The immiscible displacement of a wetting fluid by a non-wetting fluid in rough fractures is crucial in many subsurface applications. Hydrodynamic-scale modeling of such drainage flows is challenging due to the complex interaction between the forces at play, the intricate geometry, and the required modeling of moving contact lines. In addition, a remaining critical open question is to what extent not resolving the films of wetting fluid deposited on fracture walls degrades numerical predictions. We address this question by solving the Navier–Stokes equations, employing the volume-of-fluid method to capture fluid–fluid interfaces and considering numerical meshes that result in either resolved films (RF) or unresolved films (UF) in the simulations. The numerical model, implemented in OpenFOAM, is validated in the classical Saffman–Taylor (ST) viscous instability configuration using the original ST experimental measurements; at capillary numbers (Ca) larger than 103, UF simulations overpredict ST finger widths. We then address two synthetic fracture geometries: one with sinusoidally varying apertures and one with stochastic geometric properties typical of geological fractures. Predictions of RF and UF simulations are compared quantitatively for Ca ranging between 105 and 103. Wetting film thicknesses follow a power law of Ca similar to Bretherton's law. RF and UF approaches both predict similar invasion patterns, but the latter underestimates interfacial lengths and macroscopic pressure drops, as compared to RF simulations, while overpredicting invading fluid saturations and breakthrough times. These discrepancies increase with Ca, whereas the disordered nature of the geological fracture tends to limit them. For Ca<105, the discrepancies are negligible.

The immiscible displacement of a wetting fluid by a non-wetting fluid (i.e., drainage) in rough-walled, open geological fractures holds significant relevance for various applications, encompassing subsurface fluid storage (Kim , 2011; Muller , 2019), subsurface contamination remediation (Lee , 2010), and oil and gas exploitation (Bikkina , 2016; Zhang , 2017). In hydrogeological frameworks, the subsurface environment is often conceptualized as a rigid, porous medium. However, many of the above applications involve a fractured or fissured subsurface, such as rock formations (Birkholzer and Tsang, 2000; Xue , 2018). Accurate predictions of two-phase flow phenomena in these applications necessitate resolving hydrodynamic-scale physics, in which interfacial interactions profoundly impact macroscopic two-phase flow dynamics (Meakin and Tartakovsky, 2009). Although significant research has been dedicated to studying the rich phenomenology of two-phase flows in porous media over the past 20 years, the modeling of immiscible displacement in rough fractures is a comparatively emerging field.

Rough fractures consist of the space between two solid walls that result from the solid material having been fractured. In the subsurface, in particular, fracturing occurs due to constraints imposed by tectonic activity or the quick cooling of solid rock. Irrespective of the solid material and the fracturing mode, the resulting fracture surfaces have a universal geometrical property, which is a self-affine scale invariance characterized by a roughness exponent (or Hurst exponent) (Bouchaud, 1997; Schmittbuhl , 1995). This property is valid at scales ranging from micrometers in metals and ceramics to thousands of kilometers in tectonic faults. The resulting geometry is asymptotically planar at large scales but gets all the rougher as one examines smaller scales. In a freshly fractured solid material, if there is no loose material, the two facing walls are perfectly matched, i.e., they have the same geometry. However, any shift in the walls' positions along their mean planes renders the distance between the two facing walls, denoted local fracture aperture, spatially dependent. In geological fractures, chemical weathering over geological times, as well as tectonic constraints and fracturing around contact points, also results in apertures that vary spatially, with fracture walls that have maintained their self-affine geometry.

The complex fracture geometry leads to complex flow, both for single-phase Newtonian (Brown, 1987; Zimmerman and Bodvarsson, 1996; and Méheust and Schmittbuhl, 2001) and non-Newtonian (Lavrov, 2014; Lenci , 2022a, 2022b; and Zhang , 2023) fluids, with in-plane channeling, which impacts the fracture transmissivity. This structural heterogeneity also impacts immiscible two-phase flow (Glass , 2001). For two-phase flows, the displacement patterns have been studied extensively for capillary-dominated flows. Depending on the roughness and correlation structure, the patterns change from invasion percolation structures to more compact patterns, as the in-plane component of the capillary pressure smoothes out small-scale fluid clusters (Glass , 2001; Neuweiler , 2004; and Yang , 2012). The patterns were investigated in detail by Yang (2016), who showed that closing a fracture more leads to less compact patterns due to an increase in the out-of-plane component of the capillary pressure, relative to the in-plane component, in agreement with the previous findings. They also showed that the latter in-plane component suppresses the trapping of fluid clusters below the correlation length of the fractures and changes the characteristics of the size distribution of trapped clusters. Yang (2019) studied the fluid distribution for a wider range of capillary numbers, for which the patterns transitioned from capillary fingers to viscous fingers. The in-plane curvature influenced patterns for cases with low capillary numbers and large viscosity ratios, leading to not only more compact clusters but also a change in trapping behavior; the patterns' fractal dimension varied between 1.55 and 1.85 depending on the capillary number and viscosity ratio between the two fluids. Chen (2017) characterized the transition between capillary fingering and viscous fingering experimentally in detail from various observables, in a transparent replica of a fracture in granite rock; in particular, they observed different dependencies of the tip of the most advanced finger on the capillary number, depending on the displacement regime. The same group later extended these experimental findings by examining the corresponding flow velocity fields (Chen , 2018), using direct numerical simulations in the same geometry as that of Chen (2017), validated by comparisons with their experiments.

Numerical approaches for modeling two-phase flows in rough fractures have developed gradually in the past 20 years. The first models addressed very low capillary (quasi-stationary) displacements with invasion percolation (IP) schemes (Glass , 1998; Detwiler , 2009) inspired by similar approaches on two-dimensional (2D) porous media, with a constant in-plane component of the capillary pressure due to an in-plane curvature radius that was considered a characteristic length scale of the geometry. Yang (2016) also used an IP approach, but with an in-plane curvature that was computed in time and space as a purely geometrical quantity based on the geometry of the interface. Yang (2019) proposed a mesoscopic model treating each fluid phase at the continuum scale with proper boundary conditions at the fluid–fluid interface, including an in-plane curvature defined from the interface geometry. Pore network models (Fatt, 1956) have also been used for simulating fluid flow in permeable media, including rough fractures (Ferer , 2011) or fracture networks (Hughes and Blunt, 2001). Although computationally efficient, these models employ simplified representations of pore geometry and, especially in multiphase flow, simplified physics (Meakin and Tartakovsky, 2009). Recent advancements in computational capabilities have enabled the resolution of the Navier–Stokes (NS) equations, either directly (direct numerical simulations) or using methods that are equivalent to solving the NS equations, such as smoothed particle hydrodynamics (SPH) (Tartakovsky and Meakin, 2005) and lattice Boltzmann methods (LB) (Dou , 2013; Guiltinan , 2021). However, these methods face challenges in relating interaction forces to fluid properties. For example, LB methods require case-specific model calibration to match density ratio and surface tension based on LB parameters, with complex parameter interdependencies (Porter , 2012; Huang , 2011).

The direct numerical simulation (DNS) of the NS equations has superior numerical efficiency and is capable of handling large viscosity and density ratios (Meakin and Tartakovsky, 2009), hence its use in recent studies (Ferrari , 2015; Chen , 2018). Hydrodynamic scale modeling of immiscible flows in rough fractures is a complex and challenging task influenced by a multitude of factors, such as the complex interplay between the viscous, capillary, gravitational, and inertial forces (Detwiler , 2009; Glass , 1998); characteristic wall roughness and wetting properties (Auradou, 2009); intricate geometries and aperture variability (Glass , 2003); molecular scale phenomena such as moving contact lines and thin wetting films (Shi , 2018; Qiu , 2023); and interfacial fragmentation and coalescence (Amundsen , 1999; Santiago , 2016). Among these challenges, a crucial yet largely unexplored one is capturing the thin film deposited on the fracture walls by the displaced wetting fluid. In subsurface applications such as enhanced oil recovery, the wetting films have a significant impact on the macroscopic transport properties (Blunt , 1992; Hui and Blunt, 2000). For instance, films can speed up invading fluid's flow through constrictions, resulting in a lubricating effect that, depending on the viscosity ratio, can yield relative permeabilities greater than 1 (Blunt , 1992). Resolving thin films in hydrodynamics-scale models is computationally challenging, primarily due to the large aspect ratio of the different length scales (Zhao , 2019). For context, fracture dimensions along the mean fracture plane typically span centimeters to kilometers in the subsurface, with aperture thicknesses up to a few centimeters, while in laboratory replicas, the dimensions span up to tens of centimeters and the aperture up to a few millimeters. For comparison, film thicknesses range from a micrometer to ten micrometers. While some numerical studies have reported satisfying comparisons between the experimental measurements and numerical predictions using numerical models that do not resolve wetting films (Chen , 2018), to the best of our knowledge, the consequences of ignoring the film on the predicted flow and spatial distribution of fluid phases for drainage in open rough fractures have not been studied systematically. The main focus of the present study centers around this critical question.

To resolve the fluid–fluid interface, we have chosen to couple a NS solver with the volume of fluid (VOF) method (Hirt and Nichols, 1981). It is one of several interface-capturing methods, which also include the Level-Set (LS) (Sethian, 1996) and phase-field (PF) (Sun and Beckermann, 2007) methods. These Eulerian-based methods are well-suited for flows with complex interface motion and can handle large interface deformations such as breakup or coalescence (Gopala and van Wachem, 2008). The VOF method, in particular, has been effectively applied to model immiscible flow in both fractured (Chen , 2018) and porous media (Ferrari and Lunati, 2013; Ferrari , 2015). Experimental validation from these studies has substantiated the VOF model's capacity to reasonably predict invasion structures and faithfully reproduce phenomena such as viscous meniscus deformation, snap-off events, and coalescence. Valuable insights on the issue of resolving wetting films have emerged from porous media studies, where the issue of wetting films was successfully tackled using the VOF approach. For instance, Horgue (2013) conducted experimental and numerical investigations of liquid jets spreading across in-line arrays of cylinders confined within a Hele–Shaw cell. Additionally, Ferrari (2015) employed the VOF model to replicate primary drainage experiments in a 2D porous medium consisting of a Hele–Shaw cell filled with cylindrical grains. Inspired by prior studies on Taylor flow in microchannels (Gupta , 2009; Horgue , 2012), these two studies showed that to numerically capture the thin wetting film, the liquid film region must be covered by at least two grid cells.

Transferring the film resolution approach from these studies to the case of rough fractures is not straightforward. Complex fracture geometries present significant challenges for numerical grid generation procedures. As the location of the fluid–fluid interface cannot be known a priori, selective region refinements are not feasible. Abrupt changes in local aperture distribution and the presence of zero-aperture regions where fracture walls touch each other further add to the difficulties. Previous studies on bubble motion in tubes with wetting fluid film have revealed that film thickness scales as Ca2/3, where Ca is the capillary number (Hodges , 2004; Bretherton, 1961). As Ca decreases and gets into ranges <105, film thickness nears the nanoscale, requiring ultra-fine grid resolutions. To accurately describe multiphase flow in permeable media, another crucial consideration is the behavior of moving contact lines resulting from fluid–fluid interface interactions with solid surfaces (Meakin and Tartakovsky, 2009). Modeling fluid behavior around a moving contact line is complicated, as the no-slip boundary condition at the solid–fluid interface leads to a force singularity at the contact line (Pasandideh-Fard , 1996). Therefore, it is important to include “slip” when dealing with moving contact lines (Dussan , 1991). The magnitude of this slip is quantified by the so-called slip length, which is usually in the molecular range and is much smaller than any numerically achievable grid sizes (see Renardy , 2001 for details). In practice, numerical calculations implicitly introduce a slip length, approximately Δxwall/2, with Δxwall representing the typical cell dimension near solid walls (Renardy , 2001). Consequently, grid resolution near the walls significantly influences the contact line motion, making complete grid independence for a VOF code challenging in such scenarios (Rosengarten , 2006). Hence, conducting a systematic validation of the VOF model becomes paramount to ensure the accuracy and reliability of investigations of immiscible displacements in rough fractures across varying grid resolutions.

The main goal of this study is to comprehensively assess how resolving thin wetting films impacts model predictions of drainage in open rough fractures compared to not resolving them. Our study is restricted to the range of flow cases where the wetting film thicknesses do not approach the nanoscale and can be well resolved with realistic grid resolutions. We aim to identify scenarios where wetting films significantly affect results, necessitating their inclusion in the model. Our approach involves initial model validation considering a Hele–Shaw geometry based on the classical experiments of Saffman and Taylor (1958). Subsequently, we perform numerical drainage experiments with two distinct rough fracture geometries. We then compare outcomes from two configurations: one resolving the film and the other not. We analyze various hydrodynamic and macroscopic scale properties, including invasion patterns and average pressure drops, while discussing the computational requirements. Our numerical simulations utilize interFoam, a finite-volume, VOF-based code developed by OpenFOAM (OpenFOAM, 2012), which we have self-modified to address the specific numerical challenges.

The paper is structured as follows: In Sec. II, the numerical model is described in detail. The model is then validated on a Hele–Shaw geometry in Sec. III, while the results and discussions of the numerical experiments in rough fractures are presented in Sec. IV. Finally, Sec. V presents the conclusions and future outlooks.

In this section, for the sake of completeness, we briefly outline the numerical model as implemented in OpenFOAM (2012) and employed to perform the simulations presented hereafter. We describe the fundamental governing equations (Sec. II A) as well as an enhancement to the standard volume-of-fluid (VOF) OpenFOAM solver (Sec. II B) that we have implemented, and briefly present the solution procedure (Sec. II C). Readers interested in the details of the model implementation, numerical discretizations, and solution procedure are encouraged to refer to the provided supplementary material and references therein.

In the VOF method, the flow of two immiscible fluids (denoted 1 and 2) is governed by a single set of Navier–Stokes (NS) equations, treating the two fluid phases as a single effective fluid whose properties vary spatially depending on an indicator function γ. The value of γ at the center of a mesh cell corresponds to the phase fraction of fluid 1 in the cell. Consequently, its value varies between 0 (i.e., only fluid 2) and 1 (i.e., only fluid 1), with intermediate values defining the fluid–fluid interface. For two incompressible, immiscible fluids under isothermal conditions, the governing equations of the model can thus be written as
(1)
(2)
(3)
where the first two equations are the NS equations and the last one describes the advection of the indicator function by the flow (and, hence, the displacement of fluid–fluid interfaces); u is the vectorial velocity field shared by the two fluids, p is the pressure field, g is the gravitational acceleration, and fσ is the volumetric density of capillary (surface tension) forces. The density ρ and dynamic viscosity μ of the effective, single, fluid are defined as
(4)
where ρi and μi (i = 1 or 2) are, respectively, the density and viscosity of phase i.
In Eq. (2), the volumetric density of the capillary forces fσ is modeled by the Continuum Surface Force (CSF) model (Brackbill , 1992), which represents the surface tension effects as a continuous volumetric force given by fσ=σκγ. Here σ is the surface tension coefficient and κ the mean interface curvature, evaluated as κ=·n. Away from the walls, the unit interface normal vector is calculated as n=γ/||γ||. At the walls, where the fluid–fluid interfaces meet the solid, wettability effects are taken into account, and the interface normal vector at the walls, nw is thus calculated (Brackbill , 1992) as
(5)
where θ is the static equilibrium contact angle, ns is the unit vector normal to the wall pointing into the solid, and nt is the unit vector tangent to the wall and perpendicular to the triple-contact line, pointing into the non-wetting phase. The nw component is then used to adjust the interface curvature κw at the walls using κw=·nw. For a detailed discussion on wall adhesion and contact angle, see the supplementary material and references therein.
Note that in order to obtain a sharp fluid–fluid interface, the OpenFOAM (2012) based VOF solver interFoam utilizes a modified phase fraction transport equation that features an artificial “compression” term in Eq. (3) (Berberović , 2009) featuring a so-called “compression velocity” uc:
(6)
The compression velocity is defined as
(7)
where the presence of the interface unit normal n ensures that the compression velocity acts only in the direction normal to the interface. To limit the value of uc, the maximum velocity in the flow domain is chosen as the worst case value (Berberović , 2009). The compression coefficient Cγ controls the strength of the interface compression; a value of zero implies no compression, while setting it to one results in a balance between interface compression and unwanted parasitic velocities (Deshpande , 2012; Hoang , 2013). We thus use Cγ=1.
Since the phase indicator γ can abruptly change over a small region of space, the curvature calculation is susceptible to errors, leading to the appearance of so-called parasitic velocities at the interface (Williams , 1998; Deshpande , 2012). On the other hand, it is crucial to maintain a sharp interface to ensure accurate numerical solutions that effectively capture the underlying physics of the interfacial flow. The most commonly used technique to circumvent this problem is to smooth the indicator function over a finite region around the fluid interface (Gerlach , 2006). If γ̂ is the resulting smoothed indicator function, the curvature is then computed as
(8)
The non-smoothed phase fraction γ is used for all other purposes than curvature calculation. Here, we have used a smoothing technique similar to the one proposed by Lafaurie (1994), which has already been implemented in several works (Hoang , 2013; Ferrari and Lunati, 2013) and consists of computing the smoothed indicator function at the center of mesh cell c as
(9)
where the subscripts c and f denote, respectively, the cell and face index, k is the number of faces of the finite volume mesh cell, and Sf is the area of face f. This smoothing process can be applied iteratively m times to obtain the desired level of smoothness. In this study, a value of m = 2 showed sufficiently good results in terms of suppression of parasitic velocities, which is also in agreement with the benchmark study done by Hoang (2013).

Note that this smoothing procedure is not included in the original OpenFOAM VOF solver, which thus had to be modified in this respect.

The numerical model presented above has been implemented in the open-source CFD software OpenFOAM (OpenFOAM, 2012). For all the flow domains under consideration, the 3D computational mesh consisting of purely hexahedral volume elements was generated using the blockMesh utility provided in OpenFOAM. The governing equations [Eqs. (1), (2), and (6)] were discretized spatially using the cell-centered-based finite volume technique (Berberović , 2009) with a second-order accuracy in space. The implicit Euler scheme was used to perform the time integration, resulting in first-order accuracy in time.

The solution procedure starts by solving the phase fraction advection equation [Eq. (6)] with the compression velocity term [Eq. (7)] to restrict numerical diffusion. The resulting phase fraction is then used to update the bulk density and viscosity [using Eq. (4)] and to calculate the surface tension forces using the CSF model as described in Sec. II A. To limit oscillations in the surface tension force calculation and to suppress parasitic velocities, we have modified the CSF model of OpenFOAM to implement the smoothing of the phase fraction using Eq. (9) (Sec. II B). Finally, to solve for the pressure and the velocity field from the momentum equation [Eq. (2)] and the continuity equation (1), the Pressure Implicit with Splitting Operators (PISO) algorithm by Issa (1986) is used. For further details on the implementation of the PISO algorithm and the finite volume discretization of the governing equations in OpenFOAM, see Deshpande (2012), Berberović (2009), and Rusche (2003) and the supplementary material. For all simulations, the initial time step is set to a very low value (108 s), with a run-time adjustable time step scheme, which automatically adjusts the time steps based on the Courant–Friedrichs–Lewy (CFL) criterion (Rusche, 2003). The maximum CFL number for all the cases was set to a value of 0.5. All numerical simulations were conducted in parallel on the cluster system at Leibniz University of Hannover, Germany.

To validate the two-phase VOF model described in Sec. II, we consider the immiscible displacement of a wetting fluid, which initially saturates the flow cell, by a non-wetting fluid (i.e., primary drainage), in a long Hele–Shaw cell. This cell is thus a rectangular cuboid flow cell with one of its dimensions transverse to the mean flow that is very small compared to the other two dimensions. The Hele–Shaw cell (or channel) can be considered the limit model for a perfectly smooth fracture. While smooth-walled fractures are unrealistic, using Hele–Shaw cells offers valuable insights into processes that could occur within more complex fracture geometries (Meakin and Tartakovsky, 2009). In this primary drainage configuration, the displacing fluid has a lower viscosity than the displaced fluid, so the process controlling the dynamics of the fluid–fluid interface is viscous instability, which leads to the formation of a fingerlike pattern, first studied experimentally and theoretically in the seminal work by Saffman and Taylor (1958). The most important observation in their experiments was the emergence of a single dominant finger in the channel, whose width's ratio to the channel width, λ, is close to 0.5 unless the flow is very slow. At very low mean flow velocities, the measured values of λ diverge from 0.5, going to the limiting case of 1.0 as the mean velocity tends to zero.

The numerical setup for our validation case (shown in Fig. 1) is inspired by the original experimental configuration used by Saffman and Taylor (1958), with slight modifications on account of the computational effort and time, e.g., the width to thickness ratio for our case is Ly/b=31.25, which is very close to the original experimental setup (31.75). The channel length is 10 cm, which is 8 times its width. We impose an initial perturbation on the fluid interface, as shown in Fig. 1, in order to suppress the growth of smaller fingers and accelerate the propagation of one single dominant finger in the channel. We designate the two fluids as fluid 1 and fluid 2, where fluid 1 is the invading, less viscous non-wetting fluid, while fluid 2 is the defending, more viscous fluid, wetting the channel walls. The material properties for the pair of fluids are listed in Table I.

FIG. 1.

Schematic of the long Hele–Shaw cell used to validate the resolved-film (RF) and unresolved-film (UF) numerical simulations. The dimensions of the cell are as follows: length Lx = 100 mm (x[0,Lx]), width Ly=12.5 mm (y[Ly/2,Ly/2]), while the constant vertical aperture is b = 0.4 mm (z[0,b]). The width-to-aperture ratio is thus Ly/b=31.25. The initial (t = 0) distribution of the two phases is shown, with the dark region indicating the invading fluid (fluid 1) and the lighter region indicating the defending fluid (fluid 2). The initial fluid–fluid interface is not flat, but its horizontal profile is Gaussian.

FIG. 1.

Schematic of the long Hele–Shaw cell used to validate the resolved-film (RF) and unresolved-film (UF) numerical simulations. The dimensions of the cell are as follows: length Lx = 100 mm (x[0,Lx]), width Ly=12.5 mm (y[Ly/2,Ly/2]), while the constant vertical aperture is b = 0.4 mm (z[0,b]). The width-to-aperture ratio is thus Ly/b=31.25. The initial (t = 0) distribution of the two phases is shown, with the dark region indicating the invading fluid (fluid 1) and the lighter region indicating the defending fluid (fluid 2). The initial fluid–fluid interface is not flat, but its horizontal profile is Gaussian.

Close modal
TABLE I.

Fluid properties and flow parameters used for the Hele–Shaw channel simulations.

Non-wetting contact angle (θ 135 
Surface tension σ (N/m)  0.015 
Fluid densities ρ1, ρ2 (kg m−3 100, 87.5 
Fluid viscosities μ1, μ2 (kg m−1 s−1 0.005, 0.15 
Inlet injection velocity U (mm s−1 1.0, 1.5, 2.0, 3.5, 5.0, 6.0, 7.5, 10.0, 12.5 
Reynolds numbers Re/103  8, 12, 16, 28, 40, 48, 60, 80, 100 
Capillary numbers log(Ca)  3.48, 3.30, 3.18, 2.93, 2.78, 2.70, 2.60, 2.48, 2.38 
Non-wetting contact angle (θ 135 
Surface tension σ (N/m)  0.015 
Fluid densities ρ1, ρ2 (kg m−3 100, 87.5 
Fluid viscosities μ1, μ2 (kg m−1 s−1 0.005, 0.15 
Inlet injection velocity U (mm s−1 1.0, 1.5, 2.0, 3.5, 5.0, 6.0, 7.5, 10.0, 12.5 
Reynolds numbers Re/103  8, 12, 16, 28, 40, 48, 60, 80, 100 
Capillary numbers log(Ca)  3.48, 3.30, 3.18, 2.93, 2.78, 2.70, 2.60, 2.48, 2.38 

Depending on the prescribed injection velocities of the invading fluid 1 at the inlet, u(x=0,y,z)=(U,0,0), and using the definitions of the Reynolds number as Re=ρ1Ub/μ1, which quantifies the nonlinearity of the flow, and capillary number as Ca=μ1U/σ, which is an estimate of the typical ratio of the viscous forces' magnitude to that of the capillary forces, we obtain a range of different flow configurations as listed in Table I. The viscosity ratio M=μ1/μ2 is equal to 1/30, which is the value used in Saffman and Taylor's experiments where water penetrates an oil of higher viscosity. The boundary conditions that are imposed on the different domain boundaries (Fig. 1) are listed in Table II.

TABLE II.

Boundary conditions used for all numerical simulations. nb is the normal to the boundaries other than solid walls, namely, the inlet and outlet.

Boundary u(x,y,z) p(x,y,z) γ(z,y,z)
Inlet  (U,0,0)  nb·p=0  γ = 1 
Outlet  nb·u=0  p = 0  nb·γ=0 
Side, top, and bottom walls  (0, 0, 0)  ns·p=0  θ=135° 
Boundary u(x,y,z) p(x,y,z) γ(z,y,z)
Inlet  (U,0,0)  nb·p=0  γ = 1 
Outlet  nb·u=0  p = 0  nb·γ=0 
Side, top, and bottom walls  (0, 0, 0)  ns·p=0  θ=135° 

The numerically induced slip presents a challenge for achieving a truly grid-independent solution with the VOF model. Therefore, in our grid convergence study, our aim is to reasonably predict the two-phase flow behavior while considering computational constraints. We conduct two sets of numerical simulations, which differ in how they handle the film of the defending fluid attached to the top and bottom walls of the Hele–Shaw channel. The first set of simulations, which we denote “resolved-film” (RF), resolves that film, while the other set, denoted “unresolved-film” (UF), does not. The corresponding computational meshes used for these two configurations are shown in Fig. 2. The RF mesh [Fig. 2(a)] has a much higher number of cells in the z direction, with a uniform grading near the walls, while the UF mesh [Fig. 2(b)] has cells of uniform thickness in the z direction, equal to their horizontal size. Previous studies (Horgue , 2012, 2013) indicate that to capture the wetting fluid film accurately, it should be covered by at least two grid cells, and that having more than two cells in the liquid film does not significantly affect solution accuracy. Our simulations, conducted with varying vertical resolutions in a reduced domain, confirm this. However, it is essential to note that the vertical resolution at the walls impacts the numerical slip of the contact line, which can influence other flow behaviors, such as the finger growth velocity.

FIG. 2.

Computational mesh for the Hele–Shaw domain with a horizontal mesh size of Δx=b/8. (a) Resolved-film (RF) mesh: The thickness of the cells at the top and bottom walls (z=0,b) is 0.3 μm and that of cells close to the central horizontal plane (z=b/2) is 30 μm, with a cell-to-cell expansion ratio of 1.2, resulting in a total cell count of 250×2000×40=2.0×107. (b) Unresolved-film (UF) mesh: All cells have a constant thickness equal to their horizontal size, Δx=b/8, the total cell count thus being 250×2000×8=4.0×106.

FIG. 2.

Computational mesh for the Hele–Shaw domain with a horizontal mesh size of Δx=b/8. (a) Resolved-film (RF) mesh: The thickness of the cells at the top and bottom walls (z=0,b) is 0.3 μm and that of cells close to the central horizontal plane (z=b/2) is 30 μm, with a cell-to-cell expansion ratio of 1.2, resulting in a total cell count of 250×2000×40=2.0×107. (b) Unresolved-film (UF) mesh: All cells have a constant thickness equal to their horizontal size, Δx=b/8, the total cell count thus being 250×2000×8=4.0×106.

Close modal

The horizontal grid resolution is defined by the uniform horizontal mesh size Δx=Δy, which is chosen as a fraction of the constant aperture b: Δx=b/4,b/8, or b/16, depending on the simulation runs. For both RF and UF cases, these three levels of horizontal discretization were investigated while maintaining a constant vertical resolution. Figure 3 illustrates the impact of the horizontal discretization on finger evolution for the UF case with Re = 0.06 and log(Ca)=2.60. For the coarse grid (b/Δx=4), corresponding to the dashed curves in Fig. 3, insufficient grid resolution leads to numerical inaccuracies, preventing the formation of a stable finger. As the grid is refined to b/Δx=8 and b/Δx=12, a stable dominant finger evolution is observed (solid curves in Fig. 3). Similar findings regarding the impact of the horizontal grid resolution were obtained for the RF simulations (not shown here). Additionally, pressure drops ΔPio=PinPout across medium's length were recorded at fixed times. The pressure drops for b/Δx=8 and b/Δx=12 differed by less than 2%. Similar trends were observed for the RF simulations. Consequently, we have selected a horizontal discretization level of b/Δx=8 for all our simulations.

FIG. 3.

Effect of horizontal (xy) grid refinement for the unresolved-film (UF) case with flow conditions U=7.5×103 ms–1, log(Ca)=2.60, and Re = 0.06. The evolution of the fluid–fluid interface in the horizontal mid-plane cross section is shown at three different times, t0=0s,t1=1.6s, and t2=2.64 s for two levels of discretization: b/Δx=4 (dashed curves) and b/Δx=8 (solid curves). The case b/Δx=12 yields a shape identical to that of b/Δx=8 and hence is not shown here. Both computational meshes have eight cells in the vertical z direction, of identical thickness.

FIG. 3.

Effect of horizontal (xy) grid refinement for the unresolved-film (UF) case with flow conditions U=7.5×103 ms–1, log(Ca)=2.60, and Re = 0.06. The evolution of the fluid–fluid interface in the horizontal mid-plane cross section is shown at three different times, t0=0s,t1=1.6s, and t2=2.64 s for two levels of discretization: b/Δx=4 (dashed curves) and b/Δx=8 (solid curves). The case b/Δx=12 yields a shape identical to that of b/Δx=8 and hence is not shown here. Both computational meshes have eight cells in the vertical z direction, of identical thickness.

Close modal

We shall now compare the results of numerical simulations conducted for the immiscible two-phase displacement in the long Hele–Shaw channel. We specifically focus on two aspects: (i) comparing the numerical results of both RF and UF cases with the experimental observations and theoretical findings of Saffman and Taylor (1958) and (ii) examining the differences between the results of the RF and UF cases, particularly in predicting the macroscopic flow properties.

1. Wetting film and contact line dynamics

We first study the differences in invasion dynamics between the RF and UF configurations. Naturally, it is anticipated that the RF simulations are expected to better describe the physics at play and thus better predict the experimental findings, but they are more costly computationally. Hence, we want to test the comparative predictive efficiency of UF simulations.

Figure 4 depicts the initial stage of the fluid–fluid interface's displacement and highlights a crucial difference between the RF and UF predictions concerning the dynamics of the contact line at the top and bottom walls. In the RF case (solid curves in Fig. 4), where refined grid cells are employed at the walls, the contact line hardly moves, resulting in the formation of a distinct liquid film of the displaced wetting fluid on the top and bottom walls. This is what is observed experimentally. Conversely, in the UF case (dashed curves in Fig. 4), the much coarser vertical grid resolution in the vicinity of the walls causes the contact line to slip, preventing film development due to the fact that the vertical resolution is too coarse for the film to be resolved by the numerical simulation.

FIG. 4.

Vertical longitudinal cross section of the flow domain at y = 0 for corresponding RF (solid curves) and UF (dashed curves) simulations. The flow conditions are as follows: U=2.0×103 ms–1, log(Ca)=3.18, and Re = 0.016. The black lines represent the initial (t0=0s) intersection between the fluid–fluid interface and the cross-section plane, while the colored lines represent the same information at times t1=0.04 s, t2=0.08 s, and t3=0.12 s. The intersection between the cross-section plane and the contact lines on the top and bottom walls are the two points at which the fluid–fluid interface (in color) reaches the walls (z/b=0 and z/b=1).

FIG. 4.

Vertical longitudinal cross section of the flow domain at y = 0 for corresponding RF (solid curves) and UF (dashed curves) simulations. The flow conditions are as follows: U=2.0×103 ms–1, log(Ca)=3.18, and Re = 0.016. The black lines represent the initial (t0=0s) intersection between the fluid–fluid interface and the cross-section plane, while the colored lines represent the same information at times t1=0.04 s, t2=0.08 s, and t3=0.12 s. The intersection between the cross-section plane and the contact lines on the top and bottom walls are the two points at which the fluid–fluid interface (in color) reaches the walls (z/b=0 and z/b=1).

Close modal

Figure 5 displays the fluid–fluid interface in vertical cross sections of the channel, both longitudinal [Fig. 5(a)] and transverse [Fig. 5(b)] to the mean flow direction, for three Ca values: high, low, and intermediate (Table I). The defending fluid's film thickness decreases with lower capillary numbers (Ca), which is consistent with prior numerical and experimental studies (Bretherton, 1961; Horgue , 2012; and Rosengarten , 2006). Notably, when examining the fluid–fluid interface in the transverse vertical plane [Fig. 5(b)], upward bulges become apparent at the four corners, alongside a comparatively flatter interface at the center. Previous experimental and theoretical studies on viscous fingering in conventional Hele–Shaw cells (Tabeling and Libchaber, 1986; Jensen , 1987; Tabeling , 1987; Reinelt, 1987; and Tanveer, 2000) have well acknowledged the presence of such a thin wetting film and its distinct transverse shape. It is important to note that this detail is specifically captured in the RF configuration, which further strengthens our model validation efforts. However, a detailed analysis of the transverse shape and variable film thickness falls outside the scope of our current study. For readers interested in a more in-depth examination, the work of De Lozar (2008) may provide valuable insights.

FIG. 5.

Cross-sectional views of the fluid–fluid interface for three different capillary numbers showing the non-wetting fluid film at the bottom (z/b=0) and the top (z/b=1) walls: (a) longitudinal (y = 0 plane) and (b) transverse (x=Lx/2 plane) cross sections. The shape of the fluid–fluid interface is symmetric with respect to the vertical plane z/b=0.5. The defending fluid's film thicknesses measured at y = 0 for the three cases [ log(Ca)=2.48, –2.78, and –3.48] are 42.0, 31.0, and 15.0 μm, respectively; around the corners, near the bulge, they are 9.15, 6.22, and 2.91 μm, respectively.

FIG. 5.

Cross-sectional views of the fluid–fluid interface for three different capillary numbers showing the non-wetting fluid film at the bottom (z/b=0) and the top (z/b=1) walls: (a) longitudinal (y = 0 plane) and (b) transverse (x=Lx/2 plane) cross sections. The shape of the fluid–fluid interface is symmetric with respect to the vertical plane z/b=0.5. The defending fluid's film thicknesses measured at y = 0 for the three cases [ log(Ca)=2.48, –2.78, and –3.48] are 42.0, 31.0, and 15.0 μm, respectively; around the corners, near the bulge, they are 9.15, 6.22, and 2.91 μm, respectively.

Close modal

2. Dynamics of the displacement

The presence of a film of defending fluid on the horizontal walls reduces the effective aperture available to the invading fluid finger for flowing through the channel. Therefore, for a given injection flow rate, the mean velocity of the displacing fluid is necessarily larger for a RF simulation than for the corresponding UF simulation. This is reflected in the breakthrough time (t*), i.e., the time at which the displacing fluid reaches the outlet. In Fig. 6, we show the relative difference in breakthrough times between UF and RF simulations across the entire range of the investigated Ca values. RF simulations consistently predict lower breakthrough times compared to UF simulations. Importantly, as Ca decreases, the relative difference in breakthrough times diminishes, dropping to less than 5% for the lowest Ca values.

FIG. 6.

Dependence of the relative differences in breakthrough times between the UF and RF configurations, Δt*/tRF*=tUF*/tRF*1, on log(Ca).

FIG. 6.

Dependence of the relative differences in breakthrough times between the UF and RF configurations, Δt*/tRF*=tUF*/tRF*1, on log(Ca).

Close modal

Figure 7 illustrates the time dynamics of the mid-plane cross section of the fluid–fluid interface from t = 0 until 0.9t*, when the Hele–Shaw cell is initially filled by the wetting fluid, for both RF [Fig. 7(a)] and UF [Fig. 7(b)] simulations, with Re = 0.016 and log(Ca)=3.18. Both simulations exhibit a behavior consistent with experimental observations, with a singular stable finger of displacing fluid that exhibits a uniform width as it advances at constant velocity along the channel. Expectedly, the different color scales in Figs. 7(a) and 7(b) show that the finger in the RF simulation reaches the outlet earlier than that in the UF simulation. Furthermore, the differing shapes of the finger constrictions near the inlet illustrate the influence of the film and of the corresponding contact line motion. For the cases from Table I that are not shown here, similar observations are reported.

FIG. 7.

Time evolution of the fingering patterns for (a) RF and (b) UF simulation for U=2.0×103 m s–1, log(Ca)=3.18, and Re = 0.016. The colors from blue to yellow represent t/t*, the ratio of the time at which a region of the cell has been reached by the fluid–fluid interface to the corresponding breakthrough time, which is t*=23.6 s and t*=25.6 s for the shown RF and UF simulations, respectively.

FIG. 7.

Time evolution of the fingering patterns for (a) RF and (b) UF simulation for U=2.0×103 m s–1, log(Ca)=3.18, and Re = 0.016. The colors from blue to yellow represent t/t*, the ratio of the time at which a region of the cell has been reached by the fluid–fluid interface to the corresponding breakthrough time, which is t*=23.6 s and t*=25.6 s for the shown RF and UF simulations, respectively.

Close modal

3. Width and tip shape of the invading finger

In order to further assess the reliability of these results, we compare the width and tip shape of the simulated fingers with Saffman and Taylor's (1958) experimental observations of the finger width and theoretical prediction of the finger shapes for fingers of water displacing two different oils (Diala and Talpa). To determine the widths of our simulated fingers, we first obtain the intersection curve between the horizontal mid-plane of the Hele–Shaw cell and the fluid–fluid interface at time 0.9t*. The finger width is subsequently obtained as the mean width of the segment of the curve located in the region x/Lx[0.6,0.8].

We characterize the finger width by λ, which is its ratio to the Hele–Shaw channel width. In Fig. 8, we plot λ as a function of μ2Uf/σ, as presented by Saffman and Taylor (1958), where Uf is the finger-tip velocity of a stable finger. The black data points represent the historical experimental measurements involving direct and photographic assessments of finger positions and profiles, while the colored ones represent our simulation results. The colored data points in Fig. 8(b), correspond to our simulations, conducted at different levels of vertical discretization while maintaining the same horizontal grid refinement level of (b/Δx=8). As mentioned earlier in Sec. III B, vertical refinement of the RF simulations' numerical mesh has allowed us to reach grid convergence. All filled symbols indicate simulations for which we consider that grid convergence has been reached. At some values of Ca, unfilled symbols are shown, closely clustered together and with the filled symbol; they correspond to successive steps of the grid refinement procedure. Our investigations show that for the lowest Ca case, where the film thickness is approximately 3 μm, a grid distribution consisting of 40 cells in the vertical direction, a wall-adjacent cell thickness of 0.33 μm, and a cell-to-cell expansion ratio of 1.2 provide a sufficient resolution of the thin film. In contrast, the UF cases require a uniform vertical cell thickness, using the three levels of vertical discretization: b/Δz=4, 8, and 16.

FIG. 8.

Dependence of the finger width to channel width ratio, λ, on (a) the finger propagation velocity Uf and (b) the capillary number Ca. The experimental data in (a) (black symbols) are taken from Saffman and Taylor (1958); the circles and disks denote data for water displacing Diala oil (with λ measured in two manners, hence the two types of symbols), while the filled triangles denote data for water displacing Talpa oil; the dotted line denotes a best fit curve, also from Saffman and Taylor (1958). The blue star and red square symbols in both plots indicate, respectively, our RF and UF simulation results. When several symbols are present at a given Ca, they correspond to different vertical refinements of the numerical mesh; filled symbols corresponds to simulations for which grid convergence has been reached, while empty symbols correspond to intermediate steps of the grid refinement.

FIG. 8.

Dependence of the finger width to channel width ratio, λ, on (a) the finger propagation velocity Uf and (b) the capillary number Ca. The experimental data in (a) (black symbols) are taken from Saffman and Taylor (1958); the circles and disks denote data for water displacing Diala oil (with λ measured in two manners, hence the two types of symbols), while the filled triangles denote data for water displacing Talpa oil; the dotted line denotes a best fit curve, also from Saffman and Taylor (1958). The blue star and red square symbols in both plots indicate, respectively, our RF and UF simulation results. When several symbols are present at a given Ca, they correspond to different vertical refinements of the numerical mesh; filled symbols corresponds to simulations for which grid convergence has been reached, while empty symbols correspond to intermediate steps of the grid refinement.

Close modal

The RF numerical results closely match the experimental data across the entire range of investigated Ca, which points to an excellent validation of our OpenFOAM implementation of the VOF-based two-phase flow model. In comparison, the UF simulations tend to overpredict the finger width (λ) and accordingly yield lower finger velocity magnitudes (Uf) compared to both experimental data and the RF simulations [Fig. 8(a)]. These discrepancies are more noticeable when examining Fig. 8(b), where λ values are plotted against Ca, defined using the inlet velocity magnitude U. As expected, these differences diminish as Ca decreases. At higher Ca values, where a relatively thicker defending fluid film exists, the limited vertical resolution in the UF simulations fails to capture film effects. In the RF simulations, the contact line is pinned to the walls from the start of the injection [Fig. 4(a)], while the UF configuration shows contact line slippage [Fig. 4(b)], as mentioned in Sec. III C 1. Although minor in comparison to the longitudinal direction, contact line slippage also occurs in the transverse direction in the UF simulations. Consequently, with a given imposed flow rate, lateral contact line motion in UF simulation leads to an overestimation of λ, further resulting in a reduced finger velocity. As the wetting film thickness decreases with decreasing Ca, the film's influence diminishes, enhancing the agreement between UF and RF simulations.

Assuming a two-dimensional flow and neglecting surface tension, Saffman and Taylor (1958) derived a parametric equation for the shape of a single finger in the horizontal plane as
(10)
In Fig. 9, we compare the analytical profile to numerical simulations for the two extreme flow cases: log(Ca)=2.48,Re=0.08 and log(Ca)=3.48,Re=0.008. To calculate the analytical profile, for reference, we have used the λ values obtained from the RF simulations. The analytical profile closely matches the numerical results at higher Ca when surface tension effects are negligible for both RF and UF simulations [Fig. 9(a)]. However, at lower Ca, where surface tension dominates over inertial forces, there is a notable mismatch between the analytical and numerical profiles in both sets of simulations [Fig. 9(b)]. This is of course expected since the theoretical profile was obtained based on the assumption that the boundary condition at the fluid–fluid interface can be neglected (see Saffman and Taylor, 1958).
FIG. 9.

Comparison of the horizontal profile of the finger front obtained by numerical simulations (solid curves) with the analytical solution of Saffman and Taylor (1958) (red crosses). The flow conditions for plot (a) are: U=1.0×102 m s–1, log(Ca)=2.48, and Re = 0.08, and for plot (b), they are: U=1.0×103 m s–1, log(Ca)=3.48, and Re = 0.008. The fluid–fluid interface shown here is for the mid-plane (z=b/2) slice of the Hele–Shaw channel.

FIG. 9.

Comparison of the horizontal profile of the finger front obtained by numerical simulations (solid curves) with the analytical solution of Saffman and Taylor (1958) (red crosses). The flow conditions for plot (a) are: U=1.0×102 m s–1, log(Ca)=2.48, and Re = 0.08, and for plot (b), they are: U=1.0×103 m s–1, log(Ca)=3.48, and Re = 0.008. The fluid–fluid interface shown here is for the mid-plane (z=b/2) slice of the Hele–Shaw channel.

Close modal
We now focus on the macroscopic pressure gradient by comparing the UF solutions to the corresponding RF predictions. As shown in Table II, since the pressure at the outlet is set to 0 Pa, we can define an area-weighted average pressure drop ΔPio across the channel length (Ferrari and Lunati, 2013; Chen , 2018) as
(11)
where Γin is the inlet boundary and Ain its area. In Fig.10(a), ΔPio is plotted as a function of the global saturation S1 of the invading fluid, defined as (Ferrari and Lunati, 2013)
(12)
for four representative Ca cases from Table I. Note that S1 is simply a measure of the advancement of the displacement process; it is nearly proportional to time. As expected, a linear drop in the average pressure is observed as the less viscous invading fluid displaces the more viscous defending fluid and fills more space in the channel. In Fig. 10(b), the longitudinal pressure profile along the line (y=0,z=b/2) is plotted for the same data for the time at which the finger tips are at x0.6Lx. Two quasi-linear decreasing trends corresponding to the two phases are seen on the two sides of the fluid–fluid interface, which is indicated by a vertical drop corresponding to the capillary pressure drop across the interface. At high Ca, non-linearity in the inlet region (at x0.2Lx) results from the narrowing of the fingers (see Fig.7), and thus, capillary forces affect the pressure variations. It is worth noting that, in line with previous observations, the pressure drops in the UF simulations align with the RF results at lower Ca, primarily owing to the reduced film thickness as Ca decreases.
FIG. 10.

Longitudinal pressure drops: (a) dependence of the average pressure drop across the length of the Hele–Shaw channel, ΔPio, on the global saturation of the invading fluid, S1. (b) Longitudinal pressure (P) profile along the central line (y=0,z=b/2), between the inlet (x = 0) and the outlet (x = Lx ), considered at a time when the invading fluid finger has covered 60% of the channel length. The results are presented for four different cases ranging from high to low Ca, whose logarithms are given in the plot.

FIG. 10.

Longitudinal pressure drops: (a) dependence of the average pressure drop across the length of the Hele–Shaw channel, ΔPio, on the global saturation of the invading fluid, S1. (b) Longitudinal pressure (P) profile along the central line (y=0,z=b/2), between the inlet (x = 0) and the outlet (x = Lx ), considered at a time when the invading fluid finger has covered 60% of the channel length. The results are presented for four different cases ranging from high to low Ca, whose logarithms are given in the plot.

Close modal

The VOF model described in Sec. II and validated in Sec. III is now used to conduct numerical experiments on rough fracture geometries. Our objective is to compare the results from RF simulations to those from the corresponding UF simulations, aiming to elucidate how film resolution influences predictions of various flow-related quantities during drainage in a geological fracture. It allows us to quantify the prediction errors arising from the absence of resolved films. We aim to identify situations where such resolution is essential and those where, despite the lack of film resolution, predictions largely remain reliable. Additionally, this study highlights the trade-off between accuracy and computational expenses, guiding our decisions in selecting film-resolution approaches and strategies for drainage simulations within rough fractures.

Figures 11 and 12 show the schematics of the two numerically generated rough fractures that were used for the numerical simulations of drainage. The first fracture (Fig. 11) has an aperture field that varies sinusoidally (and is thus periodic); in the following, we shall refer to it as the sinusoidal fracture. The gradient of its aperture field is moderate (||a(x,y)||1). It is important to note that no actual fracture, regardless of the fracture material, will exhibit a geometry similar to this one. This fracture geometry serves as a transitional case of increased complexity from the parallel plate (Hele–Shaw) setup used as our validation (Sec. III) case to the realistic geometry shown in Fig. 12. This second fracture exhibits wall topographies that seem to be distributed at random (in particular, they are not periodic), but the fracture has statistical geometrical properties that are consistent with those of naturally occurring geological fractures. Its aperture field is also rougher at small scales (max(||a(x,y)||) ⪆ 1) than that of the sinusoidal fracture. Hereafter, we shall refer to that fracture as the geological fracture.

FIG. 11.

(a) Schematic of the sinusoidal fracture. The dimensions in the horizontal xy plane are Lx=Ly=0.05 m. The mean aperture is a¯=2.5×104 m and the mean inlet aperture is a¯in=3.53×104 m. The flat bottom surface lies at z = 0. The lateral walls of the fracture lie in the vertical xz planes at y = 0 and y = Ly, respectively. (b) Contour map of the aperture field a(x, y), also equal to the top surface topography: a(x,y)=A(sin(ωπy)+cos(ωπx))+c, where A=1.0×104, ω = 360, and c=2.5×104 m.

FIG. 11.

(a) Schematic of the sinusoidal fracture. The dimensions in the horizontal xy plane are Lx=Ly=0.05 m. The mean aperture is a¯=2.5×104 m and the mean inlet aperture is a¯in=3.53×104 m. The flat bottom surface lies at z = 0. The lateral walls of the fracture lie in the vertical xz planes at y = 0 and y = Ly, respectively. (b) Contour map of the aperture field a(x, y), also equal to the top surface topography: a(x,y)=A(sin(ωπy)+cos(ωπx))+c, where A=1.0×104, ω = 360, and c=2.5×104 m.

Close modal
FIG. 12.

(a) Schematic of the geological fracture with self-affine top and bottom walls of Hurst exponent H = 0.8. The dimensions in the horizontal xy plane are Lx=Ly=0.05 m. The correlation length is Lc=Lx/4=0.0125 m. The mean aperture is a¯=3.3×104 m, the standard deviation, σa=90μm, and the mean inlet aperture is a¯in=3.2×104 m. The lateral walls of the fracture lie in vertical xz plane at y = 0 and y = Ly, respectively. (b) Contour map of the aperture field a(x, y) of the fracture.

FIG. 12.

(a) Schematic of the geological fracture with self-affine top and bottom walls of Hurst exponent H = 0.8. The dimensions in the horizontal xy plane are Lx=Ly=0.05 m. The correlation length is Lc=Lx/4=0.0125 m. The mean aperture is a¯=3.3×104 m, the standard deviation, σa=90μm, and the mean inlet aperture is a¯in=3.2×104 m. The lateral walls of the fracture lie in vertical xz plane at y = 0 and y = Ly, respectively. (b) Contour map of the aperture field a(x, y) of the fracture.

Close modal

Both fractures consist of a top and a bottom wall whose horizontal mean planes are parallel and separated by a distance a¯, which is thus the mean aperture of the fracture (the walls are never in contact). In the case of the sinusoidal fracture, the bottom surface coincides with the plane z = 0, while the top surface is represented by a two-dimensional sinusoidal function (see Fig. 11). On the other hand, the top and bottom walls of the geological fracture are self-affine isotropic topographies, which exhibit long-range correlations (Bouchaud, 1997; Plouraboué , 1995) characterized by the Hurst exponent H (here H = 0.8). The two walls have identical large-scale variations above a characteristic length Lc, which we denote correlation length. In addition to (i) a¯, (ii) H, and (iii) Lc, the aperture field of such geological fractures is defined primarily by (iv) its standard deviation σa and (iv) the horizontal dimensions Lx and Ly of the fracture (Méheust and Schmittbuhl, 2003; Dewangan , 2022). For details on the geometrical properties of such geological fractures, see also Lenci (2022b) and Dewangan (2022). The fracture wall topographies of Fig. 12 are generated using a spectral method initially proposed by Méheust and Schmittbuhl (2003). The generation of the correlated random fields corresponding to the wall topographies is done by first generating two independent self-affine walls of zero mean and identical standard deviation, h+ and h; then attributing the same large Fourier modes (for wave number larger than 2π/Lc), taken from h+, to h (or conversely); then imposing to both a new standard deviation such that the h+h has the desired standard deviation σa; and finally adding a¯/2 to h+ and subtracting a¯/2 from h,a¯ being the desired mean aperture.

For both fracture geometries, we simulate the same drainage configuration, described in Sec. III, where a less viscous fluid 1 displaces a fluid 2 of higher viscosity (see Table III). To accurately capture the onset of instability of the fluid–fluid interface, we initialize the interface (flat) up to a small distance of 0.0025 m  =Lx/20 from the inlet, meaning at t = 0 fluid 1 occupies a small volume of the fracture domain. Here we use the same nomenclature for the different domain boundaries (see Figs. 11 and 12) as used in Sec. III A for the Hele–Shaw channel, and thus the boundary conditions for our simulation setup are the same as prescribed in Table II. We investigate five different cases of Ca which span over three orders of magnitude: log10Ca=3.0,3.5,4.0,4.5,5.0. Table III lists the inlet velocities U=σCa/μ1 obtained for these five cases and the resulting Re=ρ1Uina¯in/μ1 (we have used inlet mean aperture a¯in to calculate Re) for both the sinusoidal and geological fractures.

TABLE III.

Fluid properties and flow parameters used for the two rough fracture simulations.

Non-wetting contact angle (θ 135 
Surface tension σ (N/m)  1.0 
Fluid densities ρ1, ρ2 (kg m−3 100, 90 
Fluid viscosities μ1, μ2 (kg m−1 s−1 0.01, 1.0 
Capillary numbers log(Ca)  3.0, 3.5, 4.0, 4.5, 5.0 
Inlet injection velocities U (mm s−1 100, 31.6, 10, 3.16, 1.0 
Reynolds numbers Re/103 for:   
the sinusoidal fracture  350, 110, 35, 11, 3.5 
the geological fracture  324, 102, 32.4, 10.2, 3.24 
Non-wetting contact angle (θ 135 
Surface tension σ (N/m)  1.0 
Fluid densities ρ1, ρ2 (kg m−3 100, 90 
Fluid viscosities μ1, μ2 (kg m−1 s−1 0.01, 1.0 
Capillary numbers log(Ca)  3.0, 3.5, 4.0, 4.5, 5.0 
Inlet injection velocities U (mm s−1 100, 31.6, 10, 3.16, 1.0 
Reynolds numbers Re/103 for:   
the sinusoidal fracture  350, 110, 35, 11, 3.5 
the geological fracture  324, 102, 32.4, 10.2, 3.24 

We again conduct two sets of simulations, namely with resolved-film (RF) and unresolved-film (UF) for both the sinusoidal and geological fracture geometries. The computational mesh employed for our numerical calculations is shown in Fig. 13, where, similarly to our validation on Hele–Shaw geometry (Sec. III), the RF simulations have a substantially higher degree of resolution in the vertical direction than the UF simulations. To ensure grid independence in our numerical simulations, we used our findings from the Hele–Shaw validation case (Sec. III B) and chose three horizontal discretization levels: Δx=100μm, Δx=50μm, and Δx=25μm. We then performed grid sensitivity analyses on a smaller domain of the fracture geometries while maintaining a fixed vertical resolution. For both geometries and the corresponding RF and UF configurations, we found that the difference between the results corresponding to the discretization levels of Δx=50μm and Δx=25μm (e.g., average pressure drops ΔPio) were less than 1%–2%. Consequently, for all the simulations, we have chosen a horizontal discretization corresponding to Δx=50μm, which results in 1000 × 1000 cells in the horizontal plane of the fracture geometries.

FIG. 13.

Computational meshes for the sinusoidal fracture (a) and (b) and geological fracture (c) and (d), both with a horizontal mesh cell size of Δx=50μm. The computational domain for the RF simulation (a) and (c) has a cell count of 1000×1000×32=3.2×107, while that for the UF simulations (b) and (d) consists of 1000×1000×8=8×106 cells.

FIG. 13.

Computational meshes for the sinusoidal fracture (a) and (b) and geological fracture (c) and (d), both with a horizontal mesh cell size of Δx=50μm. The computational domain for the RF simulation (a) and (c) has a cell count of 1000×1000×32=3.2×107, while that for the UF simulations (b) and (d) consists of 1000×1000×8=8×106 cells.

Close modal

As previously indicated in Secs. III B and III C, ensuring convergence, particularly in the vertical resolution, poses challenges. Furthermore, with rough fractures, whose apertures vary spatially, excessively large grid cell numbers in the z direction are impractical due to the emergence of tiny volume cells in small aperture regions. For RF simulations, numerical experiments on reduced domains with varying numbers of cells in the z direction suggested that for both geometries, 32 uniformly graded cells (1.33 cell-to-cell expansion ratio) sufficed to resolve wall-adjacent films, even at the smallest capillary numbers. This arrangement results in mean cell thicknesses of 0.33 and 0.39 μm for the sinusoidal and geological fractures, respectively. Conversely, considering the dependence of wetting-film thickness on the Ca, we conducted analogous UF numerical experiments (i.e., without vertical grading in the z direction). For higher Ca values [ log(Ca)3.5], we used eight cells; for lower Ca cases [ log(Ca)<3.5], 16 cells were considered.

1. Invasion morphologies

The fluid displacement morphology, commonly referred to as invasion patterns, determines the macroscopic residual saturations and thus, for example, the sweep efficiency. It is controlled by the interplay between viscous and capillary forces, and possibly gravity, as well as the local aperture conditions. When the interface is viscously unstable (M < 1), this interplay results in fingerlike patterns. In porous media, if capillary forces are dominant (low Ca), the so-called capillary fingering occurs, which features patterns resulting from growth occurring at various places along the fluid–fluid interfaces, depending on where the capillary forces are more favorable to the interfaces' displacement (Lenormand and Zarcone, 1989). Conversely, at high Ca values, viscous fingering prevails, characterized by more branched fingers (Måløy 1985). A similar behavior is observed in rough fractures (Chen , 2017), but the phase diagrams of the flow patterns as a function of Ca and M are not necessarily identical to those known for porous media (Lenormand, 1990) and remain an open question. Figures 14 and 15 depict the displacement morphologies for the sinusoidal and geological fractures for different capillary numbers. The images display superimposed invasion patterns, using distinct colors to show agreement (yellow) or discrepancy (blue and red) between patterns from RF and UF simulations. In line with what was observed for the Hele–Shaw configuration (Sec. III), finger advancement time differs between film and without-film cases. To emphasize the patterns, the snapshots for RF and UF simulations are taken at different time steps to maintain approximately equidistant positions from the outlet.

FIG. 14.

Superimposed invasion morphologies obtained in the sinusoidal fracture geometry with RF and UF simulations under the same flow conditions. The three columns correspond to invasion patterns recorded at three different times at which, for both types of simulations and for all Ca, the invading tip is located at around the same distance from the outlet boundary. The rows correspond to different values of the capillary numbers: logCa=3.0(aa),3.5(bb),4.0(cc),4.5(dd),5.0(ee).

FIG. 14.

Superimposed invasion morphologies obtained in the sinusoidal fracture geometry with RF and UF simulations under the same flow conditions. The three columns correspond to invasion patterns recorded at three different times at which, for both types of simulations and for all Ca, the invading tip is located at around the same distance from the outlet boundary. The rows correspond to different values of the capillary numbers: logCa=3.0(aa),3.5(bb),4.0(cc),4.5(dd),5.0(ee).

Close modal
FIG. 15.

Superimposed invasion morphologies obtained in the geological fracture geometry with RF and UF simulations under the same flow conditions. The three columns correspond to invasion patterns recorded at three different times at which, for both types of simulations and for all Ca, the invading tip is located at around the same distance from the outlet boundary. The rows correspond to different values of the capillary numbers: logCa=3.0(aa),3.5(bb),4.0(cc),4.5(dd),5.0(ee).

FIG. 15.

Superimposed invasion morphologies obtained in the geological fracture geometry with RF and UF simulations under the same flow conditions. The three columns correspond to invasion patterns recorded at three different times at which, for both types of simulations and for all Ca, the invading tip is located at around the same distance from the outlet boundary. The rows correspond to different values of the capillary numbers: logCa=3.0(aa),3.5(bb),4.0(cc),4.5(dd),5.0(ee).

Close modal

For the sinusoidal fracture case (Fig. 14), an immediate observation is the presence of longitudinal preferential paths of advancement in the invasion front's progression at all capillary numbers. These pathways form because the initial invasion is easier along these paths as they feature a large aperture zone in contact with the fracture inlet. Once these high aperture regions have been invaded, parallel fingers have been initiated, which will then retain an advantage and thus prevent the fluid from invading other areas in the vicinity of the inlet. Hence, only these fingers grow. Due to the viscous instability of the interface, a finger that gets ahead of another neighboring one at a given time will keep this advantage at later times, so that the differences in finger propagation increase in time. For the geological fracture (Fig. 15), where aperture variations are no longer periodic but disordered, with a correlation length of the apertures that is a quarter of the fracture size, the same phenomenon will lead to the formation of a smaller number of disordered fingers exhibiting irregular patterns resembling those seen in natural fractures or their laboratory replicas, as reported in previous studies (e.g., Chen , 2017, 2018).

For both fracture geometries, a noticeable broadening of the fingers in the transverse direction is seen as the Ca decreases. Additionally, when the Ca is decreased, the capillary forces start impacting the invasion process more strongly, so that increased local surface tension effects lead to stronger shielding of the least advanced fingers' growth by the most advanced fingers, thus reducing the number of fingers that proceed effectively toward the outlet. These features are reproduced with both RF and UF simulations. The enhanced shielding mechanism at low Ca is more prominent in the sinusoidal fracture, due to the fact that the various fingers encounter the same geometries as they travel forward, so that no stochastic geometry heterogeneity competes with the shielding mechanism to control the differential propagation of fingers. In this geometry, and at the two lower Ca values of logCa=4.5 and –5.0, a considerable disagreement is thus observed between UF and RF invasion patterns, as UF simulations lack the ability to capture the full 3D effects of dominant surface tension forces. This effect is much less present in the geological fracture, due to the stochastic heterogeneity of the aperture field, as discussed above.

At higher Ca values [ log(Ca)>4.0], discrepancies between the invasion patterns predominantly appear in the figures as extended blue regions (RF) aligned with the flow direction, for both geometries. These differences tend to become more pronounced in the later stages of the invasion process [e.g., Figs. 14(a″) and 15(a″)]. The relatively thick wetting film at high Ca limits the effective aperture available for the invading fluid, accelerating the movement of the advancing fluid fingertips. The widening of the fingers in the transverse direction, though present in both RF and UF simulation results across all Ca cases, is predominately more marked in the UF cases. This is consistent with our observations in the case of a single finger propagating in the Hele–Shaw channel, where UF simulations yielded finger widths larger than those of the RF simulations for the entire range of Ca. Analogously, the lateral slip of contact lines in the UF simulations translates to slightly broader finger widths compared to the corresponding RF simulations. This is also depicted in Fig. 16, illustrating the fluid–fluid interface in small cross sections of the two fracture geometries.

FIG. 16.

Fluid–fluid interface in vertical cross sections of the sinusoidal (a) and (b) and geological (c) and (d) fractures. (a) and (c) correspond to a longitudinal cross section at y=Ly/2, while (b) and (d) represent a transverse cross section at x=Lx/5. The times at which the results are plotted are (a) and (b) t = 0.094 s and t = 0.122 s for the RF and UF simulations, respectively, and (c) and (d) t = 0.116 s and t = 0.152 s for the RF and UF simulations, respectively.

FIG. 16.

Fluid–fluid interface in vertical cross sections of the sinusoidal (a) and (b) and geological (c) and (d) fractures. (a) and (c) correspond to a longitudinal cross section at y=Ly/2, while (b) and (d) represent a transverse cross section at x=Lx/5. The times at which the results are plotted are (a) and (b) t = 0.094 s and t = 0.122 s for the RF and UF simulations, respectively, and (c) and (d) t = 0.116 s and t = 0.152 s for the RF and UF simulations, respectively.

Close modal

In the sinusoidal fracture at the two lower Ca values [ log(Ca)=4.5 in Fig. 14(d-d″) and log(Ca)=5.0 in Fig. 14(e-e″)], we also observe cutting-off of the invading fluid, both in the RF and UF simulations, by pinch-off in low aperture areas. As shown in Fig. 17 for log(Ca)=4.5, this occurs in fingers that are left behind and is more frequent and quicker in the RF simulations. In the latter configurations, such pinch-off events are associated with later re-merging of these isolated clusters with the base of the short finger from which they had separated; this is not observed in UF simulations. Further details are provided in the  Appendix, with additional pictures of the flow patterns for each of the flow configurations, and links to online videos of the corresponding experiments (Multimedia available online).

FIG. 17.

(a) Superimposed images for the invasion pattern corresponding to a capillary number of log(Ca)=4.5 for the sinusoidal fracture [Fig. 14(d″)]. (b) and (c) are close-up views of the patterns obtained from the UF (red) and RF (blue) simulations, respectively. (d) Dependence of the relative deterministic discrepancy between the UF and RF simulations at breakthrough time, characterized in terms of the portion A of the fracture plane occupied by the displacing fluid.

FIG. 17.

(a) Superimposed images for the invasion pattern corresponding to a capillary number of log(Ca)=4.5 for the sinusoidal fracture [Fig. 14(d″)]. (b) and (c) are close-up views of the patterns obtained from the UF (red) and RF (blue) simulations, respectively. (d) Dependence of the relative deterministic discrepancy between the UF and RF simulations at breakthrough time, characterized in terms of the portion A of the fracture plane occupied by the displacing fluid.

Close modal

The overall relative deterministic discrepancy between the UF and RF simulations can be quantified from the ratio δA/A of the joint area of the blue and red regions in Figs. 14 and 15 (for the sinusoidal and geological fractures, respectively) to the area A of the region of the fracture plane that is occupied by the displacing fluid in the RF simulation (i.e., the joint area of the blue and yellow regions in Figs. 14 and 15). This ratio is plotted at breakthrough time in Fig. 17(d), as a function of the capillary number and for the two types of fracture geometries. Its dependence on Ca is well explained by the phenomenology discussed above. In particular, at the lowest capillary number (logCa=5), the aforementioned shielding of less advanced fingers, coupled with the marked effect of capillary forces, leads to fingers partially developing at different places in the fracture plane in the UF simulations, as compared to the RF simulations, and thus to δA/A values between 20% and 30%. At the largest capillary number (logCa=5), on the contrary, the interface's enhanced instability leads to a similar discrepancy and δA/A values as high as 50% for the geological fracture. At intermediate Ca values, for which none of these two effects is strong, the δA/A values are typically close to 10%. Note that deterministic discrepancies between the two displacement patterns that are larger than 10%, essentially result from finger branches developing at different places in the UF and RF simulations, which is expected due to the unstable nature of the fluid–fluid interface and the relatively ordered geometries. We shall see that comparing quantities that correspond to more statistical measures of the flow behavior (e.g., breakthrough times, fluid–fluid interface length, longitudinal pressure profiles, longitudinal saturation profiles) will yield much smaller discrepancies between the invasion patterns predicted by the UF and RF simulations. This was also observed, for similar reasons, by Ferrari (2015) when comparing numerical predictions to experimental measurements of two-phase flow in random two-dimensional porous media.

2. Film thickness in the RF simulations

For the two fracture geometries, an essential observation emerges from the RF simulations when examining the respective film thicknesses. We estimate the film thickness at a given position along the fracture plane in the following manner. We first depth-average the indicator function γ(x,y,z) to obtain a two-dimensional saturation field s1(x,y) for the invading fluid 1, and thence the corresponding depth-averaged saturation of the defending fluid film, s2(x,y)=1s1(x,y). This 2D saturation s2 is either 1, in the portion of the fracture plane where no invading fluid is present, or significantly smaller than 1 in a region that we denote Ω. The value of s2 in Ω results from the presence of the two films of displaced fluid (on each of the fracture walls), and the average film thickness can be estimated as the average of (s2×a)/2 over Ω, which we write as s2aΩ.

Note that this estimate of the mean film thickness tends to slightly decrease in time. Its average over the three different times at which the invasion patterns of Figs. 14 and 15 are shown is plotted in Fig. 18 as a function of the capillary number for the two fracture geometries. For both geometries, it exhibits a power law behavior close to Ca1/3, with values ranging from 2.5 to 25 μm. In the figure, we also indicate the Bretherton law for wetting film thickness measured when a gas displaces a liquid (Bretherton, 1961); it is not expected to hold here due to the much lower viscosity ratio between the two fluids. Across all Ca, the sinusoidal fracture consistently exhibits larger film thickness than the geological fracture, but the thickness ratio between the two geometries is overall all the smaller as Ca is larger, reaching 1 at Ca=103. At Ca=105, it is larger than 2. A detailed study of the films' temporal and spatial fluctuation is out of the scope of the present study.

FIG. 18.

Dependence of the mean wetting film thickness on the capillary number, for the two fracture geometries. The power law of exponent 1/3 is consistent with most of the data, while that of exponent 2/3 corresponds to the Bretherton law.

FIG. 18.

Dependence of the mean wetting film thickness on the capillary number, for the two fracture geometries. The power law of exponent 1/3 is consistent with most of the data, while that of exponent 2/3 corresponds to the Bretherton law.

Close modal

3. Breakthrough times

Two key observations emerge when examining the relative differences between breakthrough times t* in the UF simulations and RF simulations, Δt*=tUF*tRF*. They are plotted against log(Ca) values in Fig. 19 for both fracture geometries. First, Δt* is always positive, which means that the UF simulations always overpredict the breakthrough times, i.e., underpredict the finger-tip propagation. Second, decreasing Ca leads to diminishing the relative differences in t* for both fracture geometries, which aligns with our Hele–Shaw cell validation findings (Sec. III). Third, the relative differences in t* for the sinusoidal fracture geometry, while comparable to those obtained for the Hele–Shaw cell case (see Fig. 6), are considerably larger (nearly one order of magnitude) than those for the geological fracture. This cannot be only explained by the differences in wetting film thickness presented in Sec. IV D 2. For the sinusoidal fracture in particular, the widening of the invading fluid pattern plays an important role in slowing down finger-tip propagation. In addition, not resolving the wetting films leads to the predominant propagation of a different finger in Fig. 17(a), which can result in a significant discrepancy in the breakthrough times, whereas for the geological fracture, the finger that propagates faster is the same in the UF and RF simulations.

FIG. 19.

Plot of the relative differences between the breakthrough times Δt*=tUF*− tRF* of the RF and UF simulations against log(Ca).

FIG. 19.

Plot of the relative differences between the breakthrough times Δt*=tUF*− tRF* of the RF and UF simulations against log(Ca).

Close modal

4. Average pressure drops

For the evolution of fluid pressure, we examine the variation of the area-weighted average pressure drop ΔPio across the inlet and outlet of the fracture geometries with the invading fluid saturation S1 (Fig. 20). Higher capillary numbers result in elevated and relatively steady pressure drops due to increased viscous forces. At lower Ca values, where capillary forces become dominant, considerable fluctuations in pressure drops are seen. Qualitatively, these fluctuations also mark the transition from a viscous-dominated flow to a more capillary-dominated invasion process. Similar behavior of the pressure drops has been reported in previous studies on drainage in both porous media (Ferrari and Lunati, 2013) and rough fractures (Chen , 2018; Yang , 2019).

FIG. 20.

Average pressure drop ΔPio between the inlet and outlet of the fracture, as a function of the invading fluid saturation, S1, for the sinusoidal (a) and geological (b) fractures and for different Ca values. The solid and dashed lines correspond to the RF and UF simulation results, respectively.

FIG. 20.

Average pressure drop ΔPio between the inlet and outlet of the fracture, as a function of the invading fluid saturation, S1, for the sinusoidal (a) and geological (b) fractures and for different Ca values. The solid and dashed lines correspond to the RF and UF simulation results, respectively.

Close modal

For the pressure drops, a generally good agreement is observed between those obtained from RF and UF simulations, particularly in the case of the sinusoidal fracture. This improved match is primarily attributed to a thinner film of defending fluid in the sinusoidal fracture geometry. At log(Ca)=3.0, when the film thickness is the largest, the pressure drops obtained in both fractures from the UF simulations are considerably lower than those obtained from the RF simulations. Another noteworthy observation is the larger ΔPio values in the sinusoidal fracture compared to the geological fracture. In addition to the mean aperture of the geological fracture being 1.3 times larger than that of the sinusoidal fracture, the larger quantity of wetting film in the latter case further reduces the effective aperture available to the invading fluid. Consequently, higher pressure drops are obtained in the sinusoidal fracture under the same inlet flow conditions.

5. Interfacial lengths and invading tip velocities

We conduct another comparison of the invasion morphologies by computing the length L of the fluid–fluid interface within the mean aperture plane (xy) of the fracture geometries. For this, the fluid distributions are mapped onto a constant-thickness z grid, and the interface length is obtained as the intersection length between the interface and the mid-cell plane of this grid. Starting with a flat interface (L(t=0)=Ly) and some initial saturation of the invading fluid S1(t=0), we are interested in the evolution of the interfacial length as a function of the saturation S1. Figure 21 depicts this evolution for the two fracture geometries across three representative Ca values spanning from highest to lowest. A relatively linear increase in L with saturation is observed across all cases. Higher Ca values correspond to a pattern close to viscous fingering, leading to a larger L. These trends in L are captured well by the RF simulations and are consistent with prior studies (Chen , 2018; Ferrari , 2015). Interface lengths predicted by the UF simulations match this behavior only at low capillary numbers when the influence of the wetting film's thickness diminishes. Notably, in the geological fracture at the highest Ca [ log(Ca)=3.0], a rapid growth in L is seen [Fig. 21(b)]. At large Ca, breakup events dominate the invasion morphology, leading to an increase in L (Pak , 2015; Chen , 2018). At low Ca, cutoff and trapping of the fluid cluster also impact the evolution of L. This is seen in the sinusoidal fracture at log(Ca)=5.0 [Fig. 21(a)], where trapping occurs early in the invasion process [see Fig. 14(e-e″)], temporarily increasing the interfacial length. For further details, see the  Appendix, which provides additional pictures of the flow patterns for each of the flow configurations and links to online videos of the corresponding simulations (Multimedia available online).

FIG. 21.

Interface lengths L for the RF (solid lines) and UF (dashed lines) simulations, plotted as a function of the invading fluid's saturation, S1, for the sinusoidal (a) and geological (b) fracture geometries, respectively, and for three different cases of Ca values.

FIG. 21.

Interface lengths L for the RF (solid lines) and UF (dashed lines) simulations, plotted as a function of the invading fluid's saturation, S1, for the sinusoidal (a) and geological (b) fracture geometries, respectively, and for three different cases of Ca values.

Close modal

Figure 22 shows the interface tip velocities Utip plotted as a function of the invading fluid saturation S1 for the two geometries. The tip velocities are the velocities of the most advanced interface point. With decreasing capillary numbers (Ca) and injection velocities (U), the average tip velocities also decrease. In the sinusoidal fracture, Utip exhibits a quasi-periodic behavior, while in the geological fracture, unsurprisingly, its spatial variations are irregular. Additionally, higher fluctuations in Utip can be noticed in the case of the sinusoidal fracture. At low Ca values [ log(Ca)=4.5,5.0], strong fluctuations indicative of increased capillary fingering appear in the tip velocities. In the non-smooth fracture, a clear crossover is observed between log(Ca)=4.0 and –4.5, signifying a comparable influence of capillary and viscous forces (Chen , 2017). This behavior is well captured by the UF simulations, although not quite as well as by the RF simulations.

FIG. 22.

Interface tip velocities Utip for the RF (solid lines) and UF (dashed lines) simulations, plotted as a function of the invading fluid 1 saturation S1 for the sinusoidal (a) and geological (b) fractures, respectively, and for five different Ca values.

FIG. 22.

Interface tip velocities Utip for the RF (solid lines) and UF (dashed lines) simulations, plotted as a function of the invading fluid 1 saturation S1 for the sinusoidal (a) and geological (b) fractures, respectively, and for five different Ca values.

Close modal

6. Longitudinal saturation profiles

The aforementioned parameters, such as interfacial lengths and average pressure drops, while revealing insightful flow behaviors, do not encompass the entire information regarding the overall fluid distributions. For instance, statistically different fracture geometries might exhibit contrasting invasion patterns yet exhibit the same temporal evolution of their interface lengths (Ferrari , 2015). To complement the analysis, we thus examine longitudinal saturation profiles of the invading fluid, obtained as averages of the indicator function across the transverse (y) direction. In Fig. 23, these profiles are presented as functions of the normalized longitudinal coordinate, x/Lx, for the two fracture geometries. It is noteworthy that the longitudinal saturation profile also shares a close connection with the Darcy scale depiction of flow through permeable media, which is based on the method of volume-averaging (see, e.g., Bear, 2013). Regarding the observed distinctions in saturation profiles between the RF and UF simulations, these dissimilarities can be attributed to one of the following factors: (i) the presence (or absence) of the defending fluid film, which could lead to a reduction (or increase) in the saturation of the invading fluid; (ii) the lateral extent occupied by the advancing fingers; or (iii) a mismatch between the resultant invasion patterns in the two scenarios, as observed in the sinusoidal fracture at low Ca. At high Ca [ log(Ca)=3.0], when the wetting film thickness is substantial, for the sinusoidal fracture case [Fig. 23(a)], the saturation profiles from the UF simulations exhibit higher saturation levels at all locations (x), but with larger differences at longitudinal positions where the average aperture is larger. Conversely, at the same Ca value for the geological fracture [Fig. 23(d)], the dissimilarities in the saturation profiles primarily arise due to the mismatch in the branched invasion patterns [Fig. 15(a″)] and are smaller. For both fracture geometries, at intermediate Ca values [ log(Ca)=4.0, Figs. 23(b) and 23(e)], when the influence of the wetting film is moderate and the observed alignment between the corresponding invasion patterns from the RF and UF simulations is most prominent [Figs. 14(c″) and 15(c″)], the saturation profiles also exhibit the highest level of agreement. At the lowest Ca value [ log(Ca)=5.0, Fig. 23(c)], the disparities in saturation profiles for the sinusoidal fracture scenario are attributed to the pronounced mismatch between the invasion morphologies resulting from the two film resolution configurations [refer to Fig. 14(e″)]. On the contrary, for the geological fracture, although at log(Ca)=5.0 the patterns exhibit considerable resemblance at large scales, these disagreements predominantly arise due to small-scale discrepancies, where the advancing fluid front becomes more compact and occupies greater aperture space in the transverse direction [see Fig. 15(e″)].

FIG. 23.

Average longitudinal saturation profile for the sinusoidal (a)–(c) and geological (d)–(f) fractures, respectively, for three different Ca. The solid and dashed lines correspond to the results from the RF and UF simulations, respectively.

FIG. 23.

Average longitudinal saturation profile for the sinusoidal (a)–(c) and geological (d)–(f) fractures, respectively, for three different Ca. The solid and dashed lines correspond to the results from the RF and UF simulations, respectively.

Close modal

7. A note on the computational requirements

We have investigated to which extent resolving the film left behind by the defending fluid on the fracture walls or not influences the immiscible fluid–fluid displacements in rough fracture geometries, and the impact of not resolving the film on the predicted flow behavior. The grid resolution employed in these simulations plays a significant role in capturing the intricate dynamics of fluid flow within the fractures. Thus, it is important to recognize the distinct computational demands imposed by each of the two approaches, namely, the RF and the UF. As stated in Secs. III B and IV C, the RF simulations demand a more refined grid structure compared to their UF counterparts. Specifically, the RF simulations involve nearly twice the number of cells in the vertical z direction. Moreover, the grid resolution is vertically graded, resulting in higher cell density near the top and bottom walls.

The computational time required for these simulations is proportional to the number of cells in the grid. Consequently, RF simulations demand significantly more computational time due to their higher cell count. Additionally, the computational effort also scales with the number of time steps and the time step size, which in turn is constrained by the CFL stability criterion. For such flow cases where surface tension forces are relevant and involve complex fluid interfaces, to maintain numerical stability and mitigate the occurrence of parasitic currents, a CFL value of 0.5 is typically imposed (Deshpande , 2012). The CFL criterion necessitates smaller time steps, which implies that as grid resolution increases, requiring smaller cell sizes, the time step size decreases accordingly. Thus, the increased cell count and the reduction in time step sizes compound the total computational effort required for the RF simulations. For instance, in the case of the sinusoidal fracture geometry at the lowest capillary number [ log(Ca)=5.0], the RF simulation employed a total cell count of 3.2×107, twice that of the UF simulation. Given a CFL criterion of 0.5, this led to time steps approximately an order of magnitude smaller, around 2×105 s, compared to those in the UF counterpart. While the UF simulation utilized 120 processors and demanded 927.62 CPU-hours, the RF simulation ran on 250 processors, requiring 14 461.04 CPU-hours, i.e., 15 times the CPU-hours needed for the UF simulation. This stark difference highlights the substantial increase in computational demand associated with resolving the defending fluid films. It underscores the trade-off between computational cost and film resolution accuracy, a crucial consideration when choosing the simulation strategy in rough fracture drainage studies.

A challenging yet essential consideration in hydrodynamic-scale modeling of drainage within open rough fractures is the accurate resolution of the film of defending fluid, which adheres to the fracture walls. This study employs a VOF-based CFD approach, using a custom-modified version of the open-source code OpenFOAM, to evaluate the impact of film resolution on hydrodynamic-scale drainage simulations.

We have presented a detailed validation of the numerical model utilizing the classical case of viscous instability within a conventional Hele–Shaw geometry. Depending on whether the computation mesh resolves the defending fluid film or not, the results from two sets of simulations; “resolved-film” (RF) or “unresolved-film” (UF), were compared with the experimental and analytical results of Saffman and Taylor (1958). This comparison revealed that the RF simulations demonstrated excellent agreement with experimental results across the entire range of capillary numbers (Ca), while UF simulations at higher Ca resulted in slower finger propagation velocities and increased finger widths. With smaller Ca values (Ca104) and smaller film thickness, UF results aligned more closely with the experimental and RF simulation findings.

The investigation was then extended to two numerically generated rough fracture geometries, where the first has a sinusoidally varying aperture and the other mimics a naturally occurring geological fracture with a stochastic self-affine roughness of the fracture walls and correlations between the walls' topographies at large scales. We conducted drainage simulations on these fractures, employing both RF and UF simulations and covering a broad range of Ca which spanned over three orders of magnitude (105Ca103). To assess the effect of film resolution, we performed an extensive comparison of hydrodynamic-scale and macroscopic quantities of interest between the UF and RF simulations. Consequently, our analysis leads us to the following findings:

  • In the sinusoidal fracture case, throughout all capillary numbers, the observed invasion patterns had distinct preferential flow paths along channels with overall larger apertures, leading to flow-aligned fingers. Unsurprisingly, the geological fracture, on the other hand, displayed more irregular invasion patterns, consistent with the stochastic geometry. In this case, the displacement patterns transition from more to less flow-aligned as the Ca decreases.

  • At higher capillary numbers (Ca>104) with a considerable film of the defending fluid present on the walls, overall invasion patterns are similar, but the UF simulations notably underestimate other hydrodynamic-scale predictions such as interfacial lengths and macroscopic pressure drops, but overpredict invading fluid saturations and the breakthrough times.

  • At very low capillary numbers (Ca=105), while most hydrodynamic-scale predictions from the UF simulations align well with those from the RF simulations, the detailed invasion patterns, especially at smaller scales, can vary.

  • Regarding macroscopic flow variables such as the average pressure drop and saturation, differences do exist between the RF and UF simulations, especially at higher Ca. However, not resolving the film only results in serious mispredictions at large Ca (Ca103). At low Ca (Ca<104), these differences can be considered negligible.

  • The more abrupt spatial variations in wall topographies in the geological fracture lead to a reduction in wetting film thickness as compared to the sinusoidal fracture, resulting in less pronounced film effects and thus a better match between the UF simulations' predictions and those from the RF simulations.

  • During the invasion process, the thin film on the fracture walls reduces the effective aperture perceived by the invading fluid. This film thickness is not uniform, even in the ideal parallel plate geometry. One potential approach to partially accounting for the film's influence without resolving it is to adjust macroscopic parameters such as the effective apertures available to the invading fluid. However, it is important to note that this cannot fully rectify all discrepancies. This limitation arises from the inherent variability in film thicknesses, which are not known a priori, and the broadening of the fingers in the absence of film resolution.

In summary, the choice of film resolution in modeling two-phase flow in rough fractures involves a trade-off between solution accuracy and computational resources. Our study demonstrates that accurately resolving the film can require up to tenfold more computational time compared to simulations where the wetting film is not resolved. Resolved-film simulations differ from unresolved-film simulations; however, these differences become smaller at small capillary numbers. In any case, taking these considerations into account, our study demonstrates that the employed VOF model is reliable and has the potential to serve as a valuable complement to experimental observations. Moreover, it enables exploration across a wide range of parameters and scenarios that may be difficult to emulate with experiments and field observations.

Prospects to this study are numerous and include the detailed study of the wetting film's heterogeneity in space and time and of its dependence on Ca, M, and the fracture geometry, as well as the characterization of flow regimes in the (Ca, M, Bo) parameter space (Bo being the Bond number).

See the supplementary material which presents a detailed description of the VOF model employed in this study. Readers interested in detail of the governing equations, discretization, and solution procedure implemented in OpenFOAM are encouraged to see the references included therein.

We gratefully acknowledge the financial support of the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Project No. NE 824/20-1 and the Agence Nationale de la Recherche (ANR, French National Research Agency) under Project No. ANR-20-CE92-0026 for the DFG-ANR project “2PhlowFrac.”

The authors have no conflicts to disclose.

R. Krishna: Data curation (equal); Formal analysis (equal); Investigation (lead); Methodology (equal); Software (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (supporting). Y. Méheust: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (supporting); Methodology (equal); Project administration (supporting); Resources (supporting); Software (supporting); Supervision (supporting); Validation (supporting); Visualization (supporting); Writing – original draft (supporting); Writing – review & editing (equal). I. Neuweiler: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (supporting); Methodology (equal); Project administration (lead); Resources (lead); Supervision (lead); Validation (supporting); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

The following figures show the snapshots of the fluid–fluid invasion patterns in the RF (invading fluid shown in blue) and UF (invading fluid shown in red) simulations. The light blue color indicates the defending fluid. The corresponding multimedia video files are available online.

  • The video linked to Fig. 24 (Multimedia available online) shows the momentary pinch-off and merging of the invading phase seen at high Ca [ log(Ca)=3.0] in the geological fracture when the film is resolved (RF simulation). These phenomena are not observed in the corresponding UF simulation, as shown in Fig. 25 (Multimedia available online).

  • The video linked to Fig. 26 (Multimedia available online) shows the progressive pinch-off (isolated blobs of invading fluid) of the invading phase in the RF simulation of the displacement in the sinusoidal fracture for log(Ca)=4.5. However, as shown in Fig. 27 (Multimedia available online), this phenomenon is completely absent in the UF configuration. A similar observation is made for the sinusoidal fracture at log(Ca)=5.0 where the RF simulation as shown in Fig. 28 (Multimedia available online), in addition to pinch-off, also leads to merging of two fingers of the invading phase. In the corresponding UF simulation shown in Fig. 29 (Multimedia available online), this pinch-off and the resulting isolation of the invading phase are present, but to a much lesser extent.

FIG. 24.

Snapshots at three different time-steps of the invasion pattern obtained from the RF simulation in the geological fracture at log(Ca)=3.0. Multimedia available online.

FIG. 24.

Snapshots at three different time-steps of the invasion pattern obtained from the RF simulation in the geological fracture at log(Ca)=3.0. Multimedia available online.

Close modal
FIG. 25.

Snapshots at three different time-steps of the invasion pattern obtained from the UF simulation in the geological fracture at log(Ca)=3.0. Multimedia available online.

FIG. 25.

Snapshots at three different time-steps of the invasion pattern obtained from the UF simulation in the geological fracture at log(Ca)=3.0. Multimedia available online.

Close modal
FIG. 26.

Snapshots at three different time-steps of the invasion pattern obtained from the RF simulation in the sinusoidal fracture at log(Ca)=4.5. Multimedia available online.

FIG. 26.

Snapshots at three different time-steps of the invasion pattern obtained from the RF simulation in the sinusoidal fracture at log(Ca)=4.5. Multimedia available online.

Close modal
FIG. 27.

Snapshots at three different time-steps of the invasion pattern obtained from the UF simulation in the sinusoidal fracture at log(Ca)=4.5. Multimedia available online.

FIG. 27.

Snapshots at three different time-steps of the invasion pattern obtained from the UF simulation in the sinusoidal fracture at log(Ca)=4.5. Multimedia available online.

Close modal
FIG. 28.

Snapshots at three different time-steps of the invasion pattern obtained from the RF simulation in the sinusoidal fracture at log(Ca)=5.0. Multimedia available online.

FIG. 28.

Snapshots at three different time-steps of the invasion pattern obtained from the RF simulation in the sinusoidal fracture at log(Ca)=5.0. Multimedia available online.

Close modal
FIG. 29.

Snapshots at three different time-steps of the invasion pattern obtained from the UF simulation in the sinusoidal fracture at log(Ca)=5.0. Multimedia available online.

FIG. 29.

Snapshots at three different time-steps of the invasion pattern obtained from the UF simulation in the sinusoidal fracture at log(Ca)=5.0. Multimedia available online.

Close modal
1.
Amundsen
,
H.
,
Wagner
,
G.
,
Oxaal
,
U.
,
Meakin
,
P.
,
Feder
,
J.
, and
Jøssang
,
T.
, “
Slow two-phase flow in artificial fractures: Experiments and simulations
,”
Water Resour. Res.
35
,
2619
2626
, https://doi.org/10.1029/1999WR900147 (
1999
).
2.
Auradou
,
H.
, “
Influence of wall roughness on the geometrical, mechanical and transport properties of single fractures
,”
J. Phys. D: Appl. Phys.
42
,
214015
(
2009
).
3.
Bear
,
J.
,
Dynamics of Fluids in Porous Media
(
Courier Corporation
,
2013
).
4.
Berberović
,
E.
,
van Hinsberg
,
N. P.
,
Jakirlić
,
S.
,
Roisman
,
I. V.
, and
Tropea
,
C.
, “
Drop impact onto a liquid layer of finite thickness: Dynamics of the cavity evolution
,”
Phys. Rev. E
79
,
036306
(
2009
).
5.
Bikkina
,
P.
,
Wan
,
J.
,
Kim
,
Y.
,
Kneafsey
,
T. J.
, and
Tokunaga
,
T. K.
, “
Influence of wettability and permeability heterogeneity on miscible CO2 flooding efficiency
,”
Fuel
166
,
219
226
(
2016
).
6.
Birkholzer
,
J.
and
Tsang
,
Y.
, “
Modeling the thermal-hydrologic processes in a large-scale underground heater test in partially saturated fractured tuff
,”
Water Resour. Res.
36
,
1431
1447
, https://doi.org/10.1029/2000WR900025 (
2000
).
7.
Blunt
,
M.
,
King
,
M. J.
, and
Scher
,
H.
, “
Simulation and theory of two-phase flow in porous media
,”
Phys. Rev. A
46
,
7680
(
1992
).
8.
Bouchaud
,
E.
, “
Scaling properties of cracks
,”
J. Phys.: Condens. Matter
9
,
4319
(
1997
).
9.
Brackbill
,
J. U.
,
Kothe
,
D. B.
, and
Zemach
,
C.
, “
A continuum method for modeling surface tension
,”
J. Comput. Phys.
100
,
335
354
(
1992
).
10.
Bretherton
,
F. P.
, “
The motion of long bubbles in tubes
,”
J. Fluid Mech.
10
,
166
188
(
1961
).
11.
Brown
,
S. R.
, “
Fluid flow through rock joints: The effect of surface roughness
,”
J. Geophys. Res.
92
,
1337
1347
, https://doi.org/10.1029/JB092iB02p01337 (
1987
).
12.
Chen
,
Y.-F.
,
Fang
,
S.
,
Wu
,
D.-S.
, and
Hu
,
R.
, “
Visualizing and quantifying the crossover from capillary fingering to viscous fingering in a rough fracture
,”
Water Resour. Res.
53
,
7756
7772
, https://doi.org/10.1002/2017WR021051 (
2017
).
13.
Chen
,
Y.-F.
,
Guo
,
N.
,
Wu
,
D.-S.
, and
Hu
,
R.
, “
Numerical investigation on immiscible displacement in 3D rough fracture: Comparison with experiments and the role of viscous and capillary forces
,”
Adv. Water Resour.
118
,
39
48
(
2018
).
14.
De Lozar
,
A.
,
Juel
,
A.
, and
Hazel
,
A. L.
, “
The steady propagation of an air finger into a rectangular tube
,”
J. Fluid Mech.
614
,
173
195
(
2008
).
15.
Deshpande
,
S. S.
,
Anumolu
,
L.
, and
Trujillo
,
M. F.
, “
Evaluating the performance of the two-phase flow solver interFoam
,”
Comput. Sci. Disc.
5
,
014016
(
2012
).
16.
Detwiler
,
R. L.
,
Rajaram
,
H.
, and
Glass
,
R. J.
, “
Interphase mass transfer in variable aperture fractures: Controlling parameters and proposed constitutive relationships
,”
Water Resour. Res.
45
,
W08436
, https://doi.org/10.1029/2008WR007009 (
2009
).
17.
Dewangan
,
M. K.
,
Ghosh
,
U.
,
Le Borgne
,
T.
, and
Méheust
,
Y.
, “
Coupled electrohydrodynamic transport in rough fractures: A generalized lubrication theory
,”
J. Fluid Mech.
942
,
A11
(
2022
).
18.
Dou
,
Z.
,
Zhou
,
Z.
, and
Sleep
,
B.
, “
Influence of wettability on interfacial area during immiscible liquid invasion into a 3D self-affine rough fracture: Lattice Boltzmann simulations
,”
Adv. Water Resour.
61
,
1
11
(
2013
).
19.
Dussan V
,
E. B.
,
Ramé
,
E.
, and
Garoff
,
S.
, “
On identifying the appropriate boundary conditions at a moving contact line: An experimental investigation
,”
J. Fluid Mech.
230
,
97
116
(
1991
).
20.
Fatt
,
I.
, “
The network model of porous media
,”
Trans. AIME
207
,
144
181
(
1956
).
21.
Ferer
,
M.
,
Crandall
,
D.
,
Ahmadi
,
G.
, and
Smith
,
D. H.
, “
Two-phase flow in a rough fracture: Experiment and modeling
,”
Phys. Rev. E
84
,
016316
(
2011
).
22.
Ferrari
,
A.
and
Lunati
,
I.
, “
Direct numerical simulations of interface dynamics to link capillary pressure and total surface energy
,”
Adv. Water Resour.
57
,
19
31
(
2013
).
23.
Ferrari
,
A.
,
Jimenez-Martinez
,
J.
,
Borgne
,
T. L.
,
Méheust
,
Y.
, and
Lunati
,
I.
, “
Challenges in modeling unstable two-phase flow experiments in porous micromodels
,”
Water Resour. Res.
51
,
1381
1400
, https://doi.org/10.1002/2014WR016384 (
2015
).
24.
Gerlach
,
D.
,
Tomar
,
G.
,
Biswas
,
G.
, and
Durst
,
F.
, “
Comparison of volume-of-fluid methods for surface tension-dominant two-phase flows
,”
Int. J. Heat Mass Transfer
49
,
740
754
(
2006
).
25.
Glass
,
R. J.
,
Nicholl
,
M. J.
, and
Yarrington
,
L.
, “
A modified invasion percolation model for low-capillary number immiscible displacements in horizontal rough-walled fractures: Influence of local in-plane curvature
,”
Water Resour. Res.
34
,
3215
3234
, https://doi.org/10.1029/98WR02224 (
1998
).
26.
Glass
,
R. J.
,
Rajaram
,
H.
, and
Detwiler
,
R. L.
, “
Immiscible displacements in rough-walled fractures: Competition between roughening by random aperture variations and smoothing by in-plane curvature
,”
Phys. Rev. E
68
,
061110
(
2003
).
27.
Glass
,
R. J.
,
Rajaram
,
H.
,
Nicholl
,
M. J.
, and
Detwiler
,
R. L.
, “
The interaction of two fluid phases in fractured media
,”
Curr. Opin. Colloid Interface Sci.
6
,
223
235
(
2001
).
28.
Gopala
,
V. R.
and
van Wachem
,
B. G.
, “
Volume of fluid methods for immiscible-fluid and free-surface flows
,”
Chem. Eng. J.
141
,
204
221
(
2008
).
29.
Guiltinan
,
E. J.
,
Santos
,
J. E.
,
Cardenas
,
M. B.
,
Espinoza
,
D. N.
, and
Kang
,
Q.
, “
Two-phase fluid flow properties of rough fractures with heterogeneous wettability: Analysis with lattice Boltzmann simulations
,”
Water Resour. Res.
57
,
e2020WR027943
, https://doi.org/10.1029/2020WR027943 (
2021
).
30.
Gupta
,
R.
,
Fletcher
,
D. F.
, and
Haynes
,
B. S.
, “
On the CFD modelling of Taylor flow in microchannels
,”
Chem. Eng. Sci.
64
,
2941
2950
(
2009
).
31.
Hirt
,
C. W.
and
Nichols
,
B. D.
, “
Volume of fluid (VOF) method for the dynamics of free boundaries
,”
J. Comput. Phys.
39
,
201
225
(
1981
).
32.
Hoang
,
D. A.
,
van Steijn
,
V.
,
Portela
,
L. M.
,
Kreutzer
,
M. T.
, and
Kleijn
,
C. R.
, “
Benchmark numerical simulations of segmented two-phase flows in microchannels using the volume of fluid method
,”
Comput. Fluids
86
,
28
36
(
2013
).
33.
Hodges
,
S.
,
Jensen
,
O.
, and
Rallison
,
J.
, “
The motion of a viscous drop through a cylindrical tube
,”
J. Fluid Mech.
501
,
279
301
(
2004
).
34.
Horgue
,
P.
,
Augier
,
F.
,
Duru
,
P.
,
Prat
,
M.
, and
Quintard
,
M.
, “
Experimental and numerical study of two-phase flows in arrays of cylinders
,”
Chem. Eng. Sci.
102
,
335
345
(
2013
).
35.
Horgue
,
P.
,
Augier
,
F.
,
Quintard
,
M.
, and
Prat
,
M.
, “
A suitable parametrization to simulate slug flows with the volume-of-fluid method
,”
C. R. Méc.
340
,
411
419
(
2012
).
36.
Huang
,
H.
,
Krafczyk
,
M.
, and
Lu
,
X.
, “
Forcing term in single-phase and Shan-Chen-type multiphase lattice Boltzmann models
,”
Phys. Rev. E
84
,
046710
(
2011
).
37.
Hughes
,
R. G.
and
Blunt
,
M. J.
, “
Network modeling of multiphase flow in fractures
,”
Adv. Water Resour.
24
,
409
421
(
2001
).
38.
Hui
,
M.-H.
and
Blunt
,
M. J.
, “
Effects of wettability on three-phase flow in porous media
,”
J. Phys. Chem. B
104
,
3833
3845
(
2000
).
39.
Issa
,
R. I.
, “
Solution of the implicitly discretised fluid flow equations by operator-splitting
,”
J. Comput. Phys.
62
,
40
65
(
1986
).
40.
Jensen
,
M. H.
,
Libchaber
,
A.
,
Pelcé
,
P.
, and
Zocchi
,
G.
, “
Effect of gravity on the Saffman-Taylor meniscus: Theory and experiment
,”
Phys. Rev. A
35
,
2221
(
1987
).
41.
Kim
,
J.-S.
,
Kwon
,
S.-K.
,
Sanchez
,
M.
, and
Cho
,
G.-C.
, “
Geological storage of high level nuclear waste
,”
KSCE J. Civ. Eng.
15
,
721
737
(
2011
).
42.
Lafaurie
,
B.
,
Nardone
,
C.
,
Scardovelli
,
R.
,
Zaleski
,
S.
, and
Zanetti
,
G.
, “
Modelling merging and fragmentation in multiphase flows with surfer
,”
J. Comput. Phys.
113
,
134
147
(
1994
).
43.
Lavrov
,
A.
, “
Radial flow of non-Newtonian power-law fluid in a rough-walled fracture: Effect of fluid rheology
,”
Transp. Porous Med.
105
,
559
570
(
2014
).
44.
Lee
,
H.-B.
,
Yeo
,
I. W.
,
Ji
,
S.-H.
, and
Lee
,
K.-K.
, “
Wettability-dependent DNAPL migration in a rough-walled fracture
,”
J. Contam. Hydrol.
113
,
44
55
(
2010
).
45.
Lenci
,
A.
,
Méheust
,
Y.
,
Putti
,
M.
, and
Federico
,
V. D.
, “
Shear-thinning hydraulic behavior of rough fractures: A stochastic analysis via Monte Carlo simulations
,”
Water Resour. Res
58
,
WR032024
(
2022a
).
46.
Lenci
,
A.
,
Putti
,
M.
,
Di Federico
,
V.
, and
Méheust
,
Y.
, “
A lubrication-based solver for shear-thinning flow in rough fractures
,”
Water Resour. Res.
58
,
WR031760
, https://doi.org/10.1029/2021WR031760 (
2022b
).
47.
Lenormand
,
R.
, “
Liquids in porous media
,”
J. Phys.: Condens. Matter
2
,
SA79
(
1990
).
48.
Lenormand
,
R.
and
Zarcone
,
C.
, “
Capillary fingering: Percolation and fractal dimension
,”
Transp. Porous Med.
4
,
599
612
(
1989
).
49.
Måløy
,
K. J.
,
Feder
,
J.
, and
Jøssang
,
T.
, “
Viscous fingering fractals in porous media
,”
Phys. Rev. Lett.
55
,
2688
2691
(
1985
).
50.
Meakin
,
P.
and
Tartakovsky
,
A. M.
, “
Modeling and simulation of pore-scale multiphase fluid flow and reactive transport in fractured and porous media
,”
Rev. Geophys.
47
,
RG3002
, https://doi.org/10.1029/2008RG000263 (
2009
).
51.
Méheust
,
Y.
and
Schmittbuhl
,
J.
, “
Geometrical heterogeneities and permeability anisotropy of rough fractures
,”
J. Geophys. Res.
106
,
2089
2102
, https://doi.org/10.1029/2000JB900306 (
2001
).
52.
Méheust
,
Y.
and
Schmittbuhl
,
J.
, “
Scale effects related to flow in rough fractures
,”
Pure Appl. Geophys.
160
,
1023
1050
(
2003
).
53.
Muller
,
R. A.
,
Finsterle
,
S.
,
Grimsich
,
J.
,
Baltzer
,
R.
,
Muller
,
E. A.
,
Rector
,
J. W.
,
Payer
,
J.
, and
Apps
,
J.
, “
Disposal of high-level nuclear waste in deep horizontal drillholes
,”
Energies
12
,
2052
(
2019
).
54.
Neuweiler
,
I.
,
Sorensen
,
I.
, and
Kinzelbach
,
W.
, “
Experimental and theoretical investigations of drainage in horizontal rough-walled fractures with different correlation structures
,”
Adv. Water Resour.
27
,
1217
1231
(
2004
).
55.
OpenFOAM
,
The Open Source CFD Toolkit User Guide
(
OpenFOAM Foundation
,
Dordrecht
,
2012
).
56.
Pak
,
T.
,
Butler
,
I. B.
,
Geiger
,
S.
,
Van Dijke
,
M. I.
, and
Sorbie
,
K. S.
, “
Droplet fragmentation: 3d imaging of a previously unidentified pore-scale process during multiphase flow in porous media
,”
Proc. Natl. Acad. Sci. U. S. A.
112
,
1947
1952
(
2015
).
57.
Pasandideh-Fard
,
M.
,
Qiao
,
Y.
,
Chandra
,
S.
, and
Mostaghimi
,
J.
, “
Capillary effects during droplet impact on a solid surface
,”
Phys. Fluids
8
,
650
659
(
1996
).
58.
Plouraboué
,
F.
,
Kurowski
,
P.
,
Hulin
,
J.-P.
,
Roux
,
S.
, and
Schmittbuhl
,
J.
, “
Aperture of rough cracks
,”
Phys. Rev. E
51
,
1675
(
1995
).
59.
Porter
,
M. L.
,
Coon
,
E.
,
Kang
,
Q.
,
Moulton
,
J.
, and
Carey
,
J.
, “
Multicomponent interparticle-potential lattice Boltzmann model for fluids with large viscosity ratios
,”
Phys. Rev. E
86
,
036701
(
2012
).
60.
Qiu
,
Y.
,
Xu
,
K.
,
Pahlavan
,
A. A.
, and
Juanes
,
R.
, “
Wetting transition and fluid trapping in a microfluidic fracture
,”
Proc. Natl. Acad. Sci. U. S. A.
120
,
e2303515120
(
2023
).
61.
Reinelt
,
D.
, “
The effect of thin film variations and transverse curvature on the shape of fingers in a Hele–Shaw cell
,”
Phys. Fluids
30
,
2617
2623
(
1987
).
62.
Renardy
,
M.
,
Renardy
,
Y.
, and
Li
,
J.
, “
Numerical simulation of moving contact line problems using a volume-of-fluid method
,”
J. Comput. Phys.
171
,
243
263
(
2001
).
63.
Rosengarten
,
G.
,
Harvie
,
D.
, and
Cooper-White
,
J.
, “
Contact angle effects on microdroplet deformation using CFD
,”
Appl. Math. Modell.
30
,
1033
1042
(
2006
).
64.
Rusche
,
H.
, “
Computational fluid dynamics of dispersed two-phase flows at high phase fractions
,” Ph.D. thesis (
University of London
,
2003
).
65.
Saffman
,
P. G.
and
Taylor
,
G. I.
, “
The penetration of a fluid into a porous medium or Hele-Shaw cell containing a more viscous liquid
,”
Proc. R. Soc. London Ser. A. Math. Phys. Sci.
245
,
312
329
(
1958
).
66.
Santiago
,
C.
,
Ghomeshi
,
S.
,
Kryuchkov
,
S.
, and
Kantzas
,
A.
, “
Pore level modeling of imbibition in heavy oil saturated media
,”
J. Pet. Sci. Eng.
140
,
108
118
(
2016
).
67.
Schmittbuhl
,
J.
,
Schmitt
,
F.
, and
Scholz
,
C.
, “
Scaling invariance of crack surfaces
,”
J. Geophys. Res.
100
,
5953
5973
, https://doi.org/10.1029/94JB02885 (
1995
).
68.
Sethian
,
J. A.
,
Level Set Methods and Fast Marching Methods: Evolving Interfaces in Computational Geometry, Fluid Mechanics, Computer Vision, and Materials Science
(
Cambridge University Press
,
1996
).
69.
Shi
,
Z.
,
Zhang
,
Y.
,
Liu
,
M.
,
Hanaor
,
D. A.
, and
Gan
,
Y.
, “
Dynamic contact angle hysteresis in liquid bridges
,”
Colloids Surf. A: Physicochem. Eng. Aspects
555
,
365
371
(
2018
).
70.
Sun
,
Y.
and
Beckermann
,
C.
, “
Sharp interface tracking using the phase-field equation
,”
J. Comput. Phys.
220
,
626
653
(
2007
).
71.
Tabeling
,
P.
and
Libchaber
,
A.
, “
Film draining and the Saffman-Taylor problem
,”
Phys. Rev. A
33
,
794
(
1986
).
72.
Tabeling
,
P.
,
Zocchi
,
G.
, and
Libchaber
,
A.
, “
An experimental study of the Saffman-Taylor instability
,”
J. Fluid Mech.
177
,
67
82
(
1987
).
73.
Tanveer
,
S.
, “
Surprises in viscous fingering
,”
J. Fluid Mech.
409
,
273
308
(
2000
).
74.
Tartakovsky
,
A. M.
and
Meakin
,
P.
, “
Simulation of unsaturated flow in complex fractures using smoothed particle hydrodynamics
,”
Vadose Zone J.
4
,
848
855
(
2005
).
75.
Williams
,
M.
,
Kothe
,
D.
, and
Puckett
,
E.
, Accuracy and Convergence of Continuum Surface Tension Models (Cambridge University Press,
1998
), pp.
294
305
.
76.
Xue
,
Y.
,
Dang
,
F.
,
Shi
,
F.
,
Li
,
R.
, and
Cao
,
Z.
, “
Evaluation of gas migration and rock damage characteristics for underground nuclear waste storage based on a coupled model
,”
Sci. Technol. Nucl. Install.
2018
,
2973279
.
77.
Yang
,
Z.
,
Méheust
,
Y.
,
Neuweiler
,
I.
,
Hu
,
R.
,
Niemi
,
A.
, and
Chen
,
Y.-F.
, “
Modeling immiscible two-phase flow in rough fractures from capillary to viscous fingering
,”
Water Resour. Res.
55
,
2033
2056
, https://doi.org/10.1029/2018WR024045 (
2019
).
78.
Yang
,
Z.
,
Neuweiler
,
I.
,
Méheust
,
Y.
,
Fagerlund
,
F.
, and
Niemi
,
A.
, “
Fluid trapping during capillary displacement in fractures
,”
Adv. Water Resour.
95
,
264
275
(
2016
).
79.
Yang
,
Z.
,
Niemi
,
A.
,
Fagerlund
,
F.
, and
Illangasekare
,
T.
, “
A generalized approach for estimation of in-plane curvature in invasion percolation models for drainage in fractures
,”
Water Resour. Res.
48
,
W09507
, https://doi.org/10.1029/2012WR011829 (
2012
).
80.
Zhang
,
L.
,
Yang
,
Z.
,
Méheust
,
Y.
,
Neuweiler
,
I.
,
Hu
,
R.
, and
Chen
,
Y.-F.
, “
Displacement patterns of a Newtonian fluid by a shear-thinning fluid in a rough fracture
,”
Water Resour. Res.
59
,
WR034958
, https://doi.org/10.1029/2023WR034958 (
2023
).
81.
Zhang
,
T.
,
Li
,
Z.
,
Adenutsi
,
C. D.
, and
Lai
,
F.
, “
A new model for calculating permeability of natural fractures in dual-porosity reservoir
,”
Adv. Geo-Energy Res.
1
,
86
92
(
2017
).
82.
Zhao
,
B.
,
MacMinn
,
C. W.
,
Primkulov
,
B. K.
,
Chen
,
Y.
,
Valocchi
,
A. J.
,
Zhao
,
J.
,
Kang
,
Q.
,
Bruning
,
K.
,
McClure
,
J. E.
,
Miller
,
C. T
et al, “
Comprehensive comparison of pore-scale models for multiphase flow in porous media
,”
Proc. Natl. Acad. Sci. U. S. A.
116
,
13799
13806
(
2019
).
83.
Zimmerman
,
R. W.
and
Bodvarsson
,
G. S.
, “
Hydraulic conductivity of rock fractures
,”
Transp. Porous Med.
23
,
1
30
(
1996
).