Next Article in Journal
Efficient Removal of Cadmium (II) and Arsenic (V) from Water by the Composite of Iron Manganese Oxides Loaded Muscovite
Previous Article in Journal
A Review of Analytical Methods and Technologies for Monitoring Per- and Polyfluoroalkyl Substances (PFAS) in Water
Previous Article in Special Issue
A 2D Hydraulic Simulation Model Including Dynamic Piping and Overtopping Dambreach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Unified General Resistance Formula for Uniform Coarse Porous Media

by
Juan Carlos López
1,*,
Miguel Ángel Toledo
1,2,
Rafael Moran
1,2 and
Luis Balairón
3
1
Department of Civil Engineering: Hydraulics, Energy and Environment, SERPA Dam Safety Research Group E.T.S. de Ingenieros de Caminos Canales y Puertos, Universidad Politécnica de Madrid (UPM), Profesor Aranguren, s/n, 28040 Madrid, Spain
2
Centre Internacional de Metodes Numerics en Enginyeria (CINME), Campus Norte UPC, Gran Capitán, s/n, 08003 Barcelona, Spain
3
Hydraulics Laboratory, Centro de Estudios Hidrográficos, Centro de Estudios y Experimentación de Obras Públicas (CEDEX), 28005 Madrid, Spain
*
Author to whom correspondence should be addressed.
Water 2023, 15(20), 3578; https://doi.org/10.3390/w15203578
Submission received: 15 September 2023 / Revised: 6 October 2023 / Accepted: 9 October 2023 / Published: 13 October 2023

Abstract

:
Many authors studied the nonlinear relationship between seepage velocity and hydraulic gradient in coarse granular materials. They managed different approaches and variables to define the resistance formula applicable to that type of granular media. Based on the analysis of the different approaches and experimental data obtained by the corresponding authors, we propose a unified general seepage equation applicable to large-sized granular materials. Such an equation gives unity to the main nonlinear resistance formulas developed to date. Particularly relevant are the conclusions regarding the relationship of the linear ( α * ) and quadratic ( β * ) dimensionless coefficients of the resistance formula with the representative size of the particle and the geometrical features of the porous materials.

1. Introduction and Objectives

Seepage through coarse porous media, such as gravels or ripraps, with a high pore size creates so-called non-Darcy flows, which can be represented by the quadratic equation of Forchheimer [1], Equation (1), or by the exponential equation presented by S.V. Izbash [2], Equation (2):
i = r · V + s · V 2
i = a · V b
where  V  is the mean seepage velocity defined as the mean fluid velocity over the entire cross section,  i  is the hydraulic gradient,   r  represents the coefficient of the linear term,  s  represents the coefficient of the quadratic term, and  a  and  b  are coefficients that can be determined experimentally for each porous media.
The Forchheimer equation, Equation (1), contemplates a wide range of flow regimes: nonlinear laminar, turbulent transition, and fully developed turbulent flow, which make it more suitable than the exponential Equation (2) to represent the physical phenomenon of filtration in porous media.
However, it is necessary to deepen the understanding of this physical phenomenon in these transition regimes by analyzing whether the coefficients  r  and  s  remain constant for a wide range of gradients.
On the other hand, based on Equation (1), different formulations have been developed that consider different physical parameters. J.C. López et al. [3] have recently published a comparative analysis of these formulations based on the approach of a generalized quadratic equation based on three physical parameters: the Darcy–Weisbach generalized friction factor  f , the generalized Reynolds number  R e , and the generalized characteristic length  L c .
From this generalized quadratic equation, this research work aims to develop a general unified equation for coarse porous media of uniform granulometry that can be applied to the so-called non-Darcy flows.
Within the development of the USEC, the obtaining of the dimensionless linear coefficient  α *  included in the coefficient  r  and the quadratic coefficient  β *  included in the coefficient  s , both related to the particle  D  representative size and the geometric characteristics of the porous media, is of special importance.
On the other hand, we also highlight the analysis of the application of the USEC to a wide range of gradients where the transition regimes are developed.
The research carried out in this article includes the presentation of a conceptual scheme obtained from the review of the existing formulation and its experimental verification with permeability tests carried out at the Hydraulics Laboratory of the Centro de Estudios Hidrográficos (CEDEX) on porous media composed of crushed aggregates duly selected to obtain a uniform granulometry and coming from the same limestone quarry (i.e., with the same geological origin and similar geometric properties).

2. State of the Art

2.1. Review of the Formulation and Conceptual Scheme

Most of the investigations have based the development of their formulas on the analogy with the flow in pipes by means of the applications of two dimensionless groups that we will call the generalized friction factor of the Darcy–Weisbach  f  Equation (3) and the generalized Reynolds number  R e  Equation (4) (J.C. López et al. [3] and J. Bear [4]).
f = L c · 2 g · i V p 2
R e = L c · V p v
where  L c  is the generalized characteristic length,  g  is the gravitational acceleration, i  is the hydraulic gradient,  v  is the kinematic viscosity, and  V p  is the velocity in the pores, determined by Equation (5).
V p = V n
where  n  is the porosity of the porous media.
As indicated by J.C. López et al. [3] and J. Bear [4], most of the researchers agree in pointing out that the experimental data corresponding to a granular porous media of the same geometry and size properly represented in a generalized diagram of the type  R e , f  fall on a smooth curve for all non-Darcy flow regimes that can be fitted by Equation (6).
f = A 1 R e + A 2
where  A 1  is referred to us as a linear generalized dimensionless coefficient and  A 2  as a quadratic generalized dimensionless coefficient.
Equation (6) is referred to us as a generalized equation in a  R e , f  diagram form.
However, the existence of these continuous curves in granular porous media contrast with the physical behavior of turbulent flow in pipes where sharp jumps occur in the Reynolds number of the pipe range  R e t  between 2000 and 4000 (F. W. White [5]).
Substituting the generalized values   f  (Equation (3)),  R e  (Equation (4)), and pore velocity  V p  (Equation (5)) into Equation (6), we obtain the generalized quadratic Equation (7):
i = v 2 · g · A 1 · 1 L c 2 · V n + 1 2 · g · A 2 · 1 L c · V 2 n 2
Recently, J.C. López et al. [3] have performed a comparative analysis of different existing formulations based on three characteristics lengths,  L c : the representative diameter of the particles  D  and the square root of the intrinsic permeability  K o   ( K o ) , which is a macroscopic property of porous media given by Equation (8):
K 0 = v g · 1 i · V
and the mean hydraulic radius  R h  that was first defined by Taylor [6] and is given by Equation (9):
R h = n S e ( 1 n )
where  S e  is the average specific surface area of the solid particles constituting the granular porous media and depends on the size, shape, angularity, and surface roughness of the particles (Loudon, Linford and Sounders, Crawford et al., Martins and Escarameria, Martins and Sabin and Hansen [7,8,9,10,11,12]) and is defined by Equation (10):
S e = 6 · F D
where  F  is a dimensionless coefficient that considers shape, angularity, and surface roughness of the particles, as a whole, to be constituents of the porous granular media and represents the frictional energy loss addition to that produced in granular porous media consisting of smooth spherical particles as a result of having a higher average specific surface  S e .
Next, we will analyze each of the parameters involved in Equation (7).
  • Characteristic Length Lc
Based on the analogy with the flow in pipes for a noncircular section, as considered by J.C. López et al. [3], we will consider the characteristic length  L c  the mean hydraulic diameter  D h  (Equation (11)):
D h = 4 · R h
The characteristic length  D h  represents a physical parameter within a microscale model (Ergun and Orning [13] and Huang and Ayoub [14]) that is related to the size and shape of the interconnected channels of the porous media.
Substituting Equation (10) into Equation (9) and considering Equation (11), we obtain Equation (12):
D h = 2 3 · 1 F · n ( 1 n ) · D
The mean hydraulic diameter  D h  contemplates three physical parameters: the representative diameter of the constituent particles of the porous media  D , which is related to the pore size, the dimensionless coefficient  F , and the porosity  n , both latter parameters related to the pore shape. The porosity parameter n is defined in (Equation (12)) by the porosity function  f ( n )   (Equation (13)):
f ( n ) = n ( 1 n )
If we consider as the characteristic length only the representative diameter of the particles  D , we are actually facing a simpler model within the analogy with the flow in pipes, which does not consider these last two fundamental physical parameters to properly characterize the granular porous media: the dimensionless coefficient  F   and the porosity function  f ( n ) .
On the other hand, the square root of intrinsic permeability,  K 0 , actually represents a macroscopic property of the granular porous media, as is the case with the parameters  r  and  s  considered in the quadratic Equation (1) (H. Huang and J. Ayoub [14]).
  • Generalized dimensionless coefficients A1, linear and A2, quadratic.
However, in the microscale model of the granular porous media represented by the characteristic length  D h , in contrast to the flow in pipes, the channels through which the water flows are tortuous and of a non-constant section with a very complex geometry. Therein, the flow experiences continuous cycles of acceleration and deceleration with the consequential loss of associated energy. At velocities much lower than those necessary to produce the turbulent effects, inertial effects appear because of the sinuous trajectories (Ahmed and Sunada [15]), producing nonlinearity (nonlinear laminar regime) through quadratic Equation (1). When velocities increase, the onset of turbulence is attributed to the phenomena of separation of the boundary layer and the corresponding formation of vortices (Wright, Panfilov et al. and Fourar et al. [16,17,18]), phenomena all related to particle shape and angularity.
On the other hand, McCorquodale et al. [19], using as characteristic length  L c  the mean hydraulic radius  R h , were the first to introduce the concept of particle surface roughness  ɛ  in a porous media in an analogy with the flow of pipes, indicating that “The effect of surface roughness is usually ignored; however, for the high velocities and large particle sizes, this is probably not justified”.
The conclusion of the experimental results carried out by J. Mulqueen [20] on gravels indicate that the flow velocity was declining substantially with increased surface roughness.
As a consequence, in this microscale model, based on the behavior of flow in pipes, the physical parameters, shape, angularity, and surface roughness of the particles in addition to their size, need to be considered to adequately characterize the geometry of the granular porous media. These physical parameters are contemplated through the  A 1  linear and  A 2  quadratic dimensionless coefficients represented in the generalized quadratic Equation (7).
With respect to the linear generalized dimensionless coefficient  A 1 , Terzaghi and Peck, Engelund, and Salahi et al. [21,22,23] consider them to be a function of the shape and angularity of particles. On the other hand, the value of this dimensionless coefficient for these non-Darcy flows is different from that obtained for the laminar regime (Fand et al. [24]; Barree and Conway [19,20,21,22,23,24,25]).
With respect to the quadratic generalized dimensionless coefficient  A 2 , Gupta [26], based on Shergold’s previous research [27], related both concepts, the geometry of the granular porous material and the representative size of particles  D , by function Equation (14):
η = f · C , D = C ·   D n
where  η  represents the angularity as a set of aggregates,  C  is a coefficient of form that depends solely on the shape of the particles as a whole and is constant for each type of porous material and is independent of its size, and  n  is an exponent.
Gupta used materials of various shapes, with uniform granulometry and a wide range of sizes: rounded gravels (6 m <   D   < 100 mm), subangular crushed quartzite (6 mm <  D   < 50 mm), and angular limestone (6 mm <   D   < 50 mm). He obtained the following values:  C  = 37.89 for rounded gravels,  C  = 41.44 for subangular crushed quartzite, and  C   = 43.35 for angular limestone. The value of the adjusted exponent  n  for these granular porous media was −0.032, so according to Equation (14), this exponent did not vary with the shape of the particles defined through the shape coefficient  C .
In this same line, Dudgeon [28] stated that “the angularity of the particles increases with the decrease in size”.
As indicated by J.C. López et al. [3], several authors in the development of their formulations (Ergun and Orning [13]; Engelund [22]; Stephenson [29] and Martins [11]) have limited themselves to proposing constant values of these generalized dimensionless coefficients,  A 1  and  A 2 , to be applied according to the three geometries representative of the granular porous material: smooth spheres, rolled aggregates, and crushed aggregates.
However, in this article, the knowledge of the seepage phenomenon in coarse porous media is further deepened by including in the general unified equation to be developed, based on Equation (7), empirical equations of the types,  A 1 = F 1 ( D )  and  A 2 = F 2 ( D ) , that relate the dimensionless linear  A 1  and quadratic  A 2  coefficients with the representative size of the particles  D . These equations have been grouped in the three representative geometries mentioned above and will be adjusted with the experimental data from previous studies described in the following section and experimentally validated in Section 4.

2.2. Experimental Data from Previous Studies

Experimental data have been collected from several authors in order to properly calibrate the unified overall equation (Sabri Ergun and A.A. Orning [13]; Ward [30]; Dudgeon [28]; Ahmed and Sunada [15]; Arbhabhirama and Dinoy [31]; Martins [11]; Sedghi-Als et al. [32]; M-B Salahi et al. [23]; Ferdos [33]). Such data have been appropriately selected to provide coefficients  r  and  s  of the Forchheimer equation (Equation (1)).
These experimental data have been included in this research work shown in Appendix A, and their main characteristics have been described: types of aggregates tested, size ranges  D  used, diameter of the permeameter  D x  of the installation, and hydraulic gradients  i  applied.

3. The Unified Seepage Equation for Coarse Materials

In this section, the unified seepage equation for coarse materials (USEC) from the generalized quadratic equation (Equation (7)) has been developed using the experimental data included in Appendix A. The USEC is based on the analogy with the flow in pipes and properly calibrated by obtaining empirical equations that relate the generalized dimensionless coefficients, linear  A 1  and quadratic  A 2 , with the representative size of the particles  D . Section 3.3 also includes its application to existing experimental data to verify the continuity of the curve given by Equation (6) and has served to obtain preliminary considerations on its applicability. Next, in Section 4, we have proceeded to validate the USEC with the additional experimental data obtained from new seepage tests performed in the Laboratory of Hydraulics at the Centro de Estudios Hidrográficos (CEDEX).

3.1. Methodology

The methodology for properly developing the USEC will consist of:
(a)
Selecting as characteristic length  L c  the average hydraulic diameter  D h  defined by Equation (11) with the physical parameters  D n , and  F . Equation (11), together with Equation (7), will result in a unified general equation that contemplates the unified dimensionless coefficients  α *  (linear) and  β *  (quadratic), equivalent to the coefficients  A 1  and  A 2  from generalized quadratic Equation (7). The rationale for this new formulation is developed in Section 3.2 below.
(b)
Collecting the experimental data duly selected according to the type of granular porous media, smooth spheres, rolled aggregates, and crushed aggregates, in order to calculate the linear term  r  and the quadratic term  s  of the Forchheimer equation (Equation (1)) that appear in Appendix A.
(c)
Analyzing with the available experimental data from Appendix A whether the  α *  linear and  β *  quadratic dimensionless coefficients for a granular porous media defined by a given geometry and size are constant for all non-Darcy flow regimes, nonlinear laminar, turbulent transition, and turbulent fully developed, as most research seems to confirm (J.C. López et al. [3]), or conversely, whether they need to be adjusted for each particular non-Darcy flow regime. This analysis is developed in Section 3.3 with the experimental data used, and in the absence of further research in this regard, we can consider as a hypothesis that the USEC can be applied to all non-Darcy flow regimes.
(d)
Obtaining the empirical equations relating the dimensionless USEC coefficients  α *  and  β *  to the representative particle size  D  for granular porous media included in Appendix A. These empirical equations have been grouped into three representative geometries, smooth spheres, rolled aggregates, and crushed aggregates, in order to analyze where there are similarities between them. The values obtained have been represented in the type [ D α * ] and [ D β * ] diagrams in order to analyze their evolution regarding the representative size of particles  D . Section 3.4 and Section 3.5 develop these aspects.
(e)
Finally, in Section 4, we have analyzed the results obtained by applying the newly developed USEC to the experimental data obtained in the tests carried out at the Hydraulics Laboratory of the Centro de Estudios Hidrográficos (CEDEX). The granular porous media used in these tests consist of crushed aggregates with uniform granulometry and are from the same limestone rock quarry, so the dispersions in the geometric properties of the same are minimal.

3.2. Justification for the New Formulation

According to the analysis carried out in Section 2.1, we will select as characteristic length  L c  the mean hydraulic diameter  D h . Substituting the value of  D h  from Equation (12) into the generalized quadratic Equation (7), we obtain Equation (15):
i = v 2 g · A 1 · 3 2 · F 2 2 2 · D 2 · ( 1 n ) 2 n 2 · V n + 1 2 g · A 2 · 3 · F 2 · D · ( 1 n ) n · V 2 n 2
from which, simplifying, we obtain Equation (16):
i = v g · 9 8 · A 1 · F 2 · ( 1 n ) 2 n 3 · 1 D 2 · V + 1 g · 3 4 · A 2 · F · ( 1 n ) n 3 · 1 D · V 2
Considering a similar approach to Engelund [22], we will define the unified dimensionless coefficients α* linear and β* quadratic by Equations (17) and (18), respectively:
α * = 9 8 · A 1 · F 2  
β * = 3 4 · A 2 · F  
Substituting these coefficients,  α *  and  β * , in Equation (16), we obtain Equation (19):
i = v g · α * · ( 1 n ) 2 n 3 · 1 D 2 · V + 1 g · β * · ( 1 n ) n 3 · 1 D · V 2
Equation (19) includes the porosity function of the linear term (Equation (20)):
f L ( n ) = ( 1 n ) 2 n 3
and the porosity function of the quadratic term (Equation (21)):
f T ( n ) = ( 1 n ) n 3
We will refer to Equation (19) as the USEC in quadratic form, applicable for coarse granular porous media with uniform particle size. Each granular porous media will therefore be defined by the physical parameters considered therein, representative size of particles  D , porosity functions of linear  f L ( n ) , and quadratic  f T ( n )  terms, and finally the unified dimensionless coefficients linear  α *  and quadratic  β *  that characterize the geometry of the porous material.
Knowing parameters  r  and  s  of Forchheimer’s Equation (1) of a given granular porous media, the unified dimensionless coefficients α* and β* that uniquely characterize granular porous media can be obtained by means of the expressions:
α * = r · g · n 3 · D 2 v · ( 1 n ) 2
β * = s · g · n 3 · D ( 1 n )
In Appendix A, the unified dimensionless coefficients  α *  and  β *  are calculated by applying Equations (22) and (23) based on the experimental data used in this research work.
On the other hand, USEC Equation (19) can be represented dimensionlessly in a [ R p , f p ] diagram in order to visualize the continuity of the curves given by Equation (6).
The pore Reynolds number  R p  based on  D h  will be defined by Equation (24):
R p = 4 · R h · V p v = 4 v · n S e · ( 1 n ) · V n = 2 3 · 1 v · 1 F · 1 ( 1 n ) · D · V
The pore friction factor  f p  based on  D h  will be defined by Equation (25):
f p = 8 · R h · g · i V p 2 = 8 · g · n S e · 1 n · n 2 V 2 · i = 4 3 · g · 1 F · n 3 1 n · D · i V 2
With it, Equation (6), particularized for the pore friction factor  f p  and the pore Reynolds number  R p , will remain defined by Equation (26):
f p = 8 3 · 1 3 · 1 R p · α * F 2 + 1 2 · β * F
being  A 1 :
A 1 = 8 9 · α * F 2
being  A 2 :
A 2 = 4 3 · β * F
We will refer to Equation (26) as USEC in a  R p , f p  diagram form.
Finally, we will linearize Equation (26). For this, we will consider the linear Equation (29):
λ p = t + u · R p
where  t , y, and  u  are parameters obtained by fitting the experimental data [ R p , f p ], and  λ p  is the linearized pore friction factor given by Equation (30).
λ p = f p · R p
To obtain the parameters  t  and  u , we will multiply both members of Equation (25) by the pore Reynolds number  R p   (Equation (31)):
f p · R p = 8 9 · α * F 2 + 4 3 · β * F · R p
Thus, the unified dimensionless coefficients  α *  and  β *  are determined by Equations (32) and (33):
α * = t · 9 8 · F 2
β * = u · 3 4 · F
We will call Equation (29) USEC in linearized form, which can be represented in a  R p , λ p  diagram.

3.3. Study of the Continuity of the Equation in Non-Laminar Flow Regimes

In order to verify the continuity of the curve given by Equation (26) for a given porous media in a wide range of gradients that includes all non-Darcy flows, we will use USEC expressed in the [ R p , f p ] diagram, Equation (26), and in the [ R p , λ p ] diagram, Equation (29).
  • Flow regimes: nonlinear laminar, turbulent transition
In Figure 1 are represented in a  R p , λ p  diagram the experimental data obtained by Dudgeon [28] corresponding to the granular porous media composed of smooth spheres that have the advantages of a well-defined geometry and a wide range of gradients, 1.80 × 10−4 i   < 8.00 (N = 14). In this case, the representative particle size  D  is the value  D 50  of the granulometric curve, 15.97 mm.
The porosities  n   of 36.90% (N = 5), 41.50% (N = 5), and 37.20% (N = 4) were used for each of the three tests carried out. It can be seen how these data fall on USEC in the linearized form defined by Equation (29), which implies the existence of a continuous curve as given by Equation (26). A good fit has been obtained for parameters  t  and   u  (R2 = 0.9972), obtaining the unified dimensionless coefficients α* = 232.01 and β* = 1.26 by applying Equations (32) and (33), respectively, and considering a value of the dimensionless coefficient  F  = 1.00 for smooth spheres.
Thus, it is observed that each granular porous media with a particle size determined by its representative size  D  and consisting as a whole of particles with a shape, surface, angularity, and surface roughness determined by the dimensionless coefficient  F  will be represented by a line whose slope  u  allows the calculation of the quadratic dimensionless coefficient β* in accordance with Equation (33), and the cut with the ordinate axis   t  allows the calculation of the linear dimensionless coefficient α* according to Equation (32).
  • Fully developed turbulent flow
When the fully developed turbulent flow is reached, exponent  b  of the exponential Equation (2) is equal to 2 (J.C. López et al. [3]), and as a consequence, parameters  a   of Equation (2) and  s  of Equation (1) are equalized, and parameter  r  of the linear component of the quadratic Equation (1) becomes zero. With this, USEC in quadratic form for fully developed turbulent flow will be given by the expression (Equation (34)):
i = 1 g · β * · ( 1 n ) n 3 · 1 D · V 2
On the other hand, USEC in a [ R p , f p ] diagram form will be defined by the horizontal asymptote (Equation (35)):
f p = 4 3 · β 3 F
Finally, the USEC in linearized form (Equation (29)) would be represented in a [ R p , λ p ] diagram by a straight line passing through the origin of the ordinates by the expression (Equation (36)):
λ p = u · R p
where  β *  is defined by Equation (33).
Figure 2 shows the experimental data corresponding to two tests of Martins [11] for rolled aggregates ( F  = 1.05 proposed by Martins) in a diagram [ R p , f p ]. In this case, the representative particle size corresponds to the average sieve aperture  D a  with a value of 22.00 mm. The porosities  n  of the two tests were 36.22% and 36.83%, taking place in the zone of the fully developed turbulent regime, and are represented in the diagram [ R p , f p ]. Values  R p  and  f p  have been obtained by applying Equations (24) and (25) from the experimental data [ V , i ] (N = 8).
It is observed for both tests that 7 of the 8 points are aligned on a horizontal asymptote defined by Equation (35) with mean values of  f p  = 1.62 for the porous media with a porosity of  n  = 36.22% and  f p  = 1.66 for the porous media with a porosity of  n  = 36.83. The fully developed turbulent regime has been reached in this case for a pore Reynolds number  R p  = 1049 for  n = 36.22 %  and  R p  = 1034 for  n = 36.83 %  (see Table 1).
Figure 3 shows all the points (N = 16) corresponding to the tests represented in Figure 2. A fit for parameter  u  = 1.6386 (R2 = 0.9974) has been obtained considering 14 of the 16 points. The value  t  = −0.0656 is close to zero as a consequence of the tests being carried out in the fully developed turbulent regime (Equation (35)). Thus, applying Equation (32), we obtain a value of β* = 1.29 for the set of the two porous media evaluated ( n  = 36.22% and  n  = 36.83%), with the representative particle size  D  = 22.00 mm.
Table 1 shows all the values of the unified quadratic dimensionless coefficient β* corresponding to all the rolled aggregates tested by Martins [11].
For each representative size  D , three values of the unified quadratic dimensionless coefficient β* have been obtained, two values with different porosities for each size by applying Equation (34) and one value by applying Equation (35), by fitting all the points of each size  D . The three values obtained are close to each other, which highlights the existence of similar geometries for each size  D .

3.4. Analysis of the Effect of Representative Size  D  on the Unified Coefficients α* and β*

In accordance with Section 2.1 on the state of the art, the dimensionless coefficients A1 and  F  are a function of the geometrical properties of the porous material. Therefore, the unified dimensionless coefficient α* will also depend on the geometry of the granular porous media, according to Equation (16).
Figure 4 shows the values of α* in a [ D ,α*] diagram obtained from the tests carried out by Sedghi-Als et al. [32] and M-B Salahi et al. [23] in the size range 1.77 mm <   D   < 56.80 mm, which have the advantage of having a higher density of points (N = 18) in this size range. In this figure, an increase in the unified dimensionless coefficient α* with the representative particle size  D  can be observed for all three porous media.
The values of the unified dimensionless coefficient α* for crushed aggregates are in general slightly above the values for the rolled aggregates, except for in the test for  D  = 13.08 mm. All this seems to indicate that, with the experimental data used and with a representative particle size  D , the values of the unified dimensionless coefficient α* for crushed aggregates are higher than the corresponding values for rolled aggregates, as shown in the research work of Terzaghi [21] and Engelund [22]. As a consequence, the linear dimensionless coefficient α* should vary according to the geometry of the porous material and the representative size of the particles  D .
In the case of the unified quadratic dimensionless β*, the analysis is based on the analogy with the flow in pipes. According to Colebrook [35], the calculation of the coefficient of turbulent friction  f t  in the pipes is given by the expression:
1 f t 1 / 2 = 2.0 l o g · ε t D t 3.7
where  D t  is the diameter of the pipe, and  ε t  is the absolute pipe roughness.
Equation (38) allows the obtaining of the asymptotic value of the friction coefficient  f t    and is independent of the Reynolds number  R e  of the pipe. In Figure 5, we have plotted two curves of type  f t  =  F · D t , applying Equation (37). Each curve is represented with a geometry defined by means of its absolute roughness  ε t . For similarity, we have adopted the absolute particle surface roughness ε proposed by McCorquodale et al. [18] ( ε t  = 0.1 mm for the curve corresponding to crushed aggregates and  ε t  = 0.02 mm for the curve corresponding to rolled aggregates).
It can be observed from Figure 5 that:
(a)
For the same absolute pipe roughness  ɛ t , the friction coefficient  f t  can be fitted by a smooth curve  f t = F ( D t ) . The friction coefficient  f t  decreases as the diameter of the pipe  D t  increases. For high values of  D t , the friction coefficient  f t  tends to an asymptotic value.
(b)
As the roughness of the pipe  ε t   decreases, curves  f t = F D t  approach the horizontal axis with lower values of the friction coefficients,  f t , for the same diameter of the pipe,  D t .
Since the dimensionless coefficient  F , which affects the specific surface area  S e , can be considered constant for a given geometry of the porous media according to Equation (9), we can consider by the analogy with the flow in pipes that (a) the parameter  D t  is equivalent to the parameter  D  through Equation (12) and (b) the parameter  f t  is equivalent to the parameter  f p  and, consequently, with parameter β* through Equation (35). Thus, in flows in porous media with remarkably similar fixed geometries, there should be curves of the type β*=  F ( D ) , equivalent to curves  f t = F ( D t ) .
In Table 1 for the rolled aggregates tested by Martins [11], a decrease in the quadratic dimensionless coefficient β* with the representative particle size  D  is observed

3.5. Analysis of the Effect of the Aggregate Type on the Unified Coefficients α* and β*

In line with the investigations of Engelund [22], Stephenson [29], Gupta [26], and Martins [10], we group the experimental data into three different types of coarse granular porous media, smooth spheres, rolled aggregates, and crushed aggregates, whose data are shown in Appendix A. To perform this analysis, we have used as a tool diagrams of the [ D ,α*] and [ D , β*] type.
We have considered a minimum representative size  D  of 0.50 mm as a coarse porous media, so Darcy’s law is no longer valid (Taylor [6]), placing us in the so-called non-Darcy regimes flow.

3.5.1. Tests with Smooth Spheres

Figure 6 shows the experimental data (N = 24) represented in the [ D ,α*] diagram for the interval 0.50 mm <  D   < 30.00 mm, where an increase in coefficient α* with the representative particle size  D  is observed by means of Equation (38):
α * = 136.08 · e 0.0371 D
(R2 = 0.9689 N = 19)
The curve that best fits, with the experimental data used, for the linear dimensionless coefficient α* is an exponential type.
Data for  D  = 0.77 mm (Lindquist (38)),  D  = 3.00 mm (Sunada [39]),  D  = 5.03 mm (Crawford et al. [9]), and   D  = 16.00 mm (Kirkham [40]) are close to their theoretical values using Equation (38).  D  = 3.20 mm (Blake [36]) is higher than its theoretical value.
Figure 7 shows the experimental data (N = 22) plotted on the [ D , β*] diagram for the unified quadratic dimensionless coefficient β* at the 0.50 mm <  D   < 30.00 mm interval. A decrease in coefficient β* with the representative particle size  D  is observed by means of Equation (39):
β * = 2.087 · D 0.159
(R2 = 0.8698 N = 17)
As can be seen in Figure 7, the data corresponding to the porous media tested by Blake [36] and Sunada [39] are the worst fit to the curve obtained.

3.5.2. Tests with Rolled Aggregates

In granular porous media composed of rolling and crushed aggregates, unlike smooth spheres, the geometrical properties can vary considerably, which makes it difficult to ensure that the experimental data used can be adequately grouped into porous media with similar geometrical properties. However, we have grouped the data by authors on the assumption that they obtained the materials from the same geological origin, since in this case there should be less dispersion in the parameters defining the geometry of the porous media.
In the following, we will obtain the [ D , α*] diagrams for different geometries of the porous media composed of aggregates from different origins.
Figure 8 shows the [ D , α*] diagram with the experimental data (N = 28) within the range 0.50 mm <  D  < 60.00 mm. The theoretical curve defined by Equation (38) for smooth spheres is represented by a yellow dotted line. An increase in coefficient α* is observed with the representative particle size  D . Over half of the experimental data could be grouped in the size range 0.50 mm <  D   < 30.00 mm, obtaining a fit given by Equation (40).
α * = 183.89 · e 0.523 D
(R2 = 0.950 N = 15)
The data corresponding to M-B Salahi et al. [23] with a smaller interval could be fitted by Equation (41).
α * = 158.1 · e 0.1246 D
(R2 = 0.9982 N = 6)
Both Equations (40) and (41) present a greater slope in the variation of the coefficient α* than Equation (38), with corresponding smooth spheres.
The values of coefficient α* corresponding to Arbhabhirama and Dinoy [31] for  D  = 12.00 mm are above the curve given by Equation (40). The α* values for  D  = 2.29 mm and  D  = 5.97 mm corresponding to Dudgeon are much higher than the other values because of two factors: the difference in particle angularity and shape compared to the rest of the porous media ( D   > 6.00 mm) and the high uniformity coefficient values ( C u = 3.28   m m  and  C u = 3.25   m m , respectively). The α* values corresponding to Dudgeon [28] for sizes  D  = 54.86 mm and  D  = 109.73 mm (the latter not shown in Figure 8) are below Equation (37) for smooth spheres. This may be due to the need to have a higher density of points in the low gradient region to obtain a good fit of parameter  r  of Equation (1) and, therefore, according to Equation (22), of the linear dimensionless coefficient α*.
In the case of the dimensionless coefficient β*, Figure 9 shows the experimental data (N = 25) obtained for different geometries of the porous media in the interval 0.50 mm <  D   < 110.00 mm.
It is striking that the dispersion that appears in the results obtained with the experimental data provided by M-B Salahi et al. [23], of which three are aligned with the Dudgeon data ( D  = 2.10 mm;  D  = 10.35 mm  D  = 17.78 mm), while two are of the intermediate values ( D  = 4.17 mm;  D  = 6.53 mm), are considerably separated from this alignment. This greater dispersion in the geometries could be due, among other factors, to the geological origin of the materials tested.
The rest of the experimental data could be grouped together to obtain a fit given by Equation (42):
β * = 2.5966 · D 0.258
(R2 = 0.9464 N = 10)
Equation (42) includes the experimental data of Martins [11] for rolled aggregates (see Table 1). This equation is close to the theoretical curve for smooth spheres, Equation (38), which is represented by a yellow dotted line, this result is consistent since the value proposed by Martins for the dimensionless coefficient  F  is  F  = 1.05 (López et al. [3]).
Data from Ferdos et al. [33] for  D  = 130.00 mm and  D  = 200.00 mm have been obtained in the fully developed turbulent regime with a Reynolds number  R p  interval of 3.648 <  R p  < 187.448. A decrease in the unified dimensionless coefficient β* with representative particle size is observed. However, the values of the quadratic dimensionless coefficient β* are even higher than the data obtained by Dudgeon [28].

3.5.3. Tests with Crushed Aggregates

Figure 10 shows the data (N = 20) corresponding to crushed aggregate in the interval 1.50 mm <  D   < 40.00 mm.
It has been possible to pool the experimental data corresponding to Dudgeon [28], Arbhabhirama and Dinoy [31] and M-B Salahi et al. [23], obtaining a fit given by Equation (43):
α * = 285.683 · e 0.0426 · D
(R2 = 0.890 N = 11)
The data corresponding to  D  = 16.61 (M-B Salahi et al. [23]) give a higher value of the dimensionless unified coefficient α*.
The experimental data obtained by Arbhabhirama and Dinoy [31] for angular gravels are the ones that give a much higher value of the unified dimensionless coefficient  α * .
Finally, Figure 11 represents experimental data grouped into three different geometries (Equations (44)–(46)):
(a)
Dudgeon’s data [28]:
β * = 8.7408 · D 0.252
(R2 = 0.7615; N = 5)
(b)
Arbhabhirama and Dinoy [31] and M-B Salahi et al.’s data [23]:
β * = 6.7305 · D 0.245
(R2 = 0.8111; N = 9)
(c)
Martins’ data [11]:
β * = 5.0693 · D 0.274
(R2 = 0.9429; N = 5)
Figure 11. Relationship between  β *  and  D  for angular particles; 3.00 mm <  D  < 130.00 mm. Data from Dudgeon [28]; Arbhabhirama and Dinoy [31]; Martins [11] and Salahi et al. [23].
Figure 11. Relationship between  β *  and  D  for angular particles; 3.00 mm <  D  < 130.00 mm. Data from Dudgeon [28]; Arbhabhirama and Dinoy [31]; Martins [11] and Salahi et al. [23].
Water 15 03578 g011
We have not considered the experimental data,  D  = 37.19 mm, from Dudgeon because of the high dispersion it generates ( β *  = 9.62).
In the case of Martins [11], the data that produce the greatest dispersion are those corresponding to  D  = 22.00 mm and  D  = 44.00 mm with values closer to those obtained for smooth spheres (Equation (38)).
Figure 11 also shows the data corresponding to the tests carried out by Ferdos et al. [33] with properly selected uniform granular materials composed of crushed aggregates with sizes  D  = 130.00 mm and  D  = 200.00 mm. As in the case of the rolled aggregates, the values of the unified quadratic dimensionless coefficient  β *  are once again higher than the other values obtained, which seems to indicate that there is an additional increase in energy loss in these granular porous media of large sizes and subjected to strong gradients. In these tests, the pore Reynolds number range was 3.250 <  R p    < 147.484.

4. Experimental Research

To properly verify the USEC represented by Equations (19), (26), and (29), we have proceeded to:
(a)
Obtain crushed aggregates from the same geological origin (limestone quarry in this case) with the intention of reducing the dispersion in the geometric properties of the selected porous media.
(b)
Select four sizes relatively close to each other ( D  = 1.00 mm;  D  = 2.00 mm;  D  = 3.50 mm;  D  = 4.00 mm) in order to analyze the relationship of the unified dimensionless coefficients  α *  and  β *  with the representative particle size  D .
(c)
Test at least 10 different pressure gradients for each porous material tested in order to obtain a good fit of the  r  and  s  parameters from quadratic Equation (1).
d)
Represent in a [ R p , f p ] diagram the results obtained to verify that the experimental data corresponding to each porous material tested fall on the theoretical curve defined by Equation (26) for the whole range of non-Darcy flows tested.
(e)
Apply Equations (29), (32), and (33) to obtain the unified dimensionless coefficients  α *  and  β *  for each of the four porous media.
(f)
Represent in the [ D , α * ] and [ D , β * ] diagrams the obtained values of the  α *  and  β *  unified dimensionless coefficients for the four granular materials tested with the finality of observing their evolution with the representative size of particles  D .

4.1. Description of Installation

Figure 12 shows a schematic of the installation consisting of three cylindrical pipes with an inner diameter  D x  of 388 mm and a length  L x  of 2 m. With it and given that the maximum uniform materials tested have a representative size  D  of 45 mm, the ratio  D x / D  reached a value of 8.6:1. On the other hand, the ratio between the length of the permeameter  L x  and its diameter  D x  was 5.15:1. The sample was placed to complete fill the intermediate stainless-steel tube. The water, coming from a lower reservoir, is pumped through a perforated conduit that ensures that it reaches the sample without turbulence.
Seven digital pressure readings are available for the average hydraulic pressure: one in the water inlet zone, five along the sample tube, and another one in the water outlet zone, as shown in Figure 13.
The first in-sample pressure reading is at 47.3 cm from the origin of the distance; the last in-sample pressure reading is at 204.3 cm.
To ensure the reliability of the measurements, the pressure recorded is the average of the pressures at three equidistant points of the cross-section where the measurement is taken (see Figure 14).
An overview of the elements that make up the installation is shown in Figure 15 and Figure 16.

4.2. Test Procedure

The tests were carried out at flow rates  Q  between 1 ls−1 and 10 ls−1. The measurements were taken after steady flow was reached. The experimental results were represented in a [x, y] diagram where the x-axis represents the distance from the “zero” point of the different pressure taps measured in cm, and the y-axis represents the water pressure head also measured in cm (see Figure 17, Figure 18, Figure 19 and Figure 20 (origin of distances in Figure 13)). The hydraulic gradient  i  is determined by the slope of the regression line duly adjusted for each flow rate  Q  (see Figure 17, Figure 18, Figure 19 and Figure 20). In some tests, it was not possible to obtain results for the whole range of flow rates.

4.3. Materials

The material used in the tests came from crushing in the same limestone quarry in order to reduce the dispersion in the geometry of these materials with the same geological origin.
  • Representative particle size D
The representative particle size was used as the equivalent diameter ( D e ), which corresponds to the diameter of a sphere having the same volume as the particle. Since the materials came from the same quarry and were all the same type of limestone, we proceed to calculate the specific weight of the material δ by weighing each particle with sufficient accuracy to calculate the volume of the material and, from this, the diameter of the equivalent sphere  D e  by means of Equation (47):
V e = G δ
where   V e  = volume of the equivalent sphere (cm3),   G =  weight of the particle, and   δ  = specific gravity of limestone rock.
The volume of the equivalent sphere  V e  is given by Equation (48):
V e = 4 3 · π · D e 2 3
This way, we get the equivalent diameter of each particle in Equation (49):
D e = 6 · V e π 3
We finally obtain Equation (50):
D e = δ 3 · 6 · G π 3
From the different values of the particle weights obtained, the 50% percentile of the equivalent diameter  D e  was easily calculated.
Four porous media with uniform particle size and representative diameters  D e  of 10.00 mm, 20.00 mm, 35.00 mm, and 45.00 mm have been selected.
In order to obtain experimentally the value of the porosity  n  in each of the test materials, a tank with a fixed and determined volume ( V d ), uncovered at the top, was used. First, the tank was filled with the granulometric material, and then, the volume of water entering the tank, which is the total volume of the voids ( V h ), was measured.
The porosity  n  was calculated by the expression (Equation (51)):
n = V h V d · 100
Porosities 40.74%, 39.12%, 41.33%, and 43.40% were obtained for sizes 10.00 mm, 20.00 mm, 35.00 mm, and 45.00 mm, respectively.

4.4. Tests Results and Discussion

Pressure loss graphs based on the [x, y] diagram for each of the four granular porous media tested are shown in Figure 17, Figure 18, Figure 19 and Figure 20. In each of the graphs, a good correlation is observed, except for the lowest flow rates ( Q  = 2 ls−1 Q  = 31 s−1 Q  = 4 ls−1) corresponding to the largest granular porous media,  D  = 45.00 mm (see Table 2).
Table 2 shows the results obtained for the mean seepage velocity  V  and the hydraulic gradient  i  for each of the materials tested, which correspond to the slopes of the regression lines shown in Figure 17, Figure 18, Figure 19 and Figure 20. High correlation coefficients are observed in all of them except the three lowest gradients  i  corresponding to a size of 45.00 mm, so they have not been considered for the analysis.
  • Representation of the results in a [Rp, fp] diagram
In order to verify that these experimental results fit the general unified Equation (26) in a [ R p , f p ] diagram form, we have plotted the data that have been calculated by applying Equations (24) and (25) in Figure 21. The range of the Reynolds number  R p  obtained in the tests was 76.12 <   R p   < 3586.26.
As we can see in Figure 21, the experimental data corresponding to the granular porous materials  D  = 10.00 mm,  D  = 20.00 mm, and  D  = 35.00 mm conform to a smooth curve in the [ R p , f p ] diagram. For size  D  = 45.00 mm and after discarding the tests with low correlation coefficients (see Table 2), it is observed that all values (N = 6) are in the fully developed turbulent regime. Figure 21 shows that Equation (26) in a [ R p , f p ] diagram form and consequently Equation (29) in linearized form can apply to all the non-Darcy flows in which the experiments were carried out (76.12 <   R p   < 3586.26).
On the other hand, all the points (N = 32) fall on a narrow band, which seems to indicate that there is a similarity in the geometries of the four porous media as a consequence of coming from the same geological origin.
Figure 22 shows the same values represented in a [ R p , λ p ] diagram.
  • Obtaining coefficients α * and β *
Figure 23, Figure 24, Figure 25 and Figure 26 show the experimental data [ V , i V ] for each of the four granular porous media tested, where a good fit is observed in all of them. With the values  r  and  s  I Forchheimer’s equation and applying Equations (22) and (23), we obtain the values of the unified dimensionless coefficients  α *  and  β *  that are represented in Figure 27 and Figure 28.
  • Unified linear dimensionless coefficient α*. [D,α*] diagram
Figure 27 shows the experimental data represented in the [ D , α * ] diagram, where we can observe that:
(a)
There is an increase in the unified dimensionless coefficient  α *  when going from the porous media  D  = 10.00 mm to  D  = 20.00 mm.
(b)
For the porous media  D  = 45.00 mm, the reduced value of  α *  = 9.58 indicates that we are in the fully developed turbulent regime.
(c)
The value of the coefficient  α *  = 304.05 should be a matter for discussion. A possible explanation lies in the fact that in 6 of the 9 points the quadratic component of Equation (1) represents more than 90% of the total energy loss, so these 6 points are close to the fully developed turbulent flow, and consequently, the evolution of the dimensionless coefficient  α *  is decreasing as we approach the fully developed turbulent flow.
We can therefore conclude that no clear conclusions can be drawn regarding the unified dimensionless coefficient  α * .
However, as shown in Section 3.5 for non-Darcy flow regimes not close to the fully developed turbulent regime, the linear dimensionless coefficient  α *  varies with the general expression:
α * = A · e p
where parameters  A  and  p  are related to the geometry of the granular porous media and shall be adjusted for each porous media characterized by a given geometry.
  • Unified quadratic dimensionless coefficient β*. [D, β*] diagram
Figure 28 shows the data for the quadratic dimensionless coefficient  β * . In this case, a good fit was obtained as given by Equation (53):
β * = 3.0653 · D 0.258
(R2 = 0.9764)
This equation is of the same type as those obtained in Section 3.5, which shows that the relation of the dimensionless coefficient  β *  to the representative particle size  D   is given by the general expression of the potential type:
β * = B · D q
where parameters  β  and q are related to the geometry of the granular porous media and shall be adjusted for each porous media characterized by a given geometry.

5. Conclusions

In this research work, a general unified equation for seepage in porous media has been developed which, based on the adoption of the mean hydraulic diameter  D h  as a characteristic length  L c ,  is applicable to coarse granular media with uniform granulometry for non-Darcy flows and can be represented in a quadratic form (Equation (18)), in a [ R p , f p ]  diagram form (Equation (26)), and in a linearized form [ R p , λ p ]  (Equation (28)).
The linearized equation [ R p , λ p ]  (Equation (28)) allows us to check whether, for a given granular porous media, there is a smooth curve in a [ R p , f p ]  diagram covering the different non-Darcy flow regimes: nonlinear laminar, turbulent transition, and fully developed turbulent.
Based on experimental data from different authors (Appendix A), it has been verified that the unified dimensionless coefficients linear  α *  and quadratic  β * , which are considered by the general unified equation, are a function of the geometry of the porous material through parameter  A  and exponent  p  for the linear unified dimensionless coefficient  α * and parameter  B  and exponent  q  for the quadratic unified dimensionless coefficient  β *; and of the representative diameter of the particles  D  through Equations (52) and (54). The linear  r  and quadratic  s  parameters of the Forchheimer equation (Equation (1)) are given by the following expressions:
r = v g · A · e p · D · ( 1 n ) 2 n 3 · 1 D
s = 1 g · B · D q · ( 1 n ) n 3 · 1 D 2
where  p  and  q  are exponents.
Granular porous media with similar geometrical properties will be defined by parameters  A ,   B ,   p , and  q . For the case of spherical particles and size range 0.50 mm <  D  < 30.00 mm, a good fit has been obtained with the values  A  = 136.08,  p  = 00371 (R2 = 0.9689),  B  = 2.087, and  q  = 0.159 (R2 = 0.8698). For the rest of the porous media, rolled aggregates and crushed aggregates, parameters  A  and   B  have varied according to their geometric properties in accordance with the results obtained in Section 3.5.2 and Section 3.5.3, respectively.
An application of the general unified equation has been carried out by means of tests at CEDEX on porous media duly selected in terms of size, uniformity, and origin from the same geological origin (limestone quarry) under the following principles:
(a)
Verification of the existence of smooth curves in the [ R p , f p ]  diagram for all the sizes tested, adjusting to the general unified Equation (26).
(b)
Application of the linearized unified general Equation (29) to obtain the dimensionless linear  α *  and quadratic  β *  coefficients.
(c)
Verification of the asymptote for the fully developed turbulent flow reached with the largest size, D = 45.00 mm
(d)
No clear conclusions have been drawn for the linear unified dimensionless coefficient  α * .
(e)
For the quadratic unified dimensionless coefficient  β * , a good fit is observed (R2 = 0.9764) with values  B  = 3.0653 and  q  = 0.258.
Further research work is needed to analyze the influence of the geological origin on the geometrical properties of both parameters  a  and  b  and exponents  p  and  q , as well as to analyze the evolution of the linear unified coefficient  α *  as we approach the fully developed turbulent regime.
Equation USEC can be applied to various fields in science and engineering where coarse porous media are used, such as rockfill dams, breakwaters in coastal engineering, ripraps, embankments, etc.
Of particular importance is its application in the study of the saturation process of the downstream spall of a rockfill dam subjected to overtopping.

Author Contributions

Conceptualization: J.C.L.; formal analysis: J.C.L.; investigation: J.C.L.; resources: J.C.L.; data curation: J.C.L.; writing—original draft preparation: J.C.L. and R.M.; writing—review and editing: M.Á.T., L.B. and R.M.; visualization: J.C.L.; supervision: M.Á.T., L.B. and R.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research work has been funded by the Spanish Ministry of Science and Innovation within the XPRES Research Project (grant number BIA2007-68120-C03-02).

Data Availability Statement

Data used in this report come from the mentioned references.

Acknowledgments

We would like to thank the staff of CEDEX (Centro de Estudios y Experimentación de Obras Públicas) and SERPA (Dam Safety Research Group) for the support in the tests for the development of this research work.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

  α Coefficient of the exponential equation that depends on the characteristics of the porous media and of the fluid
  A USEC linear parameter depending on the geometry of the porous material
  A 1 Generalized dimensionless coefficient of the linear expression r
  A 2 Generalized dimensionless coefficient of the quadratic expression s
  b Exponent of the exponential equation function of the flow conditions
  B USEC quadratic parameter depending on the geometry of the porous material.
  C Particle shape coefficient as set of R. D. Gupta
  D Representative size of the particles in uniform materials
  D a Average consecutive sieve aperture
  D e Equivalent diameter or diameter of a sphere with the same volume as the particle
  D h Hydraulic mean diameter
  D t Pipe diameter
  D x Permeameter diameter of an installation
  F Dimensionless coefficient that considers shape, angularity, and roughness of particles
  G Relative specific weight of solid particles
  f Generalized friction factor by Darcy–Weisbach
  f ( n ) Porosity function of hydraulic mean diameter
  f L ( n ) Porosity function of the linear term of the USEC
  f T ( n ) Porosity function of the quadratic term of the USEC
  f p Pore friction factor based on  D h  
  f r Pipe turbulent friction factor
  g Gravitational acceleration
  i Hydraulic gradient
  K 0 Intrinsic permeability of the porous media
  L c Generalized characteristic length of porous media
  L x Length of the permeameter
  η Angularity as a set of aggregates defined by R. D. Gupta
  N Number of tests performed
  n Porosity
  n R. D. Gupta exponent for the formula that defines the angularity η
  p USEC linear exponent depending on the geometry of the granular porous media
  q USEC quadratic exponent depending on the geometry of the granular porous media
  Q Mean permeameter flow rate (l/s)
  r Linear coefficient of the Forchheimer equation of function of the characteristic of the porous media and fluid
  R Correlation coefficient
  R e Generalized Reynolds number
  R e t Pipe Reynolds number
  R h Hydraulic mean radius
  R p Pore Reynolds number based on  D h  
  s Quadratic coefficient of the Forchheimer equation of function of the characteristic of the porous media
  S e Average specific surface area of solid particles
  t Independent term of the USEC linearized equation
  u Linear term of the USEC linearized equation
  v Kinematic viscosity
  V Average seepage velocity
  V d Fixed tank volume used for porosity calculation.
  V e Volume of equivalent sphere of one particle
  V h Total void volume of the reservoir for the porosity calculation
  V p Pore velocity
  x Horizontal distance in cm to obtain pressure loss
  y Measured pressure in cm to obtain pressure loss
α *USEC unified linear dimensionless coefficient
β *USEC unified quadratic dimensionless coefficient
  ɛ Particle surface roughness
  ε t Pipe absolute roughness
  δ Specific weight of limestone rock
  λ pLinearized USEC pore friction factor

Appendix A

Sabri Ergun and A.A. Orning [13] obtained the values of,  α n , and  β  from the tests of Lindquist [38] and Burke and Plummer [37] for particles with spherical shapes. The size range was 0.23 mm <  D  < 6.30 mm (Table A1).
Ward conducted 53 tests on 20 different porous media including glass beads, sands, gravel, and anthracite in a range of 0.27 mm <  D   < 16.1 mm particles. Data are available for r K 0 D , and porosity  n . The kinematic viscosity  v  ranged from 0.85 × 10−8·m2·s−1 v   < 1.39·10−6·m2·s−1. There are no data of this value for each of the 53 trials Table A1).
Dudgeon [28] conducted experimental research on coarse porous media to cover the widest possible range of flow conditions. The granular materials used were river gravel, crushed aggregates (Dolerite), and smooth balls. The advantage of this research work lies in having carried out the tests on a wide range of sizes and gradients, which allows the analyzing of the validity of Equation (6) for all non-Darcy flow regimes.
The range of applied sizes was 2.29 mm <  D  < 109.73 mm, and the range of hydraulic gradients was 5.40·10−5 i   < 13.4. The permeameter used had a diameter of  D x  571 mm and a length of  L x  1219.20 mm. The author represented the data obtained in the tests in a diagram  L o g i L o g V  to detect the slope changes that define the successive “post-linear regimes” or non-Darcy regime for each porous material. First, it obtained the critical speeds and gradients that correspond to the slope changes in the graphs. Second, for each of these lines, he obtained the values   a  and  b  of the exponential Equation (2). In non-Darcy regimes, the exponent interval  b  was in the range 1.20 <  b   < 1.91. Kinematic viscosity  v  data are available in all tests (Table A1).
Ahmed and Sunada [15] presented values of  r ,   s ,   D ,   a n d   K 0  of several authors, Ahmed, N. [41], Lindquist [38], and Francher and Lewis [42] with sands as porous media, Blake [36], Sunada D.K. [39], and Brownell and Katz [43] with glass beads, and finally, Kirkham [40] with smooth spheres as porous media. No porosity  n  data are available. The kinematic viscosity is calculated by applying Equation (8) since we know the values of  K 0  and  r  (Table A1).
Arbhabhirama and Dinoy [31] conducted tests on granular materials with a size range of 1.6 mm <  D   < 28.3 mm. The porous materials used were sands with  D  = 1.6 mm, angular gravels with  D  = 6.4 mm,  D  = 13.00 mm, and  D  = 28.3 mm, and rounded gravels from a river with a single size:  D  = 12.00 mm. They used two permeameters with  D x  = 140.8 mm and  D x  = 63.1 mm. The values  r ,   s ,   K 0 D ,  and porosity  n  are available (Table A3).
McDonald I.F. et al. [44] and Tyagi and Tood [45] directly provided the r and s values of the experimental data performed by Dudgeon [28]. These authors adjusted a single value of  r  and  s  for each porous media considering the entire interval of gradients tested, which means that they consider valid Equation (6), which defines a smooth curve for all non-Darcy flow regimes (Table A3).
Martins [11], in his research work, worked with two types of granular materials: rounded river gravels with size range 22.0 mm < 89.0 mm and crushed aggregates within the range 11 mm <  D   < 127 mm. The tests were performed with a diameter permeameter of  D x  500 mm and a length of  L x 1000 mm. Hydraulic gradient  i , pore speed  V p , and porosity  n  values are available for each granular material. The author’s purpose was to propose an exponential equation with coefficient b equal to 2 working with gradients to be in the fully developed turbulent regime. The range of gradients was in the range of 0.099 <  i   < 1.069 with a dot density ( v , i ) for  N = 8  of each porous material (Table A2 and Table A3).
Sedghi-Asl et al. [32] conducted experiments on 6 samples of alluvial road aggregates, with a size range of 2.83 mm <  D  < 56.8 mm. The range of gradients was in the range 0.008 <   i   < 0.914. The perimeter used had a diameter of  D x  300 mm and a length of  L x  1115 mm. A good density of points ( v , i N = 17  is available. With this, the authors made an adjustment for each porous material, obtaining the values of  r ,   s ,   a ,  and  b . The exponent  b  stood at a range of 1479 <   b   < 1804 (Table A2).
Salahi M-B et al. [23], whose research object was to describe the behavior of flow in coarse granular materials in the transition regime (non-Darcy flow), conducted a series of experiments on two different types of granular materials: 6 sizes of rolled aggregates with a size range of 2.1 mm <  D   < 17.78 mm and 6 sizes of crushed aggregates with 1.77 mm <  D   < 16.62 mm. The range of gradients was in the range 0.1 <   i  < 0.7. The permeameter had a diameter of  D x  200 mm and a length of  L x  1000 mm. They made the corresponding adjustment for each porous material ( N = 10  points), obtaining the values  r  and  s  (Table A2 and Table A3).
Finally, Ferdos F. et al. [33], to adequately model the mechanisms of rupture in breakwater dams, conducted a series of experiments on this type of material at high Reynolds numbers,  R d , to safely place themselves in the fully developed turbulent regime. They used two types of materials that were carefully sifted to assign low uniformity coefficients  C u . They used two types of materials: pebbles and crushed rock. For each of them, two size ranges were obtained: 100 mm <  D   < 160 mm and 160 mm <  D   < 240 mm with D50 = 130 mm and D50 = 200 mm, respectively. The permeameter is composed of three bodies: the main unit containing the porous material has a diameter of  D x   1000 mm and a length of  L x  1990 mm, and the input units ( L x  = 1500 mm) and output ( L x  = 500 mm) decrease their section linearly to a final diameter of  D x  = 276 mm in both cases (Table A2 and Table A3).
Table A1. SPHERES 0.20 mm < D < 30.00 mm.
Table A1. SPHERES 0.20 mm < D < 30.00 mm.
Data ByReferenceMedium TypeParticle Size D(m)Porosity nν (m2:s−1)r (s2.m−2)s (s2.m−2)Ko (m2)αβFα*β*Rp
Lindquist (1933)S.Ergun & A.A.Orning [13]Glass beads0.0002270.343**0.00*1.903.001.00136.802.25**
Ward (1964)Ward [30]Glass beads0.0002730.370***5.77 × 10−11**1.00164.84***
Ward (1964)Ward [30]Glass beads0.0003220.341***6.37 × 10−11**1.00148.62***
Ward (1964)Ward [30]Glass beads0.0003220.355***8.36 × 10−11**1.00133.37***
Ward (1964)Ward [30]Glass beads0.0003220.370***9.01 × 10−11**1.00146.86***
Ward (1964)Ward [30]Glass beads0.0003830.370***1.15 × 10−10**1.00162.79***
Ward (1964)Ward [30]Glass beads0.0004580.370***1.87 × 10−10**1.00143.16***
Lindquist (1933)S. Ergun & A. A. Orning [13]Iron shot0.0004720.375****2.002.501.00144.001.88**
Lindquist (1933)S. Ergun & A. A. Orning [13]Iron shot0.0004720.375****2.002.601.00144.001.95**
Lindquist (1933)S. Ergun & A. A. Orning [13]Iron shot0.0004720.375****1.802.501.00129.601.88**
Lindquist (1933)S. Ergun & A. A. Orning [13]Lead shot0.0004970.350****1.902.701.00136.802.03**
Lindquist (1933)S. Ergun & A. A. Orning [13]Lead shot0.0004970.350****1.803.101.00129.602.33**
Ward (1964)Ward [30]Glass beads0.0005450.370***2.62 × 10−10**1.00144.68***
Lindquist (1933)S. Ergun & A. A. Orning [13]Lead shot0.0005620.350****1.802.701.00129.602.03**
Lindquist (1933)S. Ergun & A. A. Orning [13]Lead shot0.0005620.352****1.803.101.00129.602.33**
Lindquist (1933)S. Ergun & A. A. Orning [13]Glass beads0.0005700.330****1.803.101.00129.602.33**
Lindquist (1933)S. Ergun & A. A. Orning [13]Glass beads0.0005700.330****1.902.801.00136.802.10**
Ward (1964)Ward [30]Glass beads0.0006500.370***3.47 × 10−10**1.00155.39***
Lindquist (1933)S. Ergun & A. A. Orning [13]Iron shot0.0007690.371****2.002.701.00144.002.03
Lindquist (1933)S. Ergun & A. A. Orning [13]Lead shot0.0007980.364****1.903.301.00136.802.48**
Lindquist (1933)S. Ergun & A. A. Orning [13]Lead shot0.0007980.364****1.903.301.00136.802.48**
Lindquist (1933)S. Ergun & A.A. Orning [13]Lead shot0.0010040.366****1.802.901.00129.602.18**
Burke and Plummer (1928)S. Ergun & A. A. Orning [13]Lead shot0.0014780.375****2.002.601.00144.001.95**
Sunada (1965)Ahmed & Sunada [15]Glass spheres0.0030000.3609.17 × 10−714.50648.006.45 × 10−9**1.00158.941.39**
Burke and Plummer (1928)S. Ergun & A. A. Orning [13]Lead shot0.0030770.383****2.302.501.00165.601.88**
Burke and Plummer (1928)S. Ergun & A. A. Orning [13]Lead shot0.0030770.390****2.202.501.00158.401.88**
Blake (1922)Ahmed & Sunada [15]Glass beads0.0032000.3609.79 × 10−714.90623.006.70 × 10−9**1.00174.091.43**
Crawford C.W et al. (1986)Crawford C.W et al. [9]Glass sheres 0.0050300.356******1.00166.101.50**
Burke & Plummer (1928)S. Ergun & A. A. Orning [13]Lead shot0.0062700.421****2.802.001.00201.601.50
Burke & Plummer (1928)S. Ergun & A. A. Orning [13]Lead shot0.0062700.393****2.802.101.00201.601.58
Kirkham (1966)Ahmed & Sunada [15]Marble0.0160000.3601.04 × 10−60.90117.001.19 × 10−7**1.00245.041.34
Dudgeon (1966)McDonald I.F, et al. [44]Marbles0.015970.3691.28 × 10−61.1000103.00001.19 × 10−7**1.00232.011.263.213590.05
Dudgeon (1966)McDonald I.F, et al. [44]Marbles0.015970.4151.21 × 10−60.500063.00002.46 × 10−7**1.00232.011.265.332342.52
Dudgeon (1966)McDonald I.F, et al. [44]Marbles0.015970.3721.15 × 10−60.760095.00001.55 × 10−7**1.00232.011.2629.162018.50
Dudgeon (1966)McDonald I.F, et al. [44]Marbles0.024870.3691.27 × 10−60.580066.00002.24 × 10−7**1.00348.911.286.296669.85
Dudgeon (1966)McDonald I.F, et al. [44]Marbles0.028960.3851.11 × 10−60.360049.00003.16 × 10−7**1.00400.871.2933.479699.04
Note: (*) No data available.
Table A2. Rounded materials. 0.50 mm < D < 200.00 mm.
Table A2. Rounded materials. 0.50 mm < D < 200.00 mm.
Data ByReferenceMedium TypeParticle Size D (m)Porosity nν (m2.s−1)r (s2.m−1)s (s2.m−2)Ko (m2)α*β*Rp
Sedghi-Asl M et al. (2013)Sedghi-Asl M et al. [32]Rounded alluvial materials0.002830.32009.12 × 10−716.242887.2005.72 × 10−999.161.194.46213.90
Sedghi-Asl M et al. (2013)Sedghi-Asl M et al. [32]Rounded alluvial materials0.005500.33008.80 × 10−75.906508.8301.52 × 10−8159.511.4728.04651.29
Sedghi-Asl M et al. (2013)Sedghi-Asl M et al. [32]Rounded alluvial materials0.008700.35008.72 × 10−73.622274.8802.46 × 10−8312.781.5555.941435.90
Sedghi-Asl M et al. (2013)Sedghi-Asl M et al. [32]Rounded alluvial materials0.015600.32008.71 × 10−71.692170.5905.25 × 10−8328.781.26134.583493.76
Sedghi-Asl M et al. (2013)Sedghi-Asl M et al. [32]Rounded alluvial materials0.031100.36008.78 × 10−71.30147.0466.88 × 10−81601.531.05435.6111,837.12
Sedghi-Asl M et al. (2013)Sedghi-Asl M et al. [32]Rounded alluvial materials0.056800.40008.77 × 10−70.53522.3671.67*10−73433.401.332424.2431,060.61
M-B Salahi et al. (2015)M-B Salahi et al. [23]Rounded materials0.002100.37501.32 × 10−642.7823548.8003.15 × 10−9189.296.1712.6134.91
M-B Salahi et al. (2015)M-B Salahi et al. [23]Rounded materials0.004170.37901.33 × 10−615.678854.8508.62 × 10−9284.723.0773.20162.01
M-B Salahi et al. (2015)M-B Salahi et al. [23]Rounded materials0.006530.38301.32 × 10−67.712501.4901.74 × 10−8360.692.93152.25376.21
M-B Salahi et al. (2015)M-B Salahi et al. [23]Rounded materials0.010350.38641.32 × 10−64.924440.7002.73 × 10−8602.164.21286.44703.25
M-B Salahi et al. (2015)M-B Salahi et al. [23]Rounded materials0.012130.39781.32 × 10−63.658242.7003.68 × 10−8694.303.02415.651081.89
M-B Salahi et al. (2015)M-B Salahi et al. [23]Rounded materials0.017780.40711.28 × 10−63.161158.5804.14 × 10−81465.413.15729.851923.77
Dudgeon (1966)McDonald I.F, et al. [44]River Gravel0.002290.4181.30 × 10−678.91002232.00001.68 × 10−9672.796.290.07111.74
Dudgeon (1966)McDonald I.F, et al. [44]River Gravel0.005790.3921.32 × 10−619.04002174.00007.06 × 10−9773.5712.230.23278.75
Dudgeon (1966)McDonald I.F, et al. [44]Rivel Gravel0.015850.3671.30 × 10−61.8900262.00007.02 × 10−8441.683.184.272239.85
Dudgeon (1966)McDonald I.F, et al. [44]River Gravel0.025950.3721.30 × 10−60.8200145.00001.62 × 10−7543.493.038.215016.46
Dudgeon (1966)McDonald I.F, et al. [44]River Gravel0.054860.3691.27 × 10−60.240051.00005.41 × 10−7702.402.19113.5512,868.54
Dudgeon (1966)McDonald I.F, et al. [44]River Gravel0.109730.4061.27 × 10−60.060015.00002.16 × 10−61055.961.82156.8236,188.89
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Sand0.0016000.3991.00 × 10−685.231750.861.196 × 10−6376.432.90**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Sand0.0016000.3911.00 × 10−695.181819.071.071 × 10−6385.252.80**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Gravel Round0.0120000.3731.00 × 10−66.04207.101.6885 × 10−81125.782.02**
Arbhabhirama & Dinoy (1973)Arbhabhirama &y Dinoy [31]Gravel Round0.0120000.3571.00 × 10−65.63187.821.8117× 10−8874.701.56**
Ward (1964)Ward [30]Sand0.0006250.407***2.98 × 10−10251.31***
Ward (1964)Ward [30]Sand0.0012600.400***1.36 × 10−9207.53***
Ward (1964)Ward [30]Gravel0.0018000.410***2.98 × 10−9215.27***
Ward (1964)Ward [30]Gravel0.0050400.389***1.69 × 10−8237.00***
Ward (1964)Ward [30]Gravel0.0092100.417***5.26 × 10−8344.04***
Ward (1964)Ward [30]Gravel0.0161000.422***1.80 × 10−7323.94***
Lindquist (1933)Ahmed & Sunada [15]Sand0.0010500.3809.14 × 10−7116.402920.008.00 × 10−10196.722.66
Lindquist (1933)Ahmed & Sunada [15]Sand0.0049200.3809.12 × 10−76.74368.001.38 × 10−8250.391.57
Martins (1990)Martins [11]Rounded materials0.022000.36221.15 × 10−6****1.29752.002146.00
Martins (1990)Martins [11]Rounded materials0.022000.36831.15 × 10−6****1.29751.002103.00
Martins (1990)Martins [11]Rounded materials0.044000.36181.15 × 10−6****0.902631.006970.00
Martins (1990)Martins [11]Rounded materials0.044000.35531.15 × 10−6****0.902583.007027.00
Martins (1990)Martins [11]Rounded materials0.089000.38231.15 × 10−6****0.838881.0023,359.00
Martins (1990)Martins [11]Rounded materials0.089000.38841.15 × 10−6****0.8310,765.0025,430.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Cobblestone0.130000.46801.00 × 10−63.71 × 10−1310.0792.74 × 10+50.002.489515.00114,176.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Cobblestone0.130000.46801.00 × 10−62.37 × 10−109.9774.31 × 10+20.002.459515.00114,176.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Cobblestone0.130000.44501.00 × 10−62.00 × 10−1011.6975.09 × 10+20.002.373648.00115,829.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Cobblestone0.200000.50601.00 × 10−62.04 × 10−113.9475.00 × 10+30.002.0315,621.00187,448.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Cobblestone0.200000.50601.00 × 10−61.49 × 10−114.2016.84 × 10+30.002.1615,621.00187,448.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Cobblestone0.200000.45901.00 × 10−61.96 × 10−145.8095.21 × 10−60.002.0411,411.00184,400.00
Note: (*) No data available.
Table A3. Crushed materials. 1.50 mm < D < 200.00 mm.
Table A3. Crushed materials. 1.50 mm < D < 200.00 mm.
Data ByReferenceMedium TypeParticle Size D (m)Porosity nν (m2.s−1)r (s.2m−1)s (s2.m−2)Ko (m2)α*β*Rp
M-B Salahi et al. (2015)M-B Salahi et al. [23]Crushed materials0.001770.42001.30 × 10−645.5152978.8002.91 × 10−9236.986.617.6625.29
M-B Salahi et al. (2015)M-B Salahi et al. [23]Crushed materials0.003550.42101.30 × 10−615.665979.7108.46 × 10−9331.594.4042.22109.77
M-B Salahi et al. (2015)M-B Salahi et al. [23]Crushed materials0.005550.42251.30 × 10−64.550514.3002.91 × 10−8239.143.66130.06278.60
M-B Salahi et al. (2015)M-B Salahi et al. [23]Crushed materials0.008690.42841.30 × 10−63.508361.7003.78 × 10−8481.054.24178.06471.97
M-B Salahi et al. (2015)M-B Salahi et al. [23]Crushed materials0.013080.43921.30 × 10−61.066174.4201.24 × 10−7370.883.38460.451045.33
M-B Salahi et al. (2015)M-B Salahi et al. [23]Crushed materials0.016610.46921.30 × 10−61.789114.0407.41 × 10−81365.183.62617.941544.84
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.003200.4771.31 × 10−616.61959.008.04 × 10−9505.386.250.97221.45
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.006400.4581.31 × 10−67.79573.001.71 × 10−8781.446.381.27565.36
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.014020.4281.29 × 10−61.43220.009.20 × 10−8512.214.153.091776.00
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.014020.5151.14 × 10−60.5197.002.28 × 10−7500.923.766.871374.02
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.015850.4551.16 × 10−61.15162.001.03 × 10−7774.834.351.832734.03
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.024990.4381.31 × 10−60.61117.002.19 × 10−7758.944.2922.994368.32
Dudgeon (1966)McDonald I.F. et al. [44]Blue metal0.037190.4831.29 × 10−60.33121.003.98 × 10−71463.219.6212.847100.78
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Sand0.0016000.3991.00 × 10−685.231750.861.196 × 10−9376.432.90**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Sand0.0016000.3911.00 × 10−695.181819.071.071 × 10−9385.252.80**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Gravel Angular 10.0064000.4671.00 × 10−68.80359.001.1581 × 10−81267.974.31**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Gravel Angular 10.0064000.4701.00 × 10−611.98390.078.51 × 10−91778.984.80**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Gravel Angular 20.0130000.4611.00 × 10−62.96176.913.4425 × 10−81655.534.10**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Gravel Angular 20.0130000.4791.00 × 10−62.71130.933.7585 × 10−81820.553.52**
Arbhabhirama & Dinoy (1973)Arbhabhirama & Dinoy [31]Gravel Angular 30.0283000.4651.00 × 10−61.1655.938.8255 × 10−83187.752.92**
Martins (1990)Martins [11]Angular materials0.011000.4621.15 × 10−6****2.74236.00683.00
Martins (1990)Martins [11]Angular materials0.011000.4601.15 × 10−6****2.74243.00692.00
Martins (1990)Martins [11]Angular materials0.016000.4281.15 × 10−6****2.19382.001123.00
Martins (1990)Martins [11]Angular materials0.016000.4271.15 × 10−6****2.19394.001101.00
Martins (1990)Martins [11]Angular materials0.022000.4311.15 × 10−6****1.72647.002010.00
Martins (1990)Martins [11]Angular materials0.022000.4241.15 × 10−6****1.72622.001945.00
Martins (1990)Martins [11]Angular materials0.032000.4291.15 × 10−6****2.101150.003146.00
Martins (1990)Martins [11]Angular materials0.032000.4361.15 × 10−6****2.101287.003405.00
Martins (1990)Martins [11]Angular materials0.044000.3971.15 × 10−6****1.371907.005176.00
Martins (1990)Martins [11]Angular materials0.044000.3951.15 × 10−6****1.371860.005188.00
Martins (1990)Martins [11]Angular materials0.063500.4141.15 × 10−6****1.533925.0010,509.00
Martins (1990)Martins [11]Angular materials0.063500.4441.15 × 10−6****1.534623.0012,210.00
Martins (1990)Martins [11]Angular materials0.089000.4741.15 × 10−6****2.048127.0019,793.00
Martins (1990)Martins [11]Angular materials0.089000.4701.15 × 10−6****2.047695.0019,820.00
Martins (1990)Martins [11]Angular materials0.127000.4811.15 × 10−6****1.3816,581.0043,687.00
Martins (1990)Martins [11]Angular materials0.127000.4751.15 × 10−6****1.3815,658.0041,238.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Crushed rock0.130000.50601.00 × 10−68.96 × 10−1510.8691.14 × 10+70.003.648273.0099,278.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Crushed rock0.130000.50601.00 × 10−62.08 × 10−1110.5464.91 × 10+30.003.538273.0099,278.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Crushed rock0.130000.49701.00 × 10−64.20 × 10−109.8602.43 × 10+20.003.073250.00105,626.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Crushed rock0.200000.49001.00 × 10−62.61 × 10−136.7793.90 × 10+50.003.0711,478.00137,741.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Crushed rock0.200000.49001.00 × 10−61.09 × 10−116.6519.40 × 10+30.003.0111,478.00137,741.00
Farzad Ferdos et al. (2015)Farzad Ferdos et al. [33]Crushed rock0.200000.48401.00 × 10−62.53 × 10−145.6764.04 × 10+60.002.456807.00147,484.00
Note: (*) No data available.

References

  1. Forchheimer, P.H. Wasserbewegung durch Boden. Z. Vereines Dtsch. Ingenieure 1901, 50, 1781–1788. [Google Scholar]
  2. Izbash, S.V. O Filtracii v Krupnozernstom Materiale; Nauchno Issled. Inst. Gidrotechniki (NIIG): Saint Petersburg, Russia, 1931. [Google Scholar]
  3. López, J.C.; Toledo, M.Á.; Moran, R. A unified view of nonlinear resistance formulas for seepage flow in coarse granular media. Water 2021, 13, 1967. [Google Scholar] [CrossRef]
  4. Bear, J. Dynamics of Fluid in Porous Media; American Elsevier Publishing Company Inc.: New York, NY, USA, 1972. [Google Scholar]
  5. White, F.M. Fluid Mechanics, 5th ed.; McGraw-Hill, Inc.: Boston, MA, USA, 2003. [Google Scholar]
  6. Taylor, D.W. Fundamentals of Soil Mechanics; John Wiley & Sons Inc.: New York, NY, USA, 1948. [Google Scholar]
  7. Loudon, A.G. The computation of permeability from simple soil test. Geotechnique 1952, 3, 165–183. [Google Scholar] [CrossRef]
  8. Lindford, A.; Saunders, D.H. A Hydraulic Investigation of through and Overflow Rockfill Dams; British Hydromechanics Research Associations: Cranfield, UK, 1967. [Google Scholar]
  9. Crawford, C.W.; Plumb, O.A. The influence of surface roughness on resistance to flow through packed beds. J. Fluids Eng. 1986, 108, 343–347. [Google Scholar] [CrossRef]
  10. Martins, R.; Escarameria, M. Caracterizaçao dos Materiais Para o Estudo em Laboratorio do Escoamento de Percolaçao Turbulento; Springer: Lisbon, Portugal, 1989. [Google Scholar]
  11. Martins, R. Seepage Flow through Rockfill Dams. In Proceedings of the 17th International Congress on Large Dams, Vienna, Austria, 17–21 June 1991; Volume Q67R14. [Google Scholar]
  12. Sabin, G.C.; Hansen, D. The effects of particle shape and surface roughness on the hydraulic mean radius of a porous medium consisting of quarried rock. Geotech. Test. J. 1994, 17, 43–49. [Google Scholar]
  13. Ergun, S.; Orning, A.A. Fluid flow through randomly packed columns and fluidized beds. Ind. Eng. Chem. 1949, 41, 1179–1184. [Google Scholar] [CrossRef]
  14. Huang, H.; Ayoub, J. Applicability of the Forchheimer equation for non-Darcy flow in porous media. SPE J. 2007, 13, 112–122. [Google Scholar] [CrossRef]
  15. Ahmed, N.; Sunada, D.K. Nonlinear flow in porous media. J. Hydraul. Div. 1969, 95, 1847–1858. [Google Scholar] [CrossRef]
  16. Wright, D.E. Nonlinear flow through granular media. J. Hydraul. Div. 1968, 94, 851–872. [Google Scholar] [CrossRef]
  17. Panfilov, M.; Oltean, C.; Panfilova, I.; Buès, M. Singular nature of nonlineal macroscale effects in high-rate flow through porous media. Comtes Rendus Mec. 2003, 331, 41–48. [Google Scholar] [CrossRef]
  18. Fourar, M.; Radilla, G.; Lenormand, R.; Moyne, C. On the non-linear behavior of a laminar single-phase flow through two and three dimensional porous media. Adv. Water Resour. 2004, 27, 669–677. [Google Scholar] [CrossRef]
  19. Mccorquodale, J.A.; Hannoura, A.A.A.; Sam Nasser, M. Hydraulic conductivity of rockfill. Hydraul. Res. 1978, 16, 123–127. [Google Scholar] [CrossRef]
  20. Mulqueen, J. The flow of water through gravels. Ir. J. Agrucyltural Food Res. 2005, 44, 83–94. [Google Scholar]
  21. Terzaghi, K.; Peck, R.B. Soil Mechanics in Engineering Practice; John Wiley & Sons: Hoboken, NJ, USA, 1964. [Google Scholar]
  22. Engelund, F. On the laminar and turbulent flows of ground water through homogeneous sand. Hydraul. Lab. 1953, 4, 356–361. [Google Scholar]
  23. Salahi, M.-B.; Sedghi-Als, M.; Parvizi, M. Nonlinear flow through a packed-column experiment. J. Hydrol. Eng. 2015, 20, 04015003. [Google Scholar] [CrossRef]
  24. Fand, R.M.; Kim, B.Y.K.; Lam, A.C.C.; Phan, R.T. Resistance to the flow of fluids through simple and complex porous media whose matrices are composed of randomly packed spheres. J. Fluid Eng. 1987, 109, 268–273. [Google Scholar] [CrossRef]
  25. Barree, R.D.; Conway, M.W. Reply to Discussion of “Beyond Beta factors: A complete model for Darcy. Forchheimer and trans-Forchheimer flow in porous media”. J. Pet. Technol. 2005, 57, 73. [Google Scholar] [CrossRef]
  26. Gupta, R.D. Angularity of Aggregate Particles as a Measure of Their Shape and Hydraulic Resistance; Technical note 438; Water Engineering Group, Aligarh Muslim University: Aligarh, India, 1985; pp. 705–716. [Google Scholar]
  27. Shergold, F.A. The percentage voids in compacted gravel as a measure of its angularity. Mag. Concr. Res. 1953, 5, 3–10. [Google Scholar] [CrossRef]
  28. Dudgeon, R.D. An experimental study of the flow of water through coarse granular media. Houllie Blanche 1966, 7, 785–802. [Google Scholar] [CrossRef]
  29. Stephenson, D. Flow through rockfill. Dev. Geotech. Eng. 1979, 27, 19–37. [Google Scholar] [CrossRef]
  30. Ward, J.C. Turbulent flow in porous media. J. Hydraul. Div. 1964, 90. [Google Scholar] [CrossRef]
  31. Arbhabhirama, A.; Dinoy, A.A. Friction factor and Reynolds number in porous media Flow. J. Hydraul. Div. 1973, 99, 901–911. [Google Scholar] [CrossRef]
  32. Sedghi-Als, M.; Rihimi, H.; Salehi, R. Non-Darcy Flow of Water through a Packed-Column; Springer Science: Berlin, Germany, 2013; pp. 215–227. [Google Scholar] [CrossRef]
  33. Ferdos, F.; Wörman, A.; Ekström, I. Hydraulic conductivity of coarse Rockfill used in hydraulic structures. Trans. Porous Med. 2015, 108, 367–391. [Google Scholar] [CrossRef]
  34. Parkin, A.K. Through and overflow rockfill dams. In Advances in Rockfill Structures; Kluwer Academic Publishers: Alphen aan den Rijn, The Netherlands, 1991; pp. 571–592. [Google Scholar]
  35. Colebrook, C.F. Turbulent Flow in Pipes, with Particular Reference to the Transition Region between the Smooth and Rough Pipe Laws. J. Inst. Civ. Eng. 1939, 12, 393–422. [Google Scholar] [CrossRef]
  36. Blake, F.C. The Resistance of Packing to Fluid Flow; Read at the Richmond Meeting; American Institute of Chemical Engineers: New York, NY, USA, 1922; pp. 415–421. [Google Scholar]
  37. Burke, S.P.; Plummer, W.B. Gas flow through packed columns. Ind. Eng. Chem. 1928, 20, 1196–1200. [Google Scholar] [CrossRef]
  38. Lindquist, E. Premier Congrès des Grandes Banages; Stockholm, Sweden, 1933; Volume V, pp. 81–99. Available online: https://fr.wikipedia.org/wiki/Premier_Congr%C3%A8s_universel_des_races (accessed on 9 October 2023).
  39. Sunada, D.K. Laminar and Turbulent Flow of Water through Homogeneous Porous Media. Ph.D. Thesis, University of California at Berkeley, Berkeley, CA, USA, 1969. [Google Scholar]
  40. Kirkham, C.E. Turbulent Flow in Porous Media—An Analytical and Experimental Model Study. Ph.D. Thesis, University of Melbourne, Melbourne, Australia, 1966. [Google Scholar]
  41. Ahmed, N. Physical Properties of Porous Medium Affecting Laminal and Turbulent Flow of Water. Ph.D. Thesis, Colorado State University, Fort Collins, CO, USA, 1967. [Google Scholar]
  42. Francher, G.H.; Lewis, J.A. Flow of simple fluids through materials. Ind. Eng. Chem. 1933, 25, 1139–1147. [Google Scholar] [CrossRef]
  43. Brownell, L.E.; Kaltz, D.L. Flow of fluids through porous media. Chem. Eng. Prog. 1974, 43, 537–548. [Google Scholar]
  44. McDonald, I.F.; El-Sayed, M.S.; Mow, K.; Dillen, F.A.L. Flow through porous media-the Ergun equation revisited. Am. Chem. Soc. 1979, 18, 199–208. [Google Scholar] [CrossRef]
  45. Tyagi, A.K.; Tood, D.K. Discussion of Non-linear flow in porous media by N Ahmed and DK Sunada. J. Hydraul. Div. 1970, 96, 1734–1738. [Google Scholar] [CrossRef]
Figure 1. Relationship between  λ p  and  R p  for smooth spheres  D  = 15.97 mm; n = 36.90%;  n  = 41.50%;  n  = 37.20%; 30 <  R p   < 3590.  F  = 1.0 Data from Dudgeon [28]. Source: authors.
Figure 1. Relationship between  λ p  and  R p  for smooth spheres  D  = 15.97 mm; n = 36.90%;  n  = 41.50%;  n  = 37.20%; 30 <  R p   < 3590.  F  = 1.0 Data from Dudgeon [28]. Source: authors.
Water 15 03578 g001
Figure 2. Relationship between  f p   a n d   R p  for rolled aggregates  D  = 22.00 mm; 730 <    R p   < 2150.  F  = 1.05 Data from Martins [11]. Source: authors.
Figure 2. Relationship between  f p   a n d   R p  for rolled aggregates  D  = 22.00 mm; 730 <    R p   < 2150.  F  = 1.05 Data from Martins [11]. Source: authors.
Water 15 03578 g002
Figure 3. Relationship between   λ p  and  R p  for rolled aggregates  D  = 22.00 mm;  n  = 36.22%;  n  = 36.83% .   F  = 1.05. Data from Martins. Source: authors.
Figure 3. Relationship between   λ p  and  R p  for rolled aggregates  D  = 22.00 mm;  n  = 36.22%;  n  = 36.83% .   F  = 1.05. Data from Martins. Source: authors.
Water 15 03578 g003
Figure 4. Relationship between  α *  and  D  for rolling and crushed aggregates, 1.77 mm <   D   < 56.80 mm. Data from Sedghi-Als et al. [32] and M-B Salahi [23]. Source: authors.
Figure 4. Relationship between  α *  and  D  for rolling and crushed aggregates, 1.77 mm <   D   < 56.80 mm. Data from Sedghi-Als et al. [32] and M-B Salahi [23]. Source: authors.
Water 15 03578 g004
Figure 5. Relationship between  f t   and   D t  for smooth curves depending on the absolute roughness of the pipe  ε t . Source: authors.
Figure 5. Relationship between  f t   and   D t  for smooth curves depending on the absolute roughness of the pipe  ε t . Source: authors.
Water 15 03578 g005
Figure 6. Relationship between α* and  D  for smooth spheres; 0.50 mm <  D   < 30.00 mm. Blake; Burke and Plummer; Lindquist [36,37,38]; Ward [30]; Sunada; Kirkham [39,40]; Dudgeon [28] and Crawford et al. [8]. Source: authors.
Figure 6. Relationship between α* and  D  for smooth spheres; 0.50 mm <  D   < 30.00 mm. Blake; Burke and Plummer; Lindquist [36,37,38]; Ward [30]; Sunada; Kirkham [39,40]; Dudgeon [28] and Crawford et al. [8]. Source: authors.
Water 15 03578 g006
Figure 7. Relationship between β* and  D  for smooth spheres; 0.50 mm <  D   < 30.00 mm. Data from Blake; Burke and Plummer; Lindquist; Sunada; Kirkham [36,37,38,39,40]; Dudgeon [28] and Crawford et al. [9]. Source: authors.
Figure 7. Relationship between β* and  D  for smooth spheres; 0.50 mm <  D   < 30.00 mm. Data from Blake; Burke and Plummer; Lindquist; Sunada; Kirkham [36,37,38,39,40]; Dudgeon [28] and Crawford et al. [9]. Source: authors.
Water 15 03578 g007
Figure 8. Relationship between α* and  D  for rounded particles; 1.00 mm <  D   < 60.00 mm. Data from: Lindquist [38]; Ward [30]; Dudgeon [28]; Arbhabhirama and Dinoy [31]; Sedghi-Als et al. [32]; and Salahi et al. [23]. Source: authors.
Figure 8. Relationship between α* and  D  for rounded particles; 1.00 mm <  D   < 60.00 mm. Data from: Lindquist [38]; Ward [30]; Dudgeon [28]; Arbhabhirama and Dinoy [31]; Sedghi-Als et al. [32]; and Salahi et al. [23]. Source: authors.
Water 15 03578 g008
Figure 9. Relationship between β* and  D  for rounded particles; 1.00 mm <  D   < 110.00 mm. Data from: Lindquist [38]; Dudgeon [28]; Arbhabhirama and Dinoy [31]; Martins [11] and Salahi et al. [23]. Source: authors.
Figure 9. Relationship between β* and  D  for rounded particles; 1.00 mm <  D   < 110.00 mm. Data from: Lindquist [38]; Dudgeon [28]; Arbhabhirama and Dinoy [31]; Martins [11] and Salahi et al. [23]. Source: authors.
Water 15 03578 g009
Figure 10. Relationship between  α *  and  D  for angular particles; 3.00 mm <   D   < 40.00 mm. Data from: Dudgeon [28] and Arbhabhirama and Dinoy [31]. Source: authors.
Figure 10. Relationship between  α *  and  D  for angular particles; 3.00 mm <   D   < 40.00 mm. Data from: Dudgeon [28] and Arbhabhirama and Dinoy [31]. Source: authors.
Water 15 03578 g010
Figure 12. Schematic diagram of test equipment.
Figure 12. Schematic diagram of test equipment.
Water 15 03578 g012
Figure 13. Positions of pressure measurement points.
Figure 13. Positions of pressure measurement points.
Water 15 03578 g013
Figure 14. Positions of pressure measuring points in a cross-section.
Figure 14. Positions of pressure measuring points in a cross-section.
Water 15 03578 g014
Figure 15. Stainless-steel sample pipe. Pressure measuring points.
Figure 15. Stainless-steel sample pipe. Pressure measuring points.
Water 15 03578 g015
Figure 16. Booster pump (left), flowmeter (center), and digital pressure readers of the seven piezometers shown in Figure 13 (right).
Figure 16. Booster pump (left), flowmeter (center), and digital pressure readers of the seven piezometers shown in Figure 13 (right).
Water 15 03578 g016
Figure 17. Relationship between pressure y and distance x in cm for the  D  = 10 mm material. Source: authors.
Figure 17. Relationship between pressure y and distance x in cm for the  D  = 10 mm material. Source: authors.
Water 15 03578 g017
Figure 18. Relationship between pressure y and distance x in cm for the  D  = 20 mm material. Source: authors.
Figure 18. Relationship between pressure y and distance x in cm for the  D  = 20 mm material. Source: authors.
Water 15 03578 g018
Figure 19. Relationship between pressure y and distance x in cm for the  D  = 35 mm material. Source: authors.
Figure 19. Relationship between pressure y and distance x in cm for the  D  = 35 mm material. Source: authors.
Water 15 03578 g019
Figure 20. Relationship between pressure y and distance x in cm for the  D  = 45 mm material. Source: authors.
Figure 20. Relationship between pressure y and distance x in cm for the  D  = 45 mm material. Source: authors.
Water 15 03578 g020
Figure 21. Relationship between  f p  and  R p  for crushed aggregates; 1.00 mm <  D   < 45.00 mm from CEDEX tests. Source: authors.
Figure 21. Relationship between  f p  and  R p  for crushed aggregates; 1.00 mm <  D   < 45.00 mm from CEDEX tests. Source: authors.
Water 15 03578 g021
Figure 22. Relationship between  λ p  and  R p . Equation  ( 29 ) ; 10.00 mm <  D   < 45.00 mm. CEDEX tests. F = 1.25. Source: authors.
Figure 22. Relationship between  λ p  and  R p . Equation  ( 29 ) ; 10.00 mm <  D   < 45.00 mm. CEDEX tests. F = 1.25. Source: authors.
Water 15 03578 g022
Figure 23. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 10 mm.
Figure 23. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 10 mm.
Water 15 03578 g023
Figure 24. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 20 mm.
Figure 24. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 20 mm.
Water 15 03578 g024
Figure 25. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 35 mm.
Figure 25. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 35 mm.
Water 15 03578 g025
Figure 26. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 45 mm.
Figure 26. Relationship between  i / V  and velocity V by use of the Forchheimer equation. D = 45 mm.
Water 15 03578 g026
Figure 27. Relationship between  α *  and  D D  = 10.00 mm;  D  = 20.00 mm;  D  = 35.00 mm;  D  = 45.00 mm. Source: authors.
Figure 27. Relationship between  α *  and  D D  = 10.00 mm;  D  = 20.00 mm;  D  = 35.00 mm;  D  = 45.00 mm. Source: authors.
Water 15 03578 g027
Figure 28. Relationship between  β *  and  D D  = 10.00 mm;  D  = 20.00 mm;  D  = 35.00 mm;  D  = 45.00 mm. Source: authors.
Figure 28. Relationship between  β *  and  D D  = 10.00 mm;  D  = 20.00 mm;  D  = 35.00 mm;  D  = 45.00 mm. Source: authors.
Water 15 03578 g028
Table 1. Values of the unified dimensionless quadratic β* coefficient for the fully developed turbulent regime. Rolled aggregates. Data from Martins [11] Source: authors.
Table 1. Values of the unified dimensionless quadratic β* coefficient for the fully developed turbulent regime. Rolled aggregates. Data from Martins [11] Source: authors.
D (mm)TestnFRh/DN   R p   f p β* (Equation (35))N u  (Equation (36))β* (Equation (33))
22.00A36.22%1.0500.09071049 1.6151.272 --
22.00B36.83%1.0500.09371034 1.6601.308141.6391.290
44.00A36.18%1.0500.09054986 1.1680.920 --
44.00B35.53%1.0500.08755019 1.1100.874101.1430.900
89.00A38.23%1.0500.098517,033 1.0550.831 --
89.00B38.84%1.0500.101518,097 1.0350.815101.0520.828
Note:  F  = 1.05 Value adopted by Martins [11] for rolled aggregates.  R h / D  = 0.1 Value proposed by Parkin [34].  N  = Number of points considered to obtain the asymptote  f p  from a total of 8 tested points on each porous media.  R p  = Reynolds number of the pores at which the fully developed turbulent flow is reached.  u  = obtained by fitting Equation (36) with the total number of points of each representative size  D  tested.
Table 2. Results obtained for the hydraulic gradients  i  from the slopes of the regression lines obtained from the [x, y] diagram (Figure 17, Figure 18, Figure 19 and Figure 20).
Table 2. Results obtained for the hydraulic gradients  i  from the slopes of the regression lines obtained from the [x, y] diagram (Figure 17, Figure 18, Figure 19 and Figure 20).
Q (ls−1) V  (ms−1)   D e = 10   mm n = 40.74 %   D e = 20   mm n = 39.12 %   D e = 35   mm n = 41.33 %   D e = 45   mm n = 43.40 %
Hydraulic Gradient
10.00850.0200(4)0.0019(4)
20.01690.06270.03010.01070.0013 (1)
30.02540.12990.06140.02270.0094 (2)
40.03380.21420.10380.03950.0180 (3)
50.04230.32630.16100.06150.0312
60.05070.44280.21820.08620.0459
70.05920.59210.28000.11410.0613
80.0677(4)0.37580.14870.0767
90.0761(4)0.46590.18800.1007
100.0846(4)0.56590.23020.1260
Note: (1) R2 = 0.0248. (2) R2 = 0.6347. (3) R2 = 0.8182. (4) No results were obtained.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

López, J.C.; Toledo, M.Á.; Moran, R.; Balairón, L. A Unified General Resistance Formula for Uniform Coarse Porous Media. Water 2023, 15, 3578. https://doi.org/10.3390/w15203578

AMA Style

López JC, Toledo MÁ, Moran R, Balairón L. A Unified General Resistance Formula for Uniform Coarse Porous Media. Water. 2023; 15(20):3578. https://doi.org/10.3390/w15203578

Chicago/Turabian Style

López, Juan Carlos, Miguel Ángel Toledo, Rafael Moran, and Luis Balairón. 2023. "A Unified General Resistance Formula for Uniform Coarse Porous Media" Water 15, no. 20: 3578. https://doi.org/10.3390/w15203578

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop