbon sequestration capacity after 10,000 days of development.
Figure 14. Impact of adsorption density multiplier on productivity and carbon sequestration capacity after 10,000 days of development.
Processes 13 02873 g014
Figure 15. Distribution of CO2 mole fraction on day 4000 of CO2 continuous injection development with radial borehole multistage fracturing.
Figure 15. Distribution of CO2 mole fraction on day 4000 of CO2 continuous injection development with radial borehole multistage fracturing.
Processes 13 02873 g015
Figure 16. Distribution of CO2 mole fraction on day 4000 of CO2 huff-n-puff development with radial borehole multistage fracturing.
Figure 16. Distribution of CO2 mole fraction on day 4000 of CO2 huff-n-puff development with radial borehole multistage fracturing.
Processes 13 02873 g016
Figure 17. Comparison of productivity and carbon sequestration capacity between CO2 continuous injection and huff-n-puff development with matrix permeability of 0.1 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 17. Comparison of productivity and carbon sequestration capacity between CO2 continuous injection and huff-n-puff development with matrix permeability of 0.1 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g017
Figure 18. Comparison of productivity and carbon sequestration capacity between CO2 continuous injection and huff-n-puff development with matrix permeability of 0.01 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 18. Comparison of productivity and carbon sequestration capacity between CO2 continuous injection and huff-n-puff development with matrix permeability of 0.01 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g018
Figure 19. Comparison of productivity and carbon sequestration capacity between CO2 continuous injection and huff-n-puff development with matrix permeability of 0.001 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Figure 19. Comparison of productivity and carbon sequestration capacity between CO2 continuous injection and huff-n-puff development with matrix permeability of 0.001 mD. (a) Oil production. (b) Gas breakthrough and sequestration.
Processes 13 02873 g019
Figure 20. Cumulative oil production under different injection strategies with and without extended domain and fracture length.
Figure 20. Cumulative oil production under different injection strategies with and without extended domain and fracture length.
Processes 13 02873 g020
Table 1. Basic reservoir characteristics.
Table 1. Basic reservoir characteristics.
ParameterValueUnit
Reservoir dimensions (length × width × height)250 × 250 × 90m
Grid discretization (length × width × height)21 × 21 × 31Unit
Reservoir permeability10−3 to 0.1mD
Porosity10%
Tortuosity2
Initial reservoir pressure48.26MPa
Reservoir temperature450K
Table 2. Basic characteristics of radial borehole fracturing.
Table 2. Basic characteristics of radial borehole fracturing.
ParameterValueUnit
Number of main wells1Unit
Number of radial boreholes4Unit
Angle between radial borehole and maximum principal stress45°
Diameter of radial borehole50mm
Length of radial borehole15
(Li et al., 2017 [63])
m
Induced length25m
1°/m
Total length of a single fracture50m
Fracture height9m
Fracture conductivity426.72md-m
Table 3. Initial and boundary conditions used in the simulation.
Table 3. Initial and boundary conditions used in the simulation.
ParameterValueDescription
Initial Pressure32 MPaReservoir initial pressure
Initial Temperature78 °CReservoir initial temperature
Initial Oil Saturation0.65Oil saturation at initial state
Initial Gas Saturation0.05Gas saturation at initial state
Lateral BoundaryConstant-pressureAllows fluid communication
Top BoundaryNo-flowRepresents caprock sealing
Bottom BoundaryNo-flowRepresents base sealing
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

Zhou, D.; Dong, H.; Wang, X.; Zhang, W.; Li, X.; Cao, Y.; Wang, Q.; Dai, J. Numerical Simulation of CO2 Injection and Production in Shale Oil Reservoirs with Radial Borehole Fracturing. Processes 2025, 13, 2873. https://doi.org/10.3390/pr13092873

AMA Style

Zhou D, Dong H, Wang X, Zhang W, Li X, Cao Y, Wang Q, Dai J. Numerical Simulation of CO2 Injection and Production in Shale Oil Reservoirs with Radial Borehole Fracturing. Processes. 2025; 13(9):2873. https://doi.org/10.3390/pr13092873

Chicago/Turabian Style

Zhou, Dongyan, Haihai Dong, Xiaohui Wang, Wen Zhang, Xiaotian Li, Yang Cao, Qun Wang, and Jiacheng Dai. 2025. "Numerical Simulation of CO2 Injection and Production in Shale Oil Reservoirs with Radial Borehole Fracturing" Processes 13, no. 9: 2873. https://doi.org/10.3390/pr13092873

APA Style

Zhou, D., Dong, H., Wang, X., Zhang, W., Li, X., Cao, Y., Wang, Q., & Dai, J. (2025). Numerical Simulation of CO2 Injection and Production in Shale Oil Reservoirs with Radial Borehole Fracturing. Processes, 13(9), 2873. https://doi.org/10.3390/pr13092873

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