This study aims to explore the influence of various geometrical and hydraulic parameters on flow behavior and hydraulic conductivity in a single artificial fracture through a series of laboratory experiments. Laboratory experiments were conducted to examine unconfined groundwater flow through an artificially constructed single fracture. The fracture model consisted of varying aperture sizes (3, 9, and 12 mm) and different surface roughness conditions (fine, medium, and coarse sand coatings). Non-Darcian turbulent flow characteristics were observed at different flow rates, and the gradient of Reynolds number versus average flow velocity increased with aperture size. Flow parameters of the Darcian, Izbash, and Forchheimer models were calculated to characterize the flow behavior. Both the Forchheimer and Izbash models were found suitable for describing the non-Darcian flow characteristics under the prevailing conditions. The study revealed that hydraulic conductivity depended on flow length for fractures with different apertures and surface roughnesses, likely due to the presence of 2-D torturous flow within the rough fracture surface. These findings contribute to a better understanding of groundwater flow in fractured rock aquifers and provide valuable insights for modeling and managing such systems.

  • Flow in a single fracture for different surface roughness and aperture.

  • Analysis of Darcy, Forchheimer, and Izbash model parameters.

  • Analysis of variation in the gradient of Reynold's number with average flow velocity.

  • Evaluation of local cubic law and evaluation of spatial dependence of hydraulic conductivity.

Symbol Description

c

Aperture size

2B

Fracture spacing

μ

Dynamic viscosity of the fluid

ρ

Fluid density

g

Acceleration due to gravity

Q

Flow rate

J

Hydraulic gradient

Re

Reynolds number

ν

Kinematic viscosity

V

Average flow velocity

K

Hydraulic conductivity

A, B

Forchheimer model coefficients (represent inertial and viscous forces)

a, b

Izbash model coefficients

h1, h2

Water levels in recharge and discharge flumes

l

The effective length of the flume

q

Unit width flux

Qlcl

Discharge from local cubic law

Qexp

Experimental discharge

H1, H2

Piezometric heads

L

Flow length between piezometers

∇P

Pressure gradients along the flow direction

β

Forchheimer coefficient (depends on the medium geometry)

F0

Forchheimer number

Fractured rock aquifers play a crucial role in various fields, such as water resources management, disposal of radioactive waste, oil and gas production, geothermal utilization, and water production (Cook 2003). These aquifers consist of a network of fractures that traverse the rock matrix, serving as a storage medium for water due to their relatively high permeability compared to the surrounding impervious blocks. However, their permeability poses a significant challenge, as pollutants can readily travel through the fracture network and contaminate adjacent aquifers. Understanding the subsurface flow through fractured rocks is essential for effectively managing water resources, preventing contamination, and optimizing various industrial processes.

In the study of fractured aquifers, fluid flow behavior is strongly influenced by fracture characteristics such as tortuosity and connectivity. The tortuosity of the fracture path, in particular, plays a significant role in determining the flow rate, especially when the aperture size is small (Tsang 1984). To accurately simulate solute transport in fractures, it is necessary to consider the variation of hydraulic conductivity (K) with scale (Carrera 1993; Neuman 1994). While K is independent of the measurement scale for a homogeneous medium, it becomes a function of heterogeneity and fluid flow for a heterogeneous medium (Schulze-Makuch et al. 1999). Researchers have observed a correlation between hydraulic conductivity (K) values and the measurement scale in different granitic rocks, indicating that K increases until a particular rock volume is reached, beyond which it remains constant (Clauser 1992). Several methods, such as slug tests, flow meter tests, permeameter tests, pumping tests, grain size analyses, and geophysical tests, can determine the hydraulic conductivity of saturated groundwater flow (Neuman & Di Federico 2003). Anand et al. (2021) investigated several field and laboratory studies on the mobility of cohesionless sediments through open channels.

Studies have shown that the aperture of natural fractures can be larger than the hydraulic aperture. The distinction between hydraulic and mechanical aperture is essential, with the latter often being larger due to surface roughness and tortuosity (Hakami & Larsson 1996; Chen et al. 2000). Cornet et al. (2003) concluded that for a mechanical aperture greater than 15 μm, both mechanical and hydraulic aperture become equivalent for flow calculations in a single fracture, and the surface roughness no longer alters the flow. Fractured aquifers also exhibit faster flow rates than porous media and non-Darcian flow has been observed in various experiments and numerical models (Choi et al. 1997).

The nonlinear flow behavior is encountered for groundwater flow through porous media when the Reynolds number exceeds 10 (Oron & Berkowitz 1998). Rough-surfaced fractures can further complicate the flow, as permeability can vary even over small distances in hard-rock terrains (Baker 2000; Berkowitz 2002). It is recommended to consider the average value of hydraulic parameters to account for this variability. Bear & Corapcioglu (2012) characterized a fracture as a set of parallel plates with a well-defined aperture for modeling flow within the fracture. Local cubic law (LCL) was applied to calculate the flow rates. However, LCL for fracture flow has been inadequate in many case studies, as flow in fractured rocks often exhibits non-Darcian behavior (Field & Nash 1997; Kohl et al. 1997). LCL, Stokes, and Navier–Stokes equations have been further explored to model flow in rough-walled fractures for different aperture sizes (Brush & Thomson 2003).

Experimental studies have been conducted to investigate flow characteristics through natural fractures and laboratory-prepared fracture samples. These studies have focussed on characterizing nonlinear flow behavior, evaluating the applicability of flow equations such as Izbash and Forchheimer's equations, and examining the influence of hydraulic gradient, scale effect, and surface roughness (Qian et al. 2005, 2007, 2011a, 2011b; Zhang & Nemcik 2013; Yan et al. 2018). Chen et al. (2015) conducted laboratory experiments on rough-walled granite fracture samples and studied the variation of Forchheimer's coefficient with confining stress. Outflow experiments on smooth and artificial rough fractures were conducted, where both Izbash and Forchheimer's equations were found capable of describing the transition flow conditions (Tzelepis et al. 2015). Li et al. (2016) studied the influence of hydraulic gradient, scale effect, and surface roughness on single-fracture intersections having nonlinear flow. The determination of geometric parameters and the development of empirical models have also been explored to understand better and predict fluid flow in a single fracture (Rong et al. 2020; Xing et al. 2021; Chen et al. 2022). Ikhsan et al. (2022) analyzed the impact of grain size uniformity and shear stress on riverbed stability for steady and uniform flow, aiming to ascertain the minimum thickness required for a non-cohesive Armor layer. Qasim et al. (2022) further examined how the discordance of bed flumes can influence the flow characteristics of hydraulic structures. Energy dissipation by porous structure was studied, which in turn can help in mitigating scour around bridge abutments (Widyastuti et al. 2022)

The present study mainly focuses on the experimental investigation of groundwater flow in an artificially built single fracture for different surface roughness (fine, medium, and coarse), fracture apertures, and hydraulic gradients. To avoid the geometrical constraints and uncertainty in data interpretation and to precisely regulate the experimental parameters (aperture, surface roughness), an artificially built fracture plane with sand glued on one surface to simulate fracture roughness has been used. A few researchers, such as Qian et al. (2005, 2011a, 2011b), Chen et al. (2009), and Tzelepis et al. (2015), have also done a similar study using a single artificial fracture.

However, a distinct departure from previous research is observed in several aspects. Firstly, the fractures used in the present study are significantly larger, approximately twice the size of those examined in previous studies (Qian et al. 2005, 2007, 2011a, 2011b; Tzelepis et al. 2015; Yan et al. 2018). Secondly, the fracture surfaces are coated with sand of different particle sizes, conforming to the Indian Standards classification of fine, medium, and coarse roughness, presenting a novel approach to surface characterization. Additionally, the present study incorporates a diverse range of aperture sizes (3, 9, and 12 mm), in contrast to the limited aperture configurations examined in earlier research. Notably, plywood sheets were utilized to simulate the fracture planes, representing a departure from the plexiglass sheets employed in previous investigations. Furthermore, a comprehensive network of piezometers was installed along the fracture length to facilitate accurate and detailed monitoring of the flow behavior, a significant improvement over the limited number of piezometers typically used in earlier studies.

The present study emphasizes the experimental investigation of flow through a single fracture, which is essentially non-Darcian. The gradient of Reynolds number with different average flow velocities for different surface roughness and aperture size is studied. The parameters for different flow-characterizing models have been calculated, finding the best-suited model for the prevailing conditions. Further, hydraulic conductivity variation with flow length is studied for different apertures, surface roughness, and hydraulic gradient. Also, the variation of the Forchheimer coefficient with hydraulic gradient is discussed.

Through these novel experimental setups and measurements, we gained valuable insights into the influence of fracture size, aperture, and surface roughness on unconfined groundwater flow characteristics. The findings contribute to a better understanding of flow behavior in large-scale fractured rock aquifers and offer important implications for groundwater management and modeling in similar geological settings.

Parallel plate models

In this approach, a parallel, evenly spaced, planar fracture system is assumed in an impermeable rock matrix. Hydraulic conductivity of the medium, parallel to the direction of fracture, is expressed as (Lamb 1932; Polubarinova-Kochina 1962):
(1)
Here, 2B [L] is fracture spacing, c [L] is aperture,[ML−1T−1] is the dynamic viscosity of the fluid, ρ [ML−3] is the density of the fluid, and g [L2T−1] is the acceleration due to gravity. Due to the dependency of hydraulic conductivity on fracture aperture, it is also referred to as the LCL. Hydraulic conductivity is zero in all other directions. Discharge in fracture per unit width for a planar, uniform aperture can be further expressed (Lamb 1932; Polubarinova-Kochina 1962):
(2)

In this equation, Q [L3T−1] represents discharge, and J (dimensionless) is the hydraulic gradient in the fracture plane. This approach is valid for laminar flow only. Zimmerman & Bodvarsson (1996) suggested that the aperture should be replaced by the hydraulic aperture in the case of rough-wall fracture for the applicability of cubic law.

Reynolds number

Reynolds number for unconfined flow in a single fracture can be expressed as (Qian et al. 2007):
(3)
Here, R [L] is the hydraulic radius for flow through the cross-section , h is the wetted height in fracture, [L2T−1] is the kinematic viscosity, and V [LT−1] is the average velocity of flow. Since c<<h, therefore Equation (3) reduces to:
(4)

For calculating the Reynolds number for different flow rates through the fracture, the above equation is used. It has been observed that at Re> 10, flow no longer remains laminar (Zimmerman & Yeo 2000). LCL becomes invalid, and Darcy's law becomes inapplicable in such flow conditions.

Darcy's law (1856)

For describing linear flow through a single fracture, Darcy's law can be expressed as
(5)
where V [L/T] is the average flow velocity through the fracture, K [L/T] is the hydraulic conductivity, and J is the hydraulic gradient. For large specific discharge values, the flow becomes non-Darcian.

Forchheimer model (1901)

This model characterizes the nonlinear flow conditions at a macroscopic scale through a quadratic equation with zero intercepts (Forchheimer 1901).
(6)
Here, J is the hydraulic gradient, and A [T2L−2] and B [TL−1] are the model coefficients (pseudo conductivity parameters) representing nonlinear and linear effects. The first and second terms represent the effect of inertial and viscous force, respectively. In case the viscous force is much smaller than the inertial force, the second term of the equation can be dropped out, and the formula becomes:

Izbash model (1931)

This model also describes nonlinear flow equations for fully developed turbulent flow. This model is generally used to describe flow processes in porous media.
(7)

In this equation, a [TnLn] is a model parameter, and b (dimensionless) is the power index whose value varies between 1 and 2. After plotting V and J on a log-log scale, if the trend comes to be linear, then the slope corresponds to the power index. However, if a nonlinear trend is observed, the Forchheimer model should be applied. Compared to the Forchheimer model, this model offers an advantage as it involves a single pseudo hydraulic conductivity parameter instead of two, as in the Forchheimer equation.

The experimental setup was fabricated in the Hydraulics Laboratory at the Indian Institute of Technology, Roorkee, India. Figure 1 shows the setup layout, which comprised a single fracture formed by two parallel plywood sheets 10.0 m long, 0.02 m wide, and 0.3 m high. On the left side, a recharge flume with dimensions of 1.18 m length, 0.3 m width, and 0.59 m height was built, and a discharge flume of similar size was constructed on the right side of the fracture. To ensure a continuous water supply for experiments, a water tank with a capacity of 500 L was connected to the recharge flume.
Figure 1

Experimental setup of the single plate fracture: (a) schematic diagram, (b) aperture, and (c) fracture plate with coating.

Figure 1

Experimental setup of the single plate fracture: (a) schematic diagram, (b) aperture, and (c) fracture plate with coating.

Close modal
The study utilized artificially created single fractures by combining cement and sand in a 1:2.5 weight ratio. This ratio strikes a balance between strength and workability, typical for general applications like bricklaying and plastering. Sand adds volume and enhances workability, while cement serves as the binder. The 1:2.5 ratio optimizes cost-effectiveness since cement is more expensive than sand while maintaining sufficient strength and workability for the rough fracture surface. The sand diameters fell into three categories based on the Indian Standards classification: fine (0.075–0.425 mm), medium (0.425–2 mm), and coarse (2–4.75 mm) (IS: 1498–1970, 2006). The soil samples representing these categories are depicted in Figure 2.
Figure 2

Sand samples used to induce the surface roughness in fracture, i.e., coarse sand, medium sand, and fine sand.

Figure 2

Sand samples used to induce the surface roughness in fracture, i.e., coarse sand, medium sand, and fine sand.

Close modal

The aperture of the fracture, which refers to the distance between the two plywood sheets, was adjusted to different sizes (3, 9, and 12 mm) for conducting the experiments. It was intended to keep the aperture size as minimal as possible while fabricating the fracture in the laboratory. Since plywood sheets were utilized to prepare the fracture, it was quite challenging to work on such small aperture sizes. The initial gap between fracture sheets was 12 mm. Thereafter, to create a 9 mm gap for creating an aperture size of 9 mm, the plywood sheets of 3 mm thickness were placed and stuck on one side of the fracture sheet. Similarly, to create a 3 mm aperture size, a 9 mm thick plywood sheet was inserted and stuck on one side of the fracture. The authors tried to create the aperture sizes in multiple of 3. However, due to the unavailability of a 6 mm thick plywood sheet, the experimental study was not done on a 6 mm aperture size. Therefore, the study focussed on 3, 9, and 12 mm aperture sizes.

The flow rate through the fracture could be altered by adjusting the hydraulic head in the recharge and discharge flumes. Nine piezometers were installed inside the fracture plane at equal distances to determine the hydraulic head. The calculations were made once the steady-state flow condition was obtained. To ensure an accurate representation of flow behavior within the fracture for turbulent flow through the media, the first piezometer was strategically positioned at a distance of 52 cm from the source. This location was chosen to minimize the impact of the boundary layer, keeping it at a distance greater than the entrance length (>10 c). As a result, the effective length under consideration extends for a total of 948 cm. The average flow velocity (V) determined the flow rate to the discharge flume.

Where h1 and h2 are the water levels at recharge and discharge flumes, l is the effective length (i.e., 9.48 m) between them. The unit width flux, denoted as q [L2T−1], is expressed as:
(9)

Here, Q represents the total discharge rate through the fracture, and (h1+h2)/2 indicates the average saturated width of the fracture (Qian et al. 2011a, 2011b).

Furthermore, the specific discharge is given by the equation:
(10)

In this equation, V represents the average flow velocity, q is the unit width flux, and c denotes the average aperture. For the experimental runs conducted at an average temperature of 20 °C, the kinematic flow viscosity is assumed to be 1.005 × 10−6m2/s (White 1990).

Variation of Reynold's number with the average flow velocity

Multiple experimental runs were conducted for different flow rates in a single fracture with aperture sizes (denoted by ‘ c ’) of 3, 9, and 12 mm, using different surface roughness conditions created by a mixture of cement and sand of different diameters (fine, medium, and coarse). The flow rates varied from 4.75 to 100.38 cm3/s, and the corresponding average flow velocities were computed for each aperture size and wall roughness. The variations of Reynolds numbers were calculated using Equation (4), and the results are presented in Figure 3. The computed slopes of the Reynolds number against average flow velocity for different surface roughness and aperture sizes are provided in Table 1.
Table 1

Fracture aperture with Reynolds number versus average flow velocity

Fracture aperture (mm)The gradient of Re v/s V
Fine-wall fractureMedium-wall fractureCoarse-wall fracture
3.0 1,492.554 1,492.539 1,492.564 
9.0 4,477.546 4,477.65 4,477.597 
12.0 5,969.917 5,970.279 5,969.814 
Fracture aperture (mm)The gradient of Re v/s V
Fine-wall fractureMedium-wall fractureCoarse-wall fracture
3.0 1,492.554 1,492.539 1,492.564 
9.0 4,477.546 4,477.65 4,477.597 
12.0 5,969.917 5,970.279 5,969.814 
Figure 3

Plot between Reynolds number and average flow velocity with different aperture sizes, c = 3, 9, and 12 mm for fine, medium, and coarse-walled fracture surfaces.

Figure 3

Plot between Reynolds number and average flow velocity with different aperture sizes, c = 3, 9, and 12 mm for fine, medium, and coarse-walled fracture surfaces.

Close modal

For all the cases, the Reynolds number fell within the range of 12.34–474.22, suggesting the flow to be non-Darcian. The gradient of Reynolds number with flow velocity increased with larger aperture sizes, regardless of the fracture roughness.

Evaluation of the LCL

LCL serves as the base to study the disparity between real and ideal flow conditions. Figure 4 depicts the comparison between the experimental flow rates and the flow rates obtained from LCL for fractures with fine, medium, and coarse surface roughness. The purpose was to check the influence of inertial terms. The ratio of Qlcl to Qexp was found to increase with aperture size ‘c.’ Moreover, the values of Qlcl/Qexp were significantly greater than 1 for c > 3 mm, suggesting the substantial influence of inertial terms. However, an exception was observed for c = 12 mm in coarse roughness fracture, where a decrease in Qlcl to Qexp value was noted, suggesting that the effect of roughness has surpassed a threshold value. These findings align with the observations by Brush & Thomson (2003), who stated that the ratio of flow rate calculated using the Stokes equation to the accurate flow rate attains its peak for greater roughness values.
Figure 4

Evaluation of the LCL.

Figure 4

Evaluation of the LCL.

Close modal

Estimation of model parameters

The hydraulic tests conducted on fractures with fine, medium, and coarse surface roughness under different confining stresses provided experimental data are shown in Figures 57.
Figure 5

Plots between hydraulic gradient and average flow velocity for fine-walled fracture.

Figure 5

Plots between hydraulic gradient and average flow velocity for fine-walled fracture.

Close modal
Figure 6

Plots between hydraulic gradient and average flow velocity for medium-walled fracture.

Figure 6

Plots between hydraulic gradient and average flow velocity for medium-walled fracture.

Close modal
Figure 7

Plots between hydraulic gradient and average flow velocity for coarse-walled fracture.

Figure 7

Plots between hydraulic gradient and average flow velocity for coarse-walled fracture.

Close modal

The obtained experimental data were compared with three popular flow models, i.e., Darcy, Izbash, and Forchheimer, to identify the most suitable model for describing flow conditions in the single fracture model. The model parameters were computed using the SPSS statistical tool and are presented in Tables 24. These parameters were then used to fit the corresponding model equations, and the agreement between experimental and model data was assessed, as depicted in Figures 6 and 7.

Table 2

Estimation of model parameters for fine surface fracture

Darcy model
Forchheimer model
Izbash model
Fracture aperture (mm)
k = (1/K)R2ABR2abR2
0.07 0.84. 0.273 0.062 0.845 0.081 1.04 0.841 3.0 
0.091 0.808 1.006 0.06 0.859 0.261 1.303 0.846 9.0 
0.079 0.71 2.017 0.047 0.758 0.362 1.367 0.749 12.0 
Darcy model
Forchheimer model
Izbash model
Fracture aperture (mm)
k = (1/K)R2ABR2abR2
0.07 0.84. 0.273 0.062 0.845 0.081 1.04 0.841 3.0 
0.091 0.808 1.006 0.06 0.859 0.261 1.303 0.846 9.0 
0.079 0.71 2.017 0.047 0.758 0.362 1.367 0.749 12.0 
Table 3

Estimation of model parameters for medium surface fracture

Darcy model
Forchheimer model
Izbash model
Fracture aperture (mm)
k = (1/K)R2ABR2abR2
0.079 0.748 0.202 0.039 0.8 0.163 1.448 0.785 3.0 
0.126 0.886 0.894 0.054 0.97 0.483 1.531 0.968 9.0 
0.108 0.815 0.031 0.956 0.956 0.722 1.701 0.815 12.0 
Darcy model
Forchheimer model
Izbash model
Fracture aperture (mm)
k = (1/K)R2ABR2abR2
0.079 0.748 0.202 0.039 0.8 0.163 1.448 0.785 3.0 
0.126 0.886 0.894 0.054 0.97 0.483 1.531 0.968 9.0 
0.108 0.815 0.031 0.956 0.956 0.722 1.701 0.815 12.0 
Table 4

Estimation of model parameters for coarse surface fracture

Darcy model
Forchheimer model
Izbash model
Fracture aperture (mm)
k = (1/K)R2ABR2abR2
0.062 0.974 0.075 0.06 0.975 0.07 1.034 0.975 
0.066 0.741 0.508 0.032 0.821 0.243 1.48 0.81 
0.076 0.818 1.37 0.028 0.938 0.589 1.607 0.931 12 
Darcy model
Forchheimer model
Izbash model
Fracture aperture (mm)
k = (1/K)R2ABR2abR2
0.062 0.974 0.075 0.06 0.975 0.07 1.034 0.975 
0.066 0.741 0.508 0.032 0.821 0.243 1.48 0.81 
0.076 0.818 1.37 0.028 0.938 0.589 1.607 0.931 12 

Regression analysis revealed that non-laminar flow conditions prevailed throughout the experiments, rendering the Darcy equation unsuitable. However, for a 3 mm aperture, the Darcy equation provided a good fit, suggesting creeping flow conditions at smaller aperture values. The coefficients of determination for the Forchheimer and Izbash models were comparable, but the Forchheimer model exhibited greater precision. Due to the complexity involved in calculations with the Forchheimer equation and considering turbulent flow, the Izbash model was chosen for subsequent hydraulic conductivity calculations.

Hydraulic conductivity dependence on the scale for different apertures and surface roughness

The surface roughness introduces heterogeneity and tortuosity to the flow within the fracture, leading to non-uniform hydraulic conductivity (K) that becomes scale-dependent. Considering fully developed turbulent flow where inertial forces are more prominent, the Izbash equation can be used to determine the value of K in a single fracture (Munson et al. 1998; Qian et al. 2005).
(11)
where J is the hydraulic gradient, and V is the average velocity of flow. The hydraulic gradient can be expressed as
(12)
where L is the flow length between two piezometers and H1 and H2 are the corresponding piezometric heads. Substituting the value of J from Equation (12) in Equation (11), the following expression of K can be obtained (Qian et al. 2005):
(13)
The variation of K with the length of flow for a particular flow velocity was observed for different surface roughnesses and aperture sizes, as shown in Figures 810. The data demonstrate a linear trend between hydraulic conductivity and flow length, and with the increase in flow length, the value of hydraulic conductivity increased.
Figure 8

Variation of hydraulic conductivity with flow length for fine-walled fracture.

Figure 8

Variation of hydraulic conductivity with flow length for fine-walled fracture.

Close modal
Figure 9

Variation of hydraulic conductivity with flow length for medium-walled fracture.

Figure 9

Variation of hydraulic conductivity with flow length for medium-walled fracture.

Close modal
Figure 10

Variation of hydraulic conductivity with flow length for coarse-walled fracture.

Figure 10

Variation of hydraulic conductivity with flow length for coarse-walled fracture.

Close modal

Table 5 represents the K–L relation approximation in the form of linear function K = aL+b, where ‘a’ is the slope and ‘b’ is the intercept.

Table 5

Regression analysis of hydraulic conductivity vs. length of flow

Wall roughnessAperture (mm)V (m/s)ReEquation [1 m ≤ L ≤ 9.48 m]R
Fine 3.0 0.0154 22.9982 K = 0.0103 + 0.4558L 0.5937 
9.0 0.0158 70.9136 K = 0.0349 + 0.2582L 0.8761 
12.0 0.0153 91.602 K = 0.0124 + 0.4143L 0.4268 
Medium 3.0 0.0909 135.7523 K = 0.0835 + 0.1297L 0.7722 
9.0 0.0892 399.5073 K = 0.099 + 0.408L 0.7436 
12.0 0.0794 474.2173 K = 0.0522 + 0.5403L 0.8201 
Coarse 3.0 0.0388 58.0076 K = 0.0508 + 0.6139L 0.5822 
9.0 0.0348 155.8857 K = 0.036 + 0.6343L 0.4053 
12.0 0.0389 232.4186 K = 0.0379 + 0.7436L 0.2741 
Wall roughnessAperture (mm)V (m/s)ReEquation [1 m ≤ L ≤ 9.48 m]R
Fine 3.0 0.0154 22.9982 K = 0.0103 + 0.4558L 0.5937 
9.0 0.0158 70.9136 K = 0.0349 + 0.2582L 0.8761 
12.0 0.0153 91.602 K = 0.0124 + 0.4143L 0.4268 
Medium 3.0 0.0909 135.7523 K = 0.0835 + 0.1297L 0.7722 
9.0 0.0892 399.5073 K = 0.099 + 0.408L 0.7436 
12.0 0.0794 474.2173 K = 0.0522 + 0.5403L 0.8201 
Coarse 3.0 0.0388 58.0076 K = 0.0508 + 0.6139L 0.5822 
9.0 0.0348 155.8857 K = 0.036 + 0.6343L 0.4053 
12.0 0.0389 232.4186 K = 0.0379 + 0.7436L 0.2741 

The slope of the KL regression lines was found to increase with wall roughness for 9 and 12 mm aperture sizes, and it exhibited an irregular trend for 3 mm wall roughness. Additionally, the slope increased with aperture size for medium and coarse-wall fractures, while a decreasing slope trend was observed for fine-wall roughness with increased aperture size. The correlation coefficient was highest for fine-wall roughness fracture and hence was considered more suitable for the present discussion. From Figure 8 and Table 5, it is evident that the slope of K–L regression lines decreases with aperture size, implying a higher scale dependency for smaller apertures. However, this trend was not observed in coarse surface fracture due to a weaker correlation between K and L.

Hydraulic conductivity dependence on the scale for different hydraulic gradients

Figures 1113 illustrate the scale dependency of hydraulic conductivity for different hydraulic gradients with the same fracture aperture for fine, medium, and coarse-wall roughness. In general, an increase in the K value was observed with scale for medium and coarse surface fracture, as depicted in Figures 12 and 13. However, for fine surface fracture, as per Figure 11, a decrease in the K value was initially detected for the initial flow length, followed by an increasing trend for the later flow length. Additionally, it was also noticed that the higher hydraulic gradient values lead to smaller hydraulic conductivity values under the same conditions. This can be attributed to the increased friction between the rough fracture surface and the flow, as well as higher turbulent flow velocity induced by the higher hydraulic gradient. Furthermore, the slope of Hydraulic conductivity versus L was found to be less sensitive to the hydraulic gradient. Therefore, hydraulic conductivity can be inferred to be more scale-dependent on apertures and surface roughness than hydraulic gradients.
Figure 11

Hydraulic conductivity versus flow length for different hydraulic gradients in fine-walled fracture. J1, J2, J3, J4, J5, J6, J7, J8, J9, and J10 denote hydraulic gradients. They are 0.001899, 0.002532, 0.002321, 0.002215, 0.001477, 0.000949, 0.00116, 0.000738, 0.000844, and 0.00116, respectively, for 3 mm aperture; 0.00443, 0.003376, 0.002321, 0.00211, 0.002426, 0.002215, 0.001793, 0.002848, 0.001582, and 0.001477, respectively, for 9 mm aperture, and 0.007595, 0.007806, 0.007595, 0.006962, 0.008966, 0.00865, 0.008755, 0.008544, and 0.008439, respectively, for 12 mm aperture.

Figure 11

Hydraulic conductivity versus flow length for different hydraulic gradients in fine-walled fracture. J1, J2, J3, J4, J5, J6, J7, J8, J9, and J10 denote hydraulic gradients. They are 0.001899, 0.002532, 0.002321, 0.002215, 0.001477, 0.000949, 0.00116, 0.000738, 0.000844, and 0.00116, respectively, for 3 mm aperture; 0.00443, 0.003376, 0.002321, 0.00211, 0.002426, 0.002215, 0.001793, 0.002848, 0.001582, and 0.001477, respectively, for 9 mm aperture, and 0.007595, 0.007806, 0.007595, 0.006962, 0.008966, 0.00865, 0.008755, 0.008544, and 0.008439, respectively, for 12 mm aperture.

Close modal
Figure 12

Hydraulic conductivity versus flow length for different hydraulic gradients in a medium-walled fracture. J1, J2, J3, J4, J5, J6, J7, J8, J9, and J10 denote hydraulic gradients. They are 0.02173, 0.011181, 0.01846, 0.014557, 0.014768, 0.01519, 0.012869, 0.017405, 0.016561, and 0.0077, respectively, for 3 mm aperture; 0.11181, 0.009283, 0.009599, 0.009916, 0.009388, 0.0012658, 0.0111814, 0.006329, 0.010021, and 0.004536, respectively, for 9 mm aperture, and 0.007911, 0.010127, 0.008122, 0.007806, 0.00654, 0.006646, 0.005485, 0.004008, and 0.003797, respectively, for 12 mm aperture.

Figure 12

Hydraulic conductivity versus flow length for different hydraulic gradients in a medium-walled fracture. J1, J2, J3, J4, J5, J6, J7, J8, J9, and J10 denote hydraulic gradients. They are 0.02173, 0.011181, 0.01846, 0.014557, 0.014768, 0.01519, 0.012869, 0.017405, 0.016561, and 0.0077, respectively, for 3 mm aperture; 0.11181, 0.009283, 0.009599, 0.009916, 0.009388, 0.0012658, 0.0111814, 0.006329, 0.010021, and 0.004536, respectively, for 9 mm aperture, and 0.007911, 0.010127, 0.008122, 0.007806, 0.00654, 0.006646, 0.005485, 0.004008, and 0.003797, respectively, for 12 mm aperture.

Close modal
Figure 13

Hydraulic conductivity versus flow length for different hydraulic gradients in a coarse-walled fracture. J1, J2, J3, J4, J5, J6, J7, J8, J9, and J10 denote hydraulic gradients. They are 0.002426, 0.00116, 0.000844, 0.001055, 0.000949, 0.000844, 0.001371, 0.001266, 0.000949, and 0.000844, respectively, for 3 mm aperture; 0.002954, 0.006329, 0.006435, 0.005907, 0.002215, 0.002532, 0.002848, 0.002637, 0.002321, and 0.00211, respectively, for 9 mm aperture, and 0.003692, 0.003376, 0.003903, 0.001266, 0.001688, 0.001371, 0.00116, 0.001477, 0.00116, and 0.001055, respectively, for 12 mm aperture.

Figure 13

Hydraulic conductivity versus flow length for different hydraulic gradients in a coarse-walled fracture. J1, J2, J3, J4, J5, J6, J7, J8, J9, and J10 denote hydraulic gradients. They are 0.002426, 0.00116, 0.000844, 0.001055, 0.000949, 0.000844, 0.001371, 0.001266, 0.000949, and 0.000844, respectively, for 3 mm aperture; 0.002954, 0.006329, 0.006435, 0.005907, 0.002215, 0.002532, 0.002848, 0.002637, 0.002321, and 0.00211, respectively, for 9 mm aperture, and 0.003692, 0.003376, 0.003903, 0.001266, 0.001688, 0.001371, 0.00116, 0.001477, 0.00116, and 0.001055, respectively, for 12 mm aperture.

Close modal

Evaluation of Forchheimer's coefficient

Forchheimer's equation has proven excellent in describing fluid flow behavior in both discrete fracture and porous media. It can be expressed in general form (Bear 1972):
(14)
where ∇P [ML−1T−2] represents the pressure gradient along the flow direction, Q is the discharge, and A [ML−8] and B [ML−5T−1] represent the coefficients for energy losses due to inertial and viscous dissipation phenomena. The coefficients A and B can be expressed as:
(15a)
(15b)

The coefficient β in A depends on the geometry of the media, which needs to be calculated from experiments. With regard to B, k=c2/12 represents the intrinsic permeability of the fractured media. The main concern in the fracture is mainly coefficient A, in contrast to porous media, where both A and B need to be parameterized simultaneously.

Equation (14) can be further reduced in the form:
(16)
where J and V are hydraulic gradients and average flow velocity, respectively.
Forchheimer number (F0) in the Forchheimer law represents the ratio of pressure gradient required to overcome inertial to viscous forces. It can be expressed as the ratio of nonlinear (first term in Equation (14)) to linear (second term in Equation (14)) pressure losses. It accounts for both the structure and geometry of the medium.
(17)
Several critical F0 values have been suggested beyond which flow enters the non-Darcy flow zone. F0 = 0.11 (Zeng & Grigg 2006), F0 = 0.31 (Ghane et al. 2014), and F0 = 0.4 for natural sand (Macini et al. 2011). Forchheimer's number can be expressed in terms of Reynolds number as
(18)
The effect of confining stress, represented by hydraulic gradient, on Forchheimer's coefficients was examined and is illustrated in Figure 14. It can be observed that there is a significant increase in the coefficients A and B with an increase in hydraulic gradient for different values of β.
Figure 14

Plots between Forchheimer coefficient and Hydraulic gradient for different β values.

Figure 14

Plots between Forchheimer coefficient and Hydraulic gradient for different β values.

Close modal

This study aimed to investigate unconfined flow in a single fracture with uniform surface roughness and aperture sizes through well-controlled laboratory experiments. The experiments were designed to minimize the influence of the boundary layer by installing the first piezometer at a distance greater than the entrance length (>10 c) for turbulent flow through the media, ensuring that the flow behavior within the fracture was accurately captured.

The results indicated that the flow through the single fracture was non-Darcian for different surface roughnesses (fine, medium, and coarse) and aperture sizes (3, 9, and 12 mm). This conclusion was reached by calculating the Reynolds number and conducting regression analyses. The non-Darcian flow behavior observed in the study suggests that the Darcy equation is not suitable for describing groundwater flow at the laboratory scale, even within a smaller range of Reynolds numbers.

It is important to note that in this study, the flow tortuosity, which can be present in natural fractures due to factors such as varied apertures, multiple fractures, and non-uniform roughness, was not considered. Therefore, future studies should take these factors into account to better simulate real-world conditions.

The influence of inertial terms on the flow behavior was found to increase with surface roughness, with the trend being coarse > medium > fine. The regression analysis indicated that the Forchheimer and Izbash models were suitable for describing the prevailing non-laminar flow conditions. However, considering the simplicity and accuracy of the Izbash model, it is recommended for estimating hydraulic conductivity.

This study also revealed that hydraulic conductivity exhibited a linear increase with flow length. This can be attributed to the 2-D torturous flow within the rough fracture surface and additional inertial effects, resulting in energy dissipation within the fracture. The slope of the hydraulic conductivity vs. flow length relationship was found to be more dependent on surface roughness than aperture size. Thus, hydraulic conductivity is more scale-dependent on the surface roughness than the aperture and hydraulic gradient.

Furthermore, the results showed that hydraulic conductivity decreased with an increase in hydraulic gradient, assuming other conditions remained constant. This can be attributed to the increased friction between the rough fracture surfaces at higher hydraulic gradients, as well as the higher turbulent flow velocity induced by the increased hydraulic gradient.

It is important to acknowledge that the present study had certain limitations that could be addressed in future research to enhance the understanding of flow behavior in real-world conditions. Firstly, the study utilized an idealized single-fracture model with parallel fracture plates and uniform surface roughness. While this provided valuable insights into the flow characteristics within the fracture, it may not fully represent the complexities observed in actual field conditions, where multiple fractures can exist with varying surface roughness. Therefore, future studies should focus on investigating flow characteristics in multiple fractures with different aperture sizes to better simulate the field conditions. Secondly, experimental runs were conducted for unconfined flow conditions, which may not fully represent the real-world scenario. Therefore, confined flow conditions must be incorporated in future studies to simulate better the dynamic interactions between fractures and surrounding rock formations.

Additionally, a comparative analysis between flow conditions in natural fractures and single artificial fractures, as utilized in this study, would provide a more comprehensive perspective on fluid migration through the geological formations. In summary, by addressing these limitations and incorporating more representative experimental setups, future studies can significantly contribute to the advancement of knowledge regarding flow behavior through the fracture. This would ultimately lead to a better understanding of fluid migration in subsurface reservoirs, improving our ability to manage natural resources and make informed decisions in various engineering and environmental applications.

Based on the study, the following conclusions were reached:

  • (1)

    The flow through a single-fracture surface is essentially non-Darcian or fully turbulent under different surface roughness conditions. Reynolds number showed an increasing trend with average flow velocity, independent of the surface roughness. Decreasing the aperture size led to a decrease in Reynolds number for the same roughness, and the gradient of Reynolds number vs. average flow velocity increased with larger apertures.

  • (2)

    Non-Darcian turbulent flow conditions were observed within the medium to extensive Reynolds number range (12.34–474.22), confirming the nonlinear relationship between average velocity and hydraulic gradient found in previous studies.

  • (3)

    The ratio of calculated discharge (Qlcl) to experimental discharge (Qexp) was significantly greater than 1 for fractures with 9 and 12 mm apertures and different surface roughnesses. This suggests that no creeping flow occurred through the fracture, and the influence of inertial terms outweighed the viscous term. The impact of inertial terms was observed to increase with higher surface roughness. However, for a coarse surface fracture with a 12 mm aperture, a decreasing trend of Qlcl vs.Qexp indicated that a threshold value has been reached.

  • (4)

    Both Forchheimer and Izbash models were suitable for describing the non-laminar flow conditions prevailing in unconfined fractures with different wall roughness and aperture sizes.

  • (5)

    Hydraulic conductivity exhibited a linear increase in flow length for fractures with different apertures and surface roughnesses. The scale dependence of hydraulic conductivity on surface roughness was attributed to the torturous flow and energy dissipation within rough fractures.

  • (6)

    Higher hydraulic gradient values corresponded to smaller hydraulic conductivity values, assuming other conditions remained constant. This can be attributed to the increased turbulent flow velocity induced by a higher hydraulic gradient and increased friction between rough surfaces.

  • (7)

    The Forchheimer coefficient increased significantly with increasing confining stress for different β values of fine, medium, and coarse surface roughnesses until a critical value was reached.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Anand
A.
,
Beg
M.
&
Kumar
N.
2021
Experimental studies and analysis on mobilization of the cohesionless sediments through alluvial channel: a review
.
Civil Engineering Journal
7
(
5
),
915
936
.
https://doi.org/10.28991/cej-2021-03091700
.
Baker
D. L.
2000
A Darcian integral approximation to interblock hydraulic conductivity means in vertical infiltration
.
Computers & Geosciences 26 (5), 581–590. https://doi.org/10.1016/S0098-3004(99)00129-6
.
Bear
J.
1972
Dynamics of Fluids in Porous Media
.
New York: American Elsevier
.
Bear
J.
&
Corapcioglu
M. Y.
2012
Advances in Transport Phenomena in Porous Media
, Vol.
128
.
Springer Science & Business Media
.
https://doi.org/10.1007/978-94-009-3625-6
.
Berkowitz
B.
2002
Characterizing flow and transport in fractured geological media: a review
.
Advances in Water Resources
25
(
8–12
),
861
884
.
https://doi.org/10.1016/S0309-1708(02)00042-8
.
Bhanja
S. N.
,
Mukherjee
A.
,
Rangarajan
R.
,
Scanlon
B. R.
,
Malakar
P.
&
Verma
S.
2019
Long-term groundwater recharge rates across India by in situ measurements
.
Hydrology and Earth System Sciences
23
(
2
),
711
722
.
https://doi.org/10.5194/hess-23-711-2019
.
Brush
D. J.
&
Thomson
N. R.
2003
Fluid flow in synthetic rough-walled fractures: Navier-Stokes, Stokes, and local cubic law simulations
.
Water Resources Research
39
(
4
).
https://doi.org/10.1029/2002WR001346
.
Carrera
J.
1993
An overview of uncertainties in modeling groundwater solute transport
.
Journal of Contaminant Hydrology
13
(
1–4
),
23
48
.
https://doi.org/10.1016/0169-7722(93)90049-Xz
.
Chen
Y. F.
,
Zhou
J. Q.
,
Hu
S. H.
,
Hu
R.
&
Zhou
C. B.
2015
Evaluation of Forchheimer equation coefficients for non-Darcy flow in deformable rough-walled fractures
.
Journal of Hydrology
529
,
993
1006
.
https://doi.org/10.1016/j.jhydrol.2015.09.021
.
Chen
Z.
Narayan
S. P.
Yang
Z.
&
Rahman
S. S.
2000
An experimental investigation of hydraulic behaviour of fractures and joints in granitic rock
.
International Journal of Rock Mechanics and Mining Sciences
37
(
7
),
1061
1071
.
https://doi.org/10.1016/S1365-1609(00)00039-3
.
Chen
Z.
Qian
J. Z.
Luo
S. H.
&
Zhan
H. B.
2009
Experimental study of friction factor for groundwater flow in a single rough fracture
.
Journal of Hydrodynamics
21
(
6
),
820
825
.
https://doi.org/10.1016/S1001-6058(08)60218-8
.
Chen
Z.
,
Tian
Z.
,
Zhan
H.
,
Huang
J.
,
Huang
Y.
,
Wei
Y.
&
Ma
X.
2022
The effect of roughness on the nonlinear flow in a single fracture with sudden aperture change
.
Lithosphere
2022
(
Special 10
).
https://doi.org/10.2113/2022/5775275
.
Choi
E. S.
,
Cheema
T.
&
Islam
M. R.
1997
A new dual-porosity/dual-permeability model with non-Darcian flow through fractures
.
Journal of Petroleum Science and Engineering
17
(
3–4
),
331
344
.
https://doi.org/10.1016/S0920-4105(96)00050-2
.
Clauser
C.
1992
Permeability of crystalline rocks
.
Eos, Transactions American Geophysical Union
73
(
21
),
233
238
.
https://doi.org/10.1029/91EO00190
.
Cook
P. G.
2003
A Guide to Regional Groundwater Flow in Fractured Rock Aquifers
.
Seaview Press
,
Henley Beach, South Australia
, p.
151
.
Cornet
F. H.
,
Li
L.
,
Hulin
J. P.
,
Ippolito
I.
&
Kurowski
P.
2003
The hydromechanical behaviour of a fracture: an in situ experimental case study
.
International Journal of Rock Mechanics and Mining Sciences
40
(
7–8
),
1257
1270
.
https://doi.org/10.1016/S1365-1609(03)00120-5
.
Field
M. S.
&
Nash
S. G.
1997
Risk assessment methodology for karst aquifers: (1) estimating Karst conduit-flow parameters
.
Environmental Monitoring and Assessment
47
(
1
),
1
21
.
https://doi.org/10.1023/A:1005753919403
.
Ghane
E.
,
Fausey
N. R.
&
Brown
L. C.
2014
Non-Darcy flow of water through woodchip media
.
Journal of hydrology 519, 3400–3409. https://doi.org/10.1016/j.jhydrol.2014.09.065
.
Hakami
E.
&
Larsson
E.
1996
Aperture measurements and flow experiments on a single natural fracture. In International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts (Vol. 33, No. 4, pp. 395–404). Pergamon. https://doi.org/10.1016/0148-9062(95)00070-4.
Ikhsan
C.
,
Permana
A. S.
&
Negara
A. S.
2022
Armor layer uniformity and thickness in stationary conditions with steady uniform flow
.
Civil Engineering Journal
8
(
06
).
http://dx.doi.org/10.28991/CEJ-2022-08-06-01
.
Kohl
T.
,
Evans
K. F.
,
Hopkirk
R. J.
,
Jung
R.
&
Rybach
L.
1997
Observation and simulation of non-Darcian flow transients in fractured rock
.
Water Resources Research
33
(
3
),
407
418
.
https://doi.org/10.1029/96WR03495
.
Lamb
H.
1932
Hydrodynamics, § 156. Aufl., Cambridge University Press 18791932, 427
.
Macini
P.
,
Mesini
E.
&
Viola
R.
2011
Laboratory measurements of non-Darcy flow coefficients in natural and artificial unconsolidated porous media
.
Journal of Petroleum Science and Engineering 77 (3–4), 365–374. https://doi.org/10.1016/j.petrol.2011.04.016
.
Munson
B. R.
,
Young
D. F.
&
Okiishi
T. H.
1998
Fundamentals of fluid mechanics, third edition, Wiley, New York, USA
.
Neuman
S. P.
1994
Generalized scaling of permeabilities: validation and effect of support scale
.
Geophysical Research Letters
21
(
5
),
349
352
.
https://doi.org/10.1029/94GL00308
.
Neuman
S. P.
&
Di Federico
V.
2003
Multifaceted nature of hydrogeologic scaling and its interpretation
.
Reviews of Geophysics
41
(
3
).
https://doi.org/10.1029/2003RG000130
.
Oron
A. P.
&
Berkowitz
B.
1998
Flow in rock fractures: the local cubic law assumption re-examined
.
Water Resources Research
34
(
11
),
2811
2825
.
https://doi.org/10.1029/98WR02285
.
Polubarinova-Kochina
P. I.
1962
Theory of water movement
.
Princeton University Press
.
Qasim
R. M.
,
Mohammed
A. A.
&
Abdulhussein
I. A.
2022
An investigating of the impact of bed flume discordance on the weir-gate hydraulic structure
.
HighTech and Innovation Journal
3
(
3
),
341
355
.
https://doi.org/10.28991/HIJ-2022-03-03-09
.
Qian
J.
,
Zhan
H.
,
Zhao
W.
&
Sun
F.
2005
Experimental study of turbulent unconfined groundwater flow in a single fracture
.
Journal of Hydrology
311
(
1–4
),
134
142
.
https://doi.org/10.1016/j.jhydrol.2005.01.013
.
Qian
J.
,
Zhan
H.
,
Luo
S.
&
Zhao
W.
2007
Experimental evidence of scale-dependent hydraulic conductivity for fully developed turbulent flow in a single fracture
.
Journal of Hydrology
339
(
3–4
),
206
215
.
https://doi.org/10.1016/j.jhydrol.2007.03.015
.
Qian
J.
,
Zhan
H.
,
Chen
Z.
&
Ye
H.
2011a
Experimental study of solute transport under non-Darcian flow in a single fracture
.
Journal of Hydrology
399
(
3–4
),
246
254
.
https://doi.org/10.1016/j.jhydrol.2011.01.003
.
Qian
J.
,
Chen
Z.
,
Zhan
H.
&
Guan
H.
2011b
Experimental study of the effect of roughness and Reynolds number on fluid flow in rough-walled single fractures: a check of local cubic law
.
Hydrological Processes
25
(
4
),
614
622
.
https://doi.org/10.1002/hyp.7849
.
Rong
G.
,
Tan
J.
,
Zhan
H.
,
He
R.
&
Zhang
Z.
2020
Quantitative evaluation of fracture geometry influence on nonlinear flow in a single rock fracture
.
Journal of Hydrology
589
,
125162
.
https://doi.org/10.1016/j.jhydrol.2020.125162
.
Schulze-Makuch
D.
,
Carlson
D. A.
,
Cherkauer
D. S.
&
Malik
P.
1999
Scale dependency of hydraulic conductivity in heterogeneous media
.
Groundwater
37
(
6
),
904
919
.
https://doi.org/10.1111/j.1745-6584.1999.tb01190.x
.
Tsang
Y. W.
1984
The effect of tortuosity on fluid flow through a single fracture
.
Water Resources Research
20
(
9
),
1209
1215
.
https://doi.org/10.1029/WR020i009p01209
.
Tzelepis
V.
,
Moutsopoulos
K. N.
,
Papaspyros
J. N.
&
Tsihrintzis
V. A.
2015
Experimental investigation of flow behavior in smooth and rough artificial fractures
.
Journal of Hydrology
521
,
108
118
.
https://doi.org/10.1016/j.jhydrol.2014.11.054
.
White
F. M.
1990
Fluid Mechanics
.
McGraw Hill, New York
.
Widyastuti
I.
,
Thaha
M. A.
,
Lopa
R. T.
&
Hatta
M. P.
2022
Dam-break energy of porous structure for scour countermeasure at bridge abutment
.
Civil Engineering Journal
8
(
12
),
3939
3951
.
https://doi.org/10.28991/CEJ-2022-08-12-019
.
Xing
K.
,
Qian
J.
,
Ma
L.
,
Ma
H.
&
Zhao
W.
2021
Characterizing the scaling coefficient ω between viscous and inertial permeability of fractures
.
Journal of Hydrology
593
,
125920
.
https://doi.org/10.1016/j.jhydrol.2020.125920
.
Yan
X.
,
Qian
J.
,
Ma
L.
,
Wang
M.
&
Hu
A.
2018
Non-Fickian solute transport in a single fracture of marble parallel plate
.
Geofluids
2018
.
https://doi.org/10.1155/2018/7418140
.
Zhang
Z.
&
Nemcik
J.
2013
Fluid flow regimes and nonlinear flow characteristics in deformable rock fractures
.
Journal of Hydrology 477, 139–151. https://doi.org/10.1016/j.jhydrol.2012.11.024
.
Zeng
Z.
&
Grigg
R.
2006
A criterion for non-Darcy flow in porous media
.
Transport in porous media 63, 57–69. https://doi.org/10.1007/s11242-005-2720-3
.
Zimmerman
R. W.
&
Bodvarsson
G. S.
1996
Hydraulic conductivity of rock fractures
.
Transport in Porous Media
23
(
1
),
1
30
.
https://doi.org/10.2172/60784
.
Zimmerman
R. W.
&
Yeo
I. W.
2000
Fluid flow in rock fractures: from the Navier-Stokes equations to the cubic law
.
Geophysical Monograph-American Geophysical Union
122
,
213
224
.
https://doi.org/10.1029/GM122p0213
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).