1. Introduction
The drilling technology of multilateral wells was first applied by the former Soviet Union in the 1930s, when a turbine drilling assembly was used to drill a multilateral well with 10 laterals. After production, the oil output of this well was approximately 20 times that of a conventional well, laying the initial foundation for the practical application of multilateral well technology [
1]. Entering the mid-to-late 1990s, the petroleum industry demanded higher development efficiency, and multilateral well technology further evolved on the basis of conventional horizontal wells, leading to the formation of multi-branch horizontal well technology [
2]. A multilateral well refers to a wellbore structure in which two or more laterals are drilled from a single main borehole. When these laterals are oriented horizontally, the well is referred to as a horizontal multilateral well [
3]. This technology significantly increases reservoir contact area and improves hydrocarbon flow paths by drilling multiple branches from the main borehole, thereby enhancing recovery efficiency, reducing development costs, and minimizing land use and environmental impacts. As a result, it has been recognized as one of the core technologies for oil and gas field development [
4].
With the deeper application of multilateral well technology, higher requirements have been raised for drilling, hole cleaning, and stability under complex downhole conditions. Consequently, the establishment of surface wellbore simulation systems has become a research focus. As early as 1974, the United States developed a testing apparatus capable of simulating bottom-hole pressures at depths of up to 9100 m, though it could not simultaneously simulate temperature [
5]. In 1986, the United Kingdom developed a system capable of simultaneously simulating both pressure and temperature conditions at a depth of 5000 m [
6]. In China, the first downhole environment simulation device (ZM-35) was developed in 1989 by the Beijing Research Institute of Petroleum Exploration and Development, capable of simulating key variables such as formation pressure, pore pressure, and wellbore fluid column pressure [
7]. This device has been widely applied in new bit testing, rock-breaking mechanism analysis, and drillability studies. The advancement of wellbore simulation technology not only improved experimental controllability of drilling processes but also provided a solid foundation for boundary condition settings and validation in CFD simulations.
Cuttings transport is essentially a complex solid–liquid two-phase flow process, and in-depth study of two-phase flow theory is therefore fundamental to understanding its mechanisms [
8]. Two main approaches are commonly used to model two-phase flows: the Eulerian–Lagrangian method and the Eulerian–Eulerian method. The former solves the liquid phase using the time-averaged Navier–Stokes equations and determines the dispersed phase by tracking individual particle trajectories. The latter treats both the liquid and solid phases as interpenetrating continua, with the total volume fractions of the two phases summing to unity [
9]. During drilling, cuttings transport in the wellbore is essentially a coupled solid–liquid two-phase flow process.
The processes of cuttings deposition, suspension, and transport are strongly influenced by flow field structure, particle properties, and wellbore geometry. According to Heshamudin et al. [
10], the factors affecting cuttings transport can be categorized into three groups: (1) input parameters, such as well inclination, drilling fluid velocity, drillstring rotation speed, borehole size, and fluid rheology; (2) internal states, such as particle size, bed height, annular concentration, and eccentricity; and (3) output indicators, such as annular pressure drop and return fluid density. Existing studies classify two-phase flow in wellbores into four basic flow regimes: fixed cuttings bed flow, moving bed flow, dune flow, and dispersed cuttings flow [
11]. Other researchers have further refined this classification into seven flow patterns based on suspension height and bed stability [
12], as shown in
Figure 1. These classification standards provide a reference for CFD modeling and lay a foundation for studying flow regime transition boundaries.
At present, research on cuttings transport is mainly based on empirical models, theoretical models, and computational fluid dynamics (CFD) models. In recent years, many scholars have carried out studies on cuttings transport [
13,
14]. Ghasemikafrudi and Hashemabadi [
15] simulated liquid–solid two-phase flow in vertical wellbores and analyzed the effects of flow velocity, drillstring rotation speed, and eccentricity on pressure drop. Sun et al. [
16] simulated the processes of cuttings transport and settling in wellbores and proposed a correlation formula for cuttings concentration based on dimensional analysis and infinite regression. Wang et al. [
17] developed an unsteady solid–liquid coupling model and dynamically analyzed the influence of cuttings transport on pressure response. Zhu et al. [
18] proposed a three-layer cuttings model and investigated the effect of drillstring rotation speed on cuttings transport efficiency. Sun et al. [
19] simulated the transport of cuttings in wellbores using an Eulerian solid–liquid model with a standard turbulence closure and analyzed the impacts of drilling fluid velocity and other factors on cuttings migration and annular pressure drop.
In addition, the kinetic theory of granular flow [
20] has been widely applied to model the behavior of cuttings at high concentrations. Xie et al. [
21] and Roy et al. [
22] validated flow behavior under different particle characteristics using the Eulerian–Eulerian two-fluid model. Peng [
23], Hu [
24], and others employed the CFD–DEM coupling method to study particle distribution, wear characteristics, and cuttings transport mechanisms. Zhu et al. [
25] established a two-layer dynamic cuttings bed transport model to simulate cuttings transport under complex conditions in extended-reach wells. Sun Xiaofeng et al. [
26] investigated the mechanism of axial drillstring movement on cuttings transport in horizontal wellbores through CFD simulations Moreover, efficient cuttings transport not only influences the flow stability in branched horizontal wellbores but also plays a crucial role in subsequent drill bit condition evaluation and real-time drilling decision-making. Previous studies have shown that drill bit diagnostics—based on downhole vibration, mechanical response, or video intelligence—can significantly enhance the accuracy of bit wear and condition assessment [
27]. Within this context, the cuttings deposition patterns, settling behavior, and flow disturbances revealed in this study provide more realistic wellbore environmental inputs for drill bit diagnostics, thereby improving the reliability of real-time operational decisions.
Although numerous studies have focused on the hydrodynamic characteristics of cuttings transport, flow regime transitions, and pressure drop prediction, CFD-based investigations specifically targeting cuttings transport in branched horizontal wellbores remain limited. In particular, Fluent-based simulations and analyses for branched wellbore structures are scarce. Compared with straight or single-section wellbores, branched wellbores feature more complex geometries, stronger eccentric annulus effects, and more pronounced particle accumulation and backflow fluctuations, making their cuttings transport characteristics considerably more complicated. Therefore, in this study, a three-dimensional branched wellbore model was constructed within an experimental system, and Fluent 2024 R2 software (version 24.2.0.3796) was employed to perform liquid–solid two-phase flow simulations. The effects of structural parameters such as wellbore diameter and eccentricity on cuttings transport, deposition, and pressure drop were systematically analyzed, providing theoretical insights and simulation-based references for improving hole-cleaning efficiency and optimizing structural design in horizontal branched wellbores.
3. Numerical Simulation Model
3.1. Mathematical Model
In this study, the liquid–solid two-phase flow is described using the Eulerian–Eulerian two-fluid model, combined with the kinetic theory of granular flow [
20], to account for the interactions between drilling fluid and cuttings particles. The governing equations include the continuity equation, momentum conservation equation, and energy conservation equation.
3.1.1. Continuity Equations
The liquid-phase continuity equation reflects the fundamental principle of local mass conservation during the liquid flow process. In other words, the rate of change in liquid mass per unit volume equals the net mass flux through the control surface. It can be expressed as follows:
where
is the liquid volume fraction, %;
is the liquid density, kg/m
3; and
is the liquid velocity vector, m/s.
The solid-phase continuity equation describes the conservation of mass for the solid phase within the flow system. It is analogous to the liquid-phase continuity equation, indicating that the solid mass cannot be created or lost within the control volume. The expression is as follows:
where
is the solid volume fraction (
+
= 1), %;
is the solid density, kg/m
3; and
is the solid velocity vector, m/s.
3.1.2. Momentum Equations
The liquid-phase momentum equation describes that during the flow process, the rate of change in momentum per unit volume of liquid originates from the pressure gradient, and the viscous stress within the liquid, gravitational forces, and the interactions between liquid and solid particles. In other words, it illustrates the motion behavior of the liquid phase under the influence of these forces, and is expressed as follows:
where p is the pressure, N/m
2; τₗ is the liquid stress tensor, Pa; g is the gravitational acceleration vector, m/s
2;
is the interaction force between the liquid and solid phases, N/m
3;
is the molecular viscosity, Pa·s;
is the turbulent viscosity, Pa·s.
The solid-phase momentum equation describes that within the flow system, the rate of change in momentum per unit volume of particles originates from the pressure exerted by the fluid, particle–particle stress, gravitational forces, and the action of the fluid on the particles. In other words, it represents the motion behavior of solid particles under the combined action of various forces in the liquid environment, and can be written as follows:
where
is the solid stress tensor, Pa;
is the interaction force of the liquid phase on the solid phase (
= −
, the interaction forces between the two phases follow Newton’s third law), N/m
3;
is the solid pressure, Pa.
3.1.3. Stress Tensor Equations
The liquid-phase stress tensor is expressed as follows:
where
The solid-phase stress tensor is given as follows:
where
where I is the unit tensor, N/m
2;
is the solid shear stress tensor, N/m
2;
is the granular temperature (a measure of particle kinetic energy), m
2/s
2; e is the coefficient of restitution for particle collisions;
is the total solid shear viscosity, Pa·s.
3.1.4. Solid Shear Stress Model
The solid-phase shear viscosity equation is used to characterize the ability of the solid particle layer to resist shear deformation during flow. Its core concept is the momentum exchange and drag interactions between particles. This equation decomposes the solid-phase viscosity into three physical mechanisms: the collisional contribution, which represents the momentum transfer during short-time elastic collisions of particles; the kinetic contribution, which denotes the shear effect caused by the free motion of particles in dilute regions; the frictional contribution, which describes the shear resistance induced by continuous contacts among particles in dense regions. The expression is given as follows [
29,
30]:
where
where
is the collisional shear viscosity, Pa·s;
is the kinetic shear viscosity, Pa·s;
is the frictional shear viscosity, Pa·s;
is the particle diameter, m;
is the radial distribution function;
is the internal friction angle, rad;
is the local shear rate, 1/s;
is the critical concentration.
3.1.5. Liquid–Solid Interphase Drag Model
In the Euler–Euler model, the momentum exchange between the liquid and solid phases is represented by a drag force source term
. The choice of the drag coefficient
directly affects the prediction of particle acceleration, settling, and pressure drop. In this study, a hybrid drag model combining the Gidaspow model [
31], Ergun model [
32], and Wen-Yu model [
33] is adopted. The governing equations are as follows:
when >0.2 (dense regime):
when ≤0.2 (dilute regime):
where the drag coefficient
is given by [
34]
and the particle Reynolds number is defined as
where
is the interphase drag coefficient, kg/m
3·s;
is the particle drag coefficient [
34];
is the particle Reynolds number.
3.2. Geometric Model
In this study, a three-dimensional horizontal wellbore structure based on a Cartesian coordinate system was established using SOLIDWORKS 2025 software (version 33.0.0.5050) to simulate the transport behavior of drilling fluid carrying cuttings within the wellbore. The inlet, outlet, and flow direction of the drilling fluid are shown in
Figure 3.
The horizontal branched wellbore consists of a section of horizontal main wellbore, a branched wellbore, and a curved section connecting them. The figure presents a top view, with the coordinate system established with the center point of the main wellbore outlet as the origin. The X-axis is defined as the vertical direction, with the negative direction pointing toward gravity (vertically downward) and the positive direction pointing vertically upward, representing the buoyancy direction of the fluid.
The
Y-axis is defined as the direction in the horizontal plane perpendicular to the
Z-axis, which corresponds to the normal expansion direction of the drilling fluid entering from the branched wellbore. The
Z-axis is defined as the extension direction of the main wellbore in the horizontal plane, with the positive direction of the
Z-axis representing the main flow path of the drilling fluid, as shown in
Figure 3.
The drilling fluid enters the system from the right-hand side of the figure, flowing along the Z–Y plane in a curved manner and ultimately exiting the system in the positive direction of the Z-axis at the main wellbore outlet. The geometric relationships between the branched wellbore, curved wellbore, and horizontal main wellbore are defined according to this coordinate system.
The length of the main wellbore is set to 5000 mm, and the length of the branched wellbore is 2300 mm, with an angle of 45° corresponding to the curve to better simulate the flow diversion characteristics in the branched wellbore structure. The model has simplified the structure of the wellbore wall and joints for simplicity and reduced the influence of small-scale boundary movements.
In this paper, a comprehensive consideration of both the calculation resources and the practical situation of the wellbore model is taken. The model parameters used are shown in
Table 1.
As shown in
Figure 4, when performing mesh generation using ANSYS Fluent 2024 R2 Meshing software (version 24.2.0.3796), the structural particularities and detailed features must be fully considered. Due to the complex internal flow structure of the wellbore, with local curvature changes and eccentric narrow gap regions, a hybrid mesh strategy combining structured and unstructured grids is used to improve simulation accuracy and computational stability.
The main wellbore and the branched wellbore are meshed using hexahedral grid combinations, while the curved sections use unstructured grids with local refinement. The outer and inner walls are treated with expansion, and the mesh growth rate is set to 1.2 to ensure overall mesh accuracy.
3.3. Grid Independence Analysis
To minimize the influence of mesh quantity on the simulation results and accurately capture the reverse circulation behavior and cuttings transport characteristics within the branched horizontal wellbore, a wellbore model with an outer diameter of 139.7 mm, an eccentricity of 0%, and a curvature of 2°/30 m was selected for mesh independence verification.
The average pressure at the cross-section connecting the horizontal and curved sections is used as the evaluation parameter.
Figure 5 presents the pressure variation curves obtained under different mesh densities. As shown in
Figure 5, the simulated pressure values gradually stabilize as the number of cells increases. When the mesh count exceeds 500,000, the average cell size is 0.1 m, while the average cell size in the locally refined regions is 0.01 m. Further refinement yields negligible improvement in accuracy. Therefore, a mesh size of approximately 500,000 elements was adopted for the numerical simulations in this study.
3.4. Boundary Conditions
In the numerical simulation conducted using ANSYS Fluent 2024 R2, software (24.2.0.3796) the wall boundary was set as a fixed no-slip condition with the standard wall function model applied. The turbulence model adopted is the RNG k-ε model, which converges faster than the standard k-ε model and is more effective in analyzing turbulent flows with high strain rates and pronounced curvature [
35]. For the discretization of convection terms, the second-order upwind scheme is used due to its higher accuracy and suitability for complex fluid–structure coupling scenarios [
36]. The SIMPLE pressure–velocity coupling algorithm, a pressure-based segregated approach, is employed for the solution process, as it is well suited for low-speed, incompressible flows and has demonstrated strong performance in complex problems such as combustion and multiphase flow [
37].
At the computational domain inlet, the boundary condition is defined as a velocity inlet, with the liquid phase inlet velocity set to 2.8 m/s and the solid phase inlet velocity set to be identical to that of the liquid phase. At the outlet, a pressure outlet boundary condition is applied, with the pressure specified as 1 MPa. To better match field conditions, the Euler–Euler two-fluid model is adopted, in which both the solid and liquid phases are treated as pseudo-continuous media that exist independently yet interpenetrate with each other.
5. Influencing Factors on Cuttings Transport in the Annulus
5.1. Effect of Wellbore Diameter
Under the conditions of an eccentricity of 0%, a branch well curvature of 2°/30 m, a flow rate of 30 L/s, a consistency index
k = 0.5 Pa·s
n, and a flow behavior index n = 0.6, simulations were conducted by varying the wellbore outer diameter to investigate its influence on cuttings transport behavior. In the model, Position 1 is located 2 m from the drilling fluid outlet, Position 2 is 5 m from the outlet at the junction between the curved and straight sections, Position 3 is at the junction between the branch and curved sections, and Position 4 is 6.5 m from the outlet near the lower end of the main wellbore. These positions are presented in
Figure 12.
Figure 12 shows the cuttings volume fraction distribution when the wellbore outer diameter is 139.7 mm. As illustrated in
Figure 12a, the cuttings are relatively uniformly distributed within the annulus, and apart from the region near the lower end of the main wellbore, no significant accumulation occurs elsewhere, indicating that the drilling fluid exhibits strong cuttings-carrying capability under this diameter condition. At Position 2, the average cuttings concentration is 10.01%, which is relatively low, forming a continuous and stable low-concentration zone that reflects effective local transport by the flow field. In contrast, at Position 4, a distinct high-concentration deposition band appears, concentrated in the lower section of the wellbore, with an average concentration as high as 17.94%, indicating severe cuttings accumulation in this confined region.
Figure 13 presents the cuttings volume fraction distribution for a wellbore with an outer diameter of 177.8 mm. As shown in
Figure 13a, the overall cuttings concentration distribution is more complex than that in
Figure 12a, with noticeably higher concentrations appearing in the curved and confluence regions. At Position 2, the average cuttings concentration reaches 10.17%, indicating that as the outer diameter increases, the local fluid’s ability to suspend cuttings weakens, making the particles more prone to retention. At Position 4, although significant deposition remains, the overall concentration decreases to 14.81%, and the extent of high-concentration regions becomes smaller than that in
Figure 12. This demonstrates that a larger wellbore diameter helps to disperse the local flow field, thereby alleviating cuttings accumulation near the lower end of the main wellbore.
Figure 14 shows the cuttings volume fraction cloud plot for a wellbore diameter of 244.5 mm. As shown in
Figure 14a, the overall color is darker, with high-concentration distributions observed in most areas of the wellbore. This indicates that as the wellbore diameter further increases, cuttings are more likely to accumulate within the annular region. At location 2, the average concentration rises to 10.28%, and the red-orange regions expand further, demonstrating a continued decrease in the fluid’s ability to carry the cuttings. However, at location 4, the extent of cuttings deposition is significantly reduced compared to the previous two cases. The average concentration decreases to 12.65%, and the cloud plot shows that the low-concentration region expands, indicating that the fluid’s diffusion effect effectively dispersed the local cuttings accumulation.
Based on the results shown in
Figure 12,
Figure 13 and
Figure 14, the influence of wellbore outer diameter on cuttings transport exhibits a dual characteristic. In most regions of the annulus, except near the lower end of the main wellbore, the cuttings concentration increases with larger diameters, indicating a gradual decline in cuttings-carrying efficiency. Conversely, at the lower end of the main wellbore, an increase in outer diameter helps reduce local cuttings concentration, thereby alleviating deposition.
Overall, the maximum cuttings concentration decreases as the wellbore diameter increases. Smaller diameters are more effective in maintaining clean flow within the branch, curved, and straight sections, while larger diameters are advantageous for mitigating accumulation near the bottom of the main wellbore. This suggests that the design of the wellbore diameter should balance cuttings transport characteristics across different regions.
5.2. Effect of Eccentricity
Under the conditions of a wellbore outer diameter of 139.7 mm, a branch well curvature of 2°/30 m, a flow rate of 30 L/s, a consistency index
k = 0.5 Pa·s
n, and a flow behavior index
n = 0.6, simulations were conducted by varying the eccentricity to investigate its influence on cuttings transport behavior.
Figure 15,
Figure 16 and
Figure 17, respectively, illustrate the cuttings volume fraction distributions under different eccentricities, the cross-sectional volume fraction profiles at various distances from the drilling fluid outlet, and the corresponding velocity field distributions.
As shown in
Figure 15, when the eccentricity is 0%, cuttings are uniformly distributed along the lower section of the main wellbore, with an overall low concentration, and deposition occurs mainly at the bottom near the lower end of the main wellbore. As the eccentricity increases to 20%, 40%, and 60%, cuttings progressively migrate toward the lower side of the annulus, showing a clear accumulation trend. At higher eccentricities, localized regions of elevated concentration appear in the lower annulus, forming distinct zones prone to cuttings deposition. Moreover, with increasing eccentricity, both the local concentration near the bottom of the main wellbore and the overall maximum cuttings volume fraction rise significantly.
Three typical cross-sections at distances of 1 m, 3 m, and 5 m from the main wellbore outlet are extracted to compare the cuttings volume fraction and transport velocity distributions under different eccentricity conditions, as shown in
Figure 16 and
Figure 17. The perspective in the figures is from the main wellbore outlet, looking inward along the centerline direction of the wellbore, with the cross-sectional plane perpendicular to the main wellbore axis.
From the cuttings volume fraction cloud maps in
Figure 16a–c, it can be observed that at 5 m, where the branched wellbore connects to the main wellbore, the drilling fluid carrying cuttings enters the main wellbore from the positive half of the
Y-axis. As a result, the cuttings concentration near the branched wellbore side is significantly higher than on the side far from the branched wellbore. As the drilling fluid transports the cuttings from 5 m to 3 m and 1 m, the distribution of the cuttings concentration gradually returns to a more typical pattern, with the bottom concentration being notably higher than at the upper part.
As the eccentricity increases, the degree of cuttings deposition in the lower annulus progressively intensifies, especially at 1 m, where the effect is most pronounced. At 60% eccentricity, a significant accumulation of cuttings is observed at the bottom of the wellbore, forming a crescent-shaped high-concentration sedimentation zone, with the volume fraction peak being significantly higher than at lower eccentricities. Overall, increasing eccentricity significantly enhances the asymmetry of the annular space, leading to an uneven distribution of cuttings in the cross-section, thereby reducing the efficiency of cuttings transport.
Correspondingly, the cuttings transport velocity cloud plots at different locations shown in
Figure 17 demonstrate that eccentricity also has a significant impact on the velocity distribution. At 5 m, where the branched wellbore connects to the main wellbore, the velocity near the branched wellbore is significantly lower than that far from the branched wellbore. As eccentricity increases, this trend becomes more pronounced. From the velocity cross-sectional cloud plots at 5 m, 3 m, and 1 m from the outlet, it can be observed that the cuttings transport velocity in the lower annular region near the wellbore wall decreases significantly, and the high-velocity region gradually shifts upward in the wellbore, showing a distinct shift in the flow field.
This asymmetric change in velocity distribution exacerbates the unevenness of cuttings deposition, especially by creating a stagnation zone in the lower part where the velocity is extremely low, further promoting cuttings accumulation and deposition in this region.
In summary, the higher the eccentricity is, the more apparent the instability of cuttings transport in the annulus is. Increased eccentricity intensifies the velocity asymmetry, reinforcing the deposition mechanism, making it one of the key factors affecting cuttings deposition. The higher the eccentricity, the more severe the cuttings deposition in the lower wellbore, with higher concentration and lower velocity, leading to reduced cuttings transport efficiency.
5.3. Influence of Branch Wellbore Curvature
Keeping all other parameters constant, the wellbore outer diameter of 139.7 mm, eccentricity of 0%, a flow rate of 30 L/s, a consistency index k = 0.5 Pa·sn, and a flow behavior index n = 0.6. This section investigates the influence of branch wellbore curvature on cuttings transport characteristics. The branch curvature was set to 1.5°/30 m, 2°/30 m, 2.5°/30 m, and 3°/30 m, corresponding to curvature radii of 1145.92 mm, 859.44 mm, 687.55 mm, and 572.96 mm, respectively.
As shown in
Figure 18, the cuttings volume fraction distributions and corresponding average value curves at a cross-section located 6 m from the main wellbore outlet (i.e., near the lower end of the main wellbore) reveal that the cuttings accumulation area in the middle of the cross-section gradually expands with increasing curvature. The cuttings volume fraction also rises, and the overall distribution becomes less uniform. When the curvature is 1.5°/30 m, cuttings mainly concentrate in the lower region of the wellbore, with relatively low volume fractions. As the curvature increases to 3°/30 m, the average cuttings volume fraction rises from approximately 14.8% to 15.9%, indicating that greater curvature significantly enhances particle settling near the bottom of the main wellbore.
This behavior can be attributed to the intensified bending of the fluid flow path caused by higher curvature, which drives the cuttings downward under centrifugal force, leading to a distinct accumulation layer near the bottom. Moreover, larger curvature results in stronger flow disturbances at the junction between the branch and main wellbores, further promoting particle enrichment in the lower region.
To further analyze the variation trend of cuttings transport capacity,
Figure 19 presents the annular velocity distribution contours and corresponding average velocity curves at the junction between the branch and main wellbores under different curvature conditions. The results indicate that as the wellbore curvature increases, the overall flow velocity at the junction slightly rises, with the cross-sectional average velocity increasing from 2.794 m/s to 2.808 m/s. This demonstrates that greater curvature intensifies local flow velocity and strengthens the flow field.
The increase in velocity can be attributed, on one hand, to the enhanced inertial effects and pressure drop within the curved section, and on the other hand, to the improved local cuttings-carrying capacity induced by the strengthened flow. However, since the cuttings’ particle size and density remain constant and the flow rate is unchanged, the elevated local velocity cannot effectively remove deposited particles from low-velocity regions. Consequently, the overall cuttings concentration increases despite the localized flow intensification.
Therefore, it can be concluded that increasing the curvature of the branch wellbore enhances local flow disturbances and velocity but does not effectively improve the overall cuttings-carrying performance. On the contrary, the curved geometry promotes greater particle retention in the lower part of the junction section, thereby intensifying the risk of accumulation near the lower end of the main wellbore.
This finding provides important insight for the optimization of branch wellbore design, suggesting that the curvature of the branch section should be carefully controlled to balance directional performance and flow stability. Maintaining a moderate curvature is recommended to reduce cuttings accumulation and minimize the risk of blockage in the junction and near-bottom regions.
In branch wells, flow disturbances at junctions and curved sections induce local recirculation and vortices, causing transient cuttings retention and release. This leads to deviations between downstream concentration measurements and actual transport efficiency. The mechanism is analogous to the wellbore storage effect in transient pressure analysis, where temporary fluid storage causes signal delay and distortion. Similarly, the junction region acts as a “cuttings storage unit”, introducing phase lag and filtering effects that may result in the apparent misjudgment of transport efficiency.
5.4. Influence of Flow Rate
With other conditions kept constant, including a wellbore diameter of 139.7 mm, eccentricity of 0%, branch wellbore curvature of 2°/30 m, consistency index k = 0.5 Pa·sn, and flow behavior index n = 0.6, this section investigates the effect of different flow rates on cuttings transport in horizontal branch wellbores. Four flow rates were considered: 25 L/s, 30 L/s, 35 L/s, and 40 L/s. The analysis focuses on variations in the cuttings volume fraction distribution along radial positions at the outlet, the concentration distribution at the cross-section of the branch and main wellbore junction, and the evolution of the annular velocity field.
As shown in
Figure 20, within the 0~5 m section of the main wellbore, the cuttings concentration exhibits a slightly decreasing trend with increasing radial distance from the outlet, regardless of the flow rate. Overall, larger flow rates correspond to lower average cuttings concentrations in this region. At 5 m, corresponding to the junction between the curved and straight wellbore sections, a sharp increase in cuttings volume fraction is observed, indicating that this location is a critical area prone to plugging, where cuttings tend to accumulate. In the 5~8 m range, corresponding to the lower end of the main wellbore, the cuttings concentration increases progressively with distance from the outlet. Unlike the upstream section, this region shows the opposite trend: as the flow rate increases, the average cuttings concentration also rises, suggesting that higher flow rates intensify recirculation and promote particle deposition near the wellbore bottom.
From the cuttings concentration contour maps at the junction of the curved and straight wellbore sections under different flow rates shown in
Figure 21, it can be observed that increasing the flow rate significantly alleviates cuttings accumulation in the lower part of the wellbore. The cross-sectional average cuttings volume fractions at flow rates of 25 L/s, 30 L/s, 35 L/s, and 40 L/s are 10.04867%, 10.00945%, 9.993403%, and 9.967008%, respectively. Although the overall reduction in average concentration is relatively small, the area of high-concentration regions decreases and the concentration gradient becomes smoother with higher flow rates, indicating an improved cuttings transport effect. In particular, at 35 L/s and 40 L/s, the cuttings distribution is more uniform and the overall concentration is reduced, demonstrating the enhanced carrying efficiency at larger flow rates.
Further analysis of
Figure 22 shows that increasing the flow rate markedly elevates the annular velocity at the junction, with the cross-sectional average rising from 2.33 m/s to 3.73 m/s. The velocity contours also indicate a pronounced shrinkage of low-velocity zones and a more uniform velocity distribution across the section. While higher velocities enhance momentum transfer and help sweep cuttings out of low-speed deposition regions, thereby reducing the likelihood of settling, high flow rates simultaneously intensify recirculation. As a result, despite improved cleaning of the upstream main wellbore, cuttings accumulation near the lower end of the main wellbore increases, leading to a higher local concentration in that region.
Based on the above analysis, under the present experimental conditions and model settings, increasing the drilling fluid flow rate effectively enhances the flow field strength and the cuttings-carrying capacity in the main wellbore section, suppressing particle accumulation at the junction and improving the overall cuttings transport performance in complex wellbore geometries. This enhancement contributes to higher hole-cleaning efficiency. However, in the lower end region of the main wellbore, the intensified recirculation associated with higher flow rates leads to an unexpected rise in the average cuttings concentration, resulting in more severe particle deposition along the wellbore bottom.
5.5. Influence of Rheological Parameters (k–n Values)
With other conditions kept constant, including a wellbore diameter of 139.7 mm, eccentricity of 0%, branch wellbore curvature of 2°/30 m, and flow rate of 30 L/s, this section further analyzes the influence of non-Newtonian fluid rheological parameters on cuttings transport within the wellbore.
Figure 23 illustrates the effects of different flow behavior indices (
n) on cuttings concentration, pressure, and velocity at 5 m from the drilling fluid outlet, corresponding to the junction between the branch and main wellbores, under the fixed condition of a consistency index (
k) of 0.5 Pa·s
n.
As n increases from 0.4 to 0.8, the average cuttings volume fraction rises from 9.96% to 10.08%, showing a gradual upward trend, which indicates that as the shear-thinning property of the fluid weakens, cuttings in the annulus are more easily suspended and carried. Meanwhile, the wellbore pressure increases from 1,010,894 Pa to 1,042,880 Pa, reflecting a significant rise in pressure drop. The average velocity, however, decreases slightly from 2.801 m/s to 2.794 m/s. These results suggest that a larger flow behavior index enhances cuttings-carrying capacity, but at the expense of higher-pressure losses and a slight reduction in velocity.
Figure 24 presents the influence of different consistency indices (
k) on cuttings transport in the wellbore under the fixed condition of a flow behavior index (
n) of 0.6. As
k increases from 0.2 Pa·s
n to 0.8 Pa·s
n, the average cuttings volume fraction gradually decreases from 10.13% to 9.99%. This indicates that a larger consistency index enhances the overall viscosity of the fluid, leading to a more uniform and stable distribution of cuttings, but at the expense of reduced suspension capacity. Meanwhile, the wellbore pressure rises significantly from 1,011,539 Pa to 1,022,134 Pa, reflecting a marked increase in flow resistance with higher
k. The average velocity decreases slightly from 2.797 m/s to 2.795 m/s, showing only a minor change but with a clear downward trend.
In summary, the rheological parameters of non-Newtonian fluids exert a significant influence on the cuttings transport process. An increase in the flow behavior index (n) enhances the suspension and carrying capacity of cuttings but leads to greater annular pressure drops. Conversely, an increase in the consistency index (k) primarily manifests as elevated flow resistance and suppressed velocity, while the cuttings volume fraction shows a slight decrease.