You are on page 1of 9

Assessment of the Ability of a Volume of Fluid Model to

Reproduce the Efficiency of a Continuous


Transverse Gully with Grate
Pedro Lopes 1; Jorge Leandro, Ph.D. 2; Rita F. Carvalho, Ph.D. 3;
Beniamino Russo, Ph.D. 4; and Manuel Gómez, Ph.D. 5
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

Abstract: This paper deals with the numerical investigation of the drainage efficiency of a continuous transverse gully with the grate’s slots
aligned in the flow direction and compared with experimental data sets. The gully efficiency attained with a three-dimensional (3D) numerical
model is compared and validated against experimental data. The numerical simulations are performed using a computational fluid dynamics
volume of fluid solver. Different slopes, from 0 to 10%, and a wide range of drainage flows, from 6.67 to 66.67 L=s=m, are simulated. The
linear relation between Froude number and efficiency of the gully is in agreement to the one experimentally obtained. DOI: 10.1061/(ASCE)
IR.1943-4774.0001058. © 2016 American Society of Civil Engineers.
Author keywords: Numerical validation; Volume of fluid; Transverse gully; Drainage efficiency.

Introduction as the pavement, covered, in most of cases, by a grate that can have
different sizes and shapes. In some countries, gullies are replaced
The complete and safe drainage in a flooding situation caused by by drop sewer manholes covered with a resistant grate, and placed
extreme rainfall events is one of the most challenging concerns for on streets following the same rules as for the gullies.
hydraulic engineers in urban areas, especially because a large part The current design guidelines for United States of America
of cities are covered with impermeable areas. To drain such flows, gullies are based upon a report published by Federal Highway
an efficient urban drainage system (UDS) needs to be designed, Administration (FHWA) with topics about efficiency and size
giving special attention to gullies and manholes. These elements, requirements (FHWA 1984, 2001). In Portugal, urban drainage
also known as “linking elements,” contribute to the drainage of the elements design follows the by-law RGSPPDADAR (1995) that
flow from the surface to the underground pipe systems. Theoretical contains general rules to implement gullies but nothing related to
and numerical studies on linking elements’ efficiency are important their efficiency. Occasionally, the hydraulic performance and effi-
to create constitutive relations between the underground and sur- ciency of gullies can be found in some catalogues of grates man-
face systems that can be used to calibrate urban drainage models ufacturers (e.g., NFCO 1998). Given this worldwide diversity of
(Djordjević et al. 2005; Leandro et al. 2009, 2016). grate inlets and implementation, detailed studies are needed.
In areas where it is possible to define one point where the major Efficiency studies of several grate types and curb inlets have
portion of the flow drains to (e.g., in curves along streets, side- been conducted by WEF and ASCE (1992), Spaliviero and May
walks, or gardens), the gullies are typically efficient structures, cap- (1998), Comport and Thornton (2009), Russo and Gómez (2011) or
turing part of the flow from the surface to the sewer systems. Comport et al. (2012). The efficiency of Portuguese gullies was nu-
Gullies’ inlets are normally rectangular and placed at the same level merically studied by Carvalho et al. (2011) using a two-dimensional
volume of fluid/fractional area volume Obstacle Representation
1
Ph.D. Candidate, Dept. of Civil Engineering, Univ. of Coimbra, (VOF/FAVOR) model (https://www.flow3d.com/home/resources/
Portugal; MARE—Marine and Environmental Sciences Centre, Dept. of cfd-101/modeling-techniques/favor-vs-body-fitted-coordinates). The
Life Sciences, Univ. of Coimbra, 3004-517 Coimbra, Portugal (corresponding latter study was further extended by Martins et al. (2014) with three-
author). E-mail: pmlopes@student.dec.uc.pt dimensional (3D) numerical simulations using the OpenFOAM and
2
Professor, Dept. of Civil Engineering, Univ. of Coimbra, Portugal;
the interFoam solver. The authors compared water depths through
MARE—Marine and Environmental Sciences Centre, Dept. of Life Sciences,
Univ. of Coimbra, 3004-517 Coimbra, Portugal. E-mail: leandro@dec.uc.pt
the gully and the discharge coefficients for a large set of flow rates.
3
Professor, Dept. of Civil Engineering, Univ. of Coimbra, Portugal; In the case of drop manholes, the studies are mainly focused on their
MARE—Marine and Environmental Sciences Centre, Dept. of Life Sciences, efficiency, performance, shape, and energy dissipation (Carvalho and
Univ. of Coimbra, 3004-517 Coimbra, Portugal. E-mail: ritalmfc@dec.uc.pt Leandro 2012; Rubinato et al. 2013; Granata et al. 2014).
4
Full Professor, Hydraulic and Environmental Engineering, Technical When underground pipes get pressurized, the flow can emerge
School of La Almunia, Univ. of Zaragoza, Mayor St., La Almunia de Doña to the surface through the gullies and manholes forming a jet.
Godina, 50100 Zaragoza, Spain. E-mail: brusso@unizar.es Romagnoli et al. (2013) used an acoustic Doppler velocimeter
5
Full Professor, Flumen Research Institute, Technical Univ. of Catalonia (ADV) to measure the flow behavior and the turbulent structure of
(UPC-ETSECCPB), 1-3 Jordi Girona St., Bldg. D1, 08034 Barcelona,
the flow under reverse conditions in Portuguese gullies. Lopes et al.
Spain. E-mail: manuel.gomez@upc.edu
Note. This manuscript was submitted on April 16, 2015; approved on
(2015) and Leandro et al. (2014) furthered the study of Portuguese
February 29, 2016; published online on May 18, 2016. Discussion period gullies and found numerical relations between the flow from the
open until October 18, 2016; separate discussions must be submitted ground pipe system and the height of the jet, and characterized the
for individual papers. This paper is part of the Journal of Irrigation 2D vertical velocity profile inside a gully using the interFoam
and Drainage Engineering, © ASCE, ISSN 0733-9437. solver within OpenFOAM. Vertical frequency and preferential

© ASCE 04016039-1 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


direction of the jet at the surface were also studied. An experimental installation described in Gómez and Russo (2009). It consists of
and numerical study in a typical United Kingdom gully under sur- a 1∶1 scale model of a 1.5 m wide and 5 m long rectangular surface
charge was made by Djordjević et al. (2013) to replicate the interac- and a transversal grate placed downstream [Fig. 1(a)]. The platform
tions between the surface flow and the ground drainage systems. is able to simulate lanes with a transversal slope up to 4% and lon-
In large paved areas, such as squares, airports, or high sloping gitudinal slope up to 14% and a wide set of flow rates from 20 to
roads, the common gully inlets are ineffective to capture all of the 200 L=s. A motorized slide valve regulates the flow discharged to
rainfall water. In such cases, the continuous longitudinal transverse the model, and an electromagnetic flow meter measures the flow
gullies represent a widely accepted solution because the main design rates with an accuracy of 1 L=s. Upstream of the platform, the flow
concern is on the positioning of the grate in the direction perpendicular passes through a tank to dissipate the energy, providing horizontal
to the flow. Because of the lack of tests in transverse gullies and their flow conditions to surface water. The flow intercepted by the gully
efficiency in different conditions, Gómez and Russo (2009) experi- is conveyed to a V-notch triangular weir, and the flow measurement
mentally studied four types of grates, found typically in Spain and is carried out through a limnimeter with an accuracy of 0.1 mm.
differing in the alignment and distribution of the slots, under different Flow depths on the platform are visually obtained by reading a thin
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

longitudinal slopes and five approaching flows. They formulated four graduated scale.
linear relations, one for each grate type, which link the hydraulic ef- The transverse grate used in this study is a composition of three
ficiency to some particular flow conditions (Froude number and water single type 2 grates [Fig. 1(b)]. This work comprises forty combi-
depth) and the grate length. The study was further extended by Russo nations of eight different longitudinal slopes ix ¼ 0, 0.5, 1, 2, 4, 6,
et al. (2013) with the formulation of empirical expressions to relate 8, and 10% and five unit inflows q ¼ 6.67, 16.67, 33.33, 50.00, and
grate hydraulic performance to flow parameters and grate geometry. 66.67 L=s=m. The transversal slope is iy ¼ 0%.
The advances of the last decades in computational fluid dynam-
ics (CFD) allow the prediction of the flow in some components of
the UDS and the evaluation of the different design factors for an Numerical Model
efficient project and operation (Jarman et al. 2008; Tabor 2010;
Bennett et al. 2011; Djordjević et al. 2013). The freeware and open-
Mathematical Formulation
source OpenFOAM is one of these CFD packages. Special attention
is given to the solver interFoam because of its capability to deal The numerical simulations were performed using the solver inter-
with free-surface flow with arbitrary configurations. Foam within the freeware and open-source CFD OpenFOAM
This paper investigates the ability of the interFoam VOF model to v.2.1.0. The mass and momentum equations were solved for iso-
reproduce the drainage efficiency of a continuous transverse gully thermal, incompressible, and immiscible two phase flows, written
with a grate under different flow rates and slopes. A particular aim in their conservative form
is to assess the numerical model’s ability to predict the hydraulic
efficiency of urban drainage elements. The “Experimental Model” ∇·u¼0 ð1Þ
section describes the experimental installation and procedure. The
numerical mesh and the numerical model section are presented in ∂ρu
the “Numerical Model” section, along with the simulations per- þ ∇ · ðρuuÞ ¼ −∇p þ ∇ · τ þ ρg þ fσ ð2Þ
∂t
formed, the different meshes tested, the boundary conditions of the
numerical model, and the numerical simplifications. “Results” section where ρ = fluid density; g = gravitational
compares and validates the grate efficiency obtained numerically with acceleration; u = fluid velocity vector; fσ = volumetric surface
the experimental datasets, further discussed in “Discussion.” Finally, tension force; τ = viscous stress tensor; and ∇ · τ
the “Conclusions” section summarizes and concludes the work.
∇ · τ ¼ ∇ · ðμ∇uÞ þ ð∇uÞ · ∇μ ð3Þ

Experimental Model where μ = fluid viscosity. The modified pressure p% is adopted


by removing the hydrostatic pressure from the total pressure (p).
The experiments were performed in the laboratory of Hydraulic This is advantageous for the specification of pressure at the boun-
Department of Technical University of Catalonia using the same daries (Rusche 2002).

Fig. 1. Photographs of the experimental installation built at UPC hydraulic department: (a) rectangular platform with transverse grate downstream
(image by Beniamino Russo); (b) single type 2 grate (reprinted from Russo et al. 2013, © ASCE)

© ASCE 04016039-2 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


∇p ¼ ∇p% þ ρg þ g · x∇ρ ð4Þ in the simulations by an input of uniform velocity and flow depths,
whereas the different slopes are carried out by numerically chang-
The interFoam uses the VOF technique, first developed by Hirt
ing the direction of the gravity vector.
and Nichols (1981), to follow and capture the interface between
Earlier work on modeling gullies (Carvalho et al. 2012; Lopes
two different fluids by using a transport/advection equation. The
et al. 2015) proved that setting the interFoam to laminar character-
transport/advection equation is given by Eq. (5); the last term is the
istics can provide accurate results with fast convergence. The lam-
compressive term (Weller 2008; Berberović et al. 2009), which is
inar model solves the mass and momentum equations together with
used here to mitigate the effects of numerical diffusion and to keep
the advection equation for α without any turbulence approach. To
the gas-liquid interface sharp rather than using interface reconsti-
achieve a quick convergence to the steady state flow, the domain is
tution schemes. This term uses the compressive velocity (uc )
initially filled by water up to a height corresponding to the expected
∂α flow depth. The steady state is achieved after 5 s of simulation, time
þ ∇ · ðαuÞ þ ∇ · ½uc αð1 − αÞ' ¼ 0 ð5Þ in which the volume of water in the domain becomes constant. The
∂t
results presented are a further average of 5–6 s of simulation. The
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

Each cell of the domain is attributed an α value, ranging from 0 simulation time step starts at 5 × 10−5 s and is progressively in-
to 1 depending on which portion of the cell is occupied by fluid 1. creased to values that do not allow the Courant number to exceed
Giving to fluid 1 the physical characteristics of water and to fluid 2 the prefixed value of 0.5. Each simulation is carried out in parallel
the physical characteristics of air, cells with α ¼ 1 only have water over 3 subdomains using two Quad-Core processors of the Centau-
whereas cells with α ¼ 0 only have air, and therefore, cells with rus Cluster, installed in the Laboratory for Advanced Computing of
intermediate values are interface cells (or free-surface cells) the University of Coimbra. The 3 subdomains are vertically divided
(Ubbink 1997). The physical properties of the mixture are defined in the longitudinal direction of the mesh.
as a weighted average of the physical properties of the two fluids
denoted by subscripts 1 and 2, on the fraction occupied on each cell
Numerical Mesh
ρ ¼ αρ1 þ ð1 − αÞρ2 ð6Þ
A 3D section of the grate is built using the open-source Salome
v.6.4.0 software (Salome 2011). In this study two different meshes
μ ¼ αμ1 þ ð1 − αÞμ2 ð7Þ
are used: Mesh 1 is built with the NETGEN tetrahedral algorithm
with edges ranging from 2 to 3 mm, making a total of 42,772 cells
The volumetric surface force function is explicitly estimated by
the continuum surface force model developed by Brackbill et al. [Fig. 3(a)]; and Mesh 2 is refined both closer to the channel bottom
(1991), which is also function of the volume fraction and on the gully inlet [Fig. 3(b)]. This refinement limits the edges to
lengths lower than 1 mm, resulting in a total of 94,215 cells. Earlier
fσ ≈ σκ∇α ð8Þ work on this study showed that in order to assure a good represen-
tation of the flow field, the mesh near to the grate slots needs to be
After discretization in space and time domains, the pressure as small as 1 mm. This value is in line with other researchers’ work
implicit with splitting of operators (PISO) algorithm (Issa 1985) is on gullies (Djordjević et al. 2013) that also used such fine meshes to
used for the pressure-velocity coupling. model urban drainage devices. Extrapolating this edge size to the
full domain, the total number of cells can quickly rise to approx-
imately 45 million cells. In order to reduce the number of cells in
Model Description the domain, this study simplifies the grate geometry to one grate
Fig. 2 shows the full domain experimentally studied by Gómez slot with 27.72 mm width and assures its representativeness by rec-
and Russo (2009) for the continuous grate with slots parallel to ognizing the symmetry planes between slots and applying proper
the longitudinal flow direction. The different inflows are included boundary conditions.

(b)

(a) (c)

Fig. 2. Sketch of the experimental installation: (a) top view; (b) detail of grate; (c) cut B-B’; units in millimeters

© ASCE 04016039-3 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


(a) (b)
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Detail of the meshes used in this study: (a) Mesh 1 = homogeneous mesh with ranging spaces between 2 and 3 mm; (b) Mesh 2 = mesh refined
at the channel bottom with cells 1 mm spaced

Boundary Conditions slot as representative of the real grate could lead to an overestima-
tion of the grate efficiency because of an increased inlet area and
Five types of boundary conditions (BC) are included in the simu-
therefore intercepted flow. To overcome this issue, the total inter-
lations (Fig. 4). The inflow BC allows to specify the flow entering
cepted flow is calculated as a function of the uncovered area used
to the channel. Because of the wide quantity of simulations per-
for drainage by the use of correction coefficients implemented on
formed with different inflow conditions (flow depth and velocity),
the numerical intercepted flow (NIF), based on area proportionality,
the inflow is defined according to flow field conditions using a dy-
resulting in a corrected intercepted flow (CIF), Eq. (9)
namic BC—extension swak4Foam (OpenFOAMWiki 2013). Two
outflows are defined, one on the channel platform and the other on CIF ¼ β × NIF ð9Þ
the bottom. The atmosphere BC allows the air to leave and to enter
in the domain and is defined on the top of the domain. The total where β = coefficient that takes into account the obstructed area,
pressure on both the outflow and atmosphere BCs is set to zero. The which can take the value of coefficient C1 , C2 , or a multiplication of
wall BC, which represents the bottom of the channel and the grate C1 and C2 , as shown in Fig. 5. The reasoning behind the C coef-
walls, is set to slip-BC. The lateral walls are symmetry planes to ficients will now be explained. Given that Qi , Qif and Qout are the
assume the continuity of the domain. inflow, intercepted flow, and outflow for just one slot, respectively,
and QI and QIF are the inflow and intercepted flow for the group
of five slots, respectively, coefficient C1 takes into account the
Numerical Approaches
blocked slots. When one slot is blocked, QIF is calculated as five
The complete grate, i.e., the 3 groups of type 2 grate, exhibit non- times C1 times Qif , where C1 is given by the number of uncovered
homogeneous characteristics along the y direction (Fig. 2). Some slots, divided by the total number of slots. In this case, the QIF (or
grate slots are completely blocked with concrete (marked in the CIF) becomes equal to four times the Qif . Coefficient C2 takes into
Fig. 2 in black color) whereas others feature different lengths account the lower drainage efficiency of the shorter slots (i.e., those
(marked in Fig. 2 with gray color). In a homogeneous grate, i.e., if with 120 mm length). If two slots are 0.5 shorter in length, QIF (or
all the grate slots had the same dimensions, the single grate slot CIF) is calculated as 5 × C2 × Qif , where C2 is equal to three slots
could numerically replicate the efficiency of the complete grate. with length Ls , plus two slots with 0.5 × Ls , divided by the simu-
However, in a nonhomogeneous grate, using the 150 mm length lated five slots with length Ls .
In this study, coefficients C1 and C2 are
No: of unblocked slots in the complete grate 50
C1 ¼ ¼
N total of slots in 1.5 m 55.5
¼ 0.9009 ð10Þ

Length of slots with 120 mm þ Length of slots with 150 mm


C2 ¼
Total length simulated numerically
6 × 120 þ 11 × 150
¼ ¼ 0.9294 ð11Þ
17 × 150
C1 is always applied to the numerical intercepted flow unless the
experimental efficiency of the grate is 100%. C2 is applied if during
the numerical simulations Lw =Ls > 4=5 (i.e., 120 mm=150 mm),
where Lw is the slot length occupied with water (Fig. 6). Ideally
both coefficients should be verified numerically by running the
complete grate; however, the computational time required to do
Fig. 4. Sketch of the simplified geometry of the grate; flow comes from
so, and the existence of real data upon the coefficients could be
the inflow boundary to the outflow; all measurements are in mm
validated, deemed the test-and-try process as a preferable solution.

© ASCE 04016039-4 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


Fig. 5. Sketch of the process to calculate the coefficients C1 and C2 , used to calculated the corrected intercepted flow (CIF) using the numerical
intercepted flow (NIF); Qi , Qif , and Qout are respectively the inflow, intercepted flow, and outflow for one slot; QI , QIF , and QOUT are respectively the
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

inflow, intercepted flow, and outflow for a group of slots

Fig. 6. 3D image of free-surface average position over the slot and the correspondent choice of coefficients; for all the simulations, ix ¼ 4%; Lw is the
slot length occupied with water; Ls is the length of the grate slot

Table 1. Distribution of the Applied Correction Coefficients (C1 ) and (C2 ),


Respectively Described by (10) and (11) Table 1 summarizes the distribution of the coefficients in simula-
tions, and Fig. 6 presents the 3D average free-surface position for
Relative simulations with ix ¼ 4% and the corresponding coefficients to be
deviation (%) q (L=s=m)
employed in each case.
ix (%) 6.67 16.67 33.33 50.00 66.67
a
10.00 — C1 C1 C1 × C2 C1 × C2
8.00 —a C1 C1 C1 × C2 C1 × C2 Results
6.00 —a C1 C1 C1 × C2 C1 × C2
4.00 —a C1 C1 C1 × C2 C1 × C2
2.00 —a C1 C1 C1 × C2 C1 × C2
Influence of Mesh Refinement on Inflows
1.00 —a C1 C1 C1 × C2 C1 × C2 The inflow values acquired with Mesh 1 and Mesh 2 are assessed
0.50 —a —a C1 C1 C1 × C2 against the inflows imposed on the model, i.e., experimental in-
0.00 —a —a C1 C1 C1 × C2 flows. Table 2 presents the relative deviations between the numeri-
Note: ix = longitudinal slope of the channel in %; cal and experimental inflows for the entire set of drainage flows and
q = unit discharge in L=s=m. the highest slopes (>2%). The relative deviation for inflow (RDI) is
a
Simulations where the experimental efficiency is 1. calculated through Eq. (12)

Table 2. Relative Deviations on Inflows (RDI) for Two Meshes Proposed—Mesh 1 and Mesh 2
RDI (%) MESH 1—q (L=s=m) MESH 2—q (L=s=m)
ix (%) 6.67 16.67 33.33 50.00 66.67 6.67 16.67 33.33 50.00 66.67
10.00 22.22 37.78 16.67 3.70 1.67 0.00 0.00 1.11 0.00 0.56
8.00 38.89 13.33 7.78 6.67 0.00 0.00 0.00 0.00 0.74 0.00
6.00 16.67 11.11 7.78 0.74 2.78 0.00 0.00 1.11 0.00 0.56
4.00 22.22 6.67 8.89 2.96 0.56 0.00 2.22 0.00 0.00 1.67
2.00 22.22 13.33 5.56 1.48 0.00 0.00 0.00 0.00 0.74 2.22
Note: ix = longitudinal slope of the channel; q = unit discharge.

© ASCE 04016039-5 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


Experimental Numerical
1
y=-0.8101x+1.2233
0.8

Efficiency
0.6

0.4

0.2 Exp R2 = 0.942 (Gomez and Russo, 2009)


Num R2 = 0.753
0
0 0.001 0.002 0.003 0.004
F(h/Bg)0.812
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Contour graph of RDE; q is the unit discharge and ix is the Fig. 9. Graph relating the efficiency with Froude number (F), flow
longitudinal slope of the channel depth (h), and width of the grate in the flow direction (Bg )

1.0
Table 3. Quantitative Statistical Coefficients to Investigate Model
0.9 Accuracy
-10% +10%
-28% Coefficient Obtained value Optimal value
Numerical Efficiency

0.8
d 0.887 1
0.7 NSE 0.712 1
+19% RMSD 0.087 0
0.6
PBIAS −0.909 0
0.5 Note: The coefficients relate the gully efficiency (experimental versus
numerical) for the range of flows from 16.67 to 66.67 L=s=m; d = index
0.4
of agreement; NSE = Nash-Sutcliffe efficiency; PBIAS = percent bias;
0.3 RMSD = root mean square deviation.
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Experimental Efficiency

Fig. 8. Relative deviations between numerical and experimental


efficiency is tested against the experimental for the flows from
efficiency
16.67 to 66.67 L=s=m. The index of agreement (d), developed
by Willmott (1981), measures the degree of model prediction error
and varies between 0 and 1. A value of 1 indicates perfect agree-
ment between the observed and predicted values. The Nash-
Experimental Inflow-Numerical Inflow Sutcliffe efficiency (NSE) coefficient (Nash and Sutcliffe 1970)
RDIð%Þ ¼ ð12Þ ranges between −∞ and 1 and reaches the optimal value when
Experimental Inflow
NSE ¼ 1. Values between 0 and 1 are viewed as good levels of
performance whereas values lower than 0 indicate unacceptable
performance. The RMSD (root-mean-square deviation) presents
Grate Efficiency (Numerical versus Experimental) good agreement when the residuals are closer to zero. The PBIAS
Fig. 7 shows the contour graph of the relative deviations between (percent-bias) indicates, in percentage, the average tendency of the
the numerical and experimental efficiencies of the grate for the simulated values to be larger or smaller than observed ones (Gupta
complete set of simulations using Mesh 2. The relative deviation et al. 1999). Positive values indicate overestimation, whereas neg-
for efficiency (RDE) is calculated in a similar way as RDI— ative values indicate underestimation. The optimal value is 0.
Eq. (12), but instead of Experimental and Numerical Inflows, the
variables are Experimental and Numerical Efficiencies.
Fig. 8 presents the numerical efficiency plotted against the ex- Discussion
perimental values. Limits of (10%, −28% (maximum negative),
and þ19% (maximum positive) errors are added to the graph.
Influence of Mesh Refinement on Inflows
Gómez and Russo (2009) related the efficiency (E) to hydraulic
and dimensional parameters such as the Froude number (F), the Analysis of Table 2 shows that for Mesh 1, in the lowest flows
width of the grate in the flow direction (Bg ) (in our case Bg ¼ (q ¼ 6.67 L=s=m), RDIs assume values higher than 15%. This
195 mm), and the flow depth upstream of the grate (h). The pro- high discrepancy can be attributed to the lack of cells in the bottom
cedure is shown in Fig. 9. The R2 correlation coefficient of the of the channel, and consequently the vertical velocity profiles either
more adequate linear trend line, adapted to the experimental results the turbulent boundary layer are not accurately modeled. For the
(ExpR2 ) is presented in Fig. 9 as well. The NumR2 coefficient is the highest flow (q ¼ 50.00 and q ¼ 66.67 L=s=m), the relative devi-
correlation between the experimental trend line and the numerical ations are not too high, assuming values that do not exceed 7%.
data. The simulations with 6.67 L=s are discarded from the graphic Using a second mesh generation, refined near the channel bottom
because the constant efficiency is equal to 1. This disposal was also with cells not larger than 1 mm (Mesh 2), the RDIs are far better
adopted in the study of Gómez and Russo (2009). than in Mesh 1. The values are now lower than 2% both for the
Table 3 presents four quantitative statistical coefficients used highest and lower values of flow rates. This result shows that the
to verify the accuracy of the numerical model. The numerical refinement near to the bottom of the channel highly influences

© ASCE 04016039-6 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


the accuracy of the data included on the model and consequently the experimental data, showing that the expression established by
the results obtained. Gómez and Russo (2009) for this type of continuous grate with bars
parallel to the flow direction is similar to the one obtained applying
the interFoam model. The statistical coefficients shown on Table 3
Grate Efficiency (Numerical versus Experimental) support the previous conclusion. The numerical efficiency (pre-
Fig. 7 shows that RDEs are much more sensible to the inflow than dicted) obtained is similar to the experimental one (observed). For
to the slope conditions, although the global RDE increases slightly all the tests simulated, the obtained coefficients are similar to the
with the increment of both variables. The maximum positive RDE optimal value of the test (Moriasi et al. 2007).
value occurs for the intermediate slope, ix ¼ 4% and for a rel-
ative small inflow, q ¼ 16.67 L=s=m (RDE ¼ þ19%) whereas the
maximum negative RDE occurs for the maximum slope, ix ¼ 10% Conclusions
and the maximum inflow, q ¼ 66.67 L=s=m (RDE ¼ −28%). The
interFoam model is more efficient in the medium range inflows This article presents the assessment of the ability of a VOF model to
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

(about q ¼ 33.33 L=s=m) and medium-small slopes (which we de- reproduce the efficiency of a continuous transverse gully with grate
fine as ≤2%) than in the range of high and low flow rates and high with bars parallel to the flow direction. The validation of the VOF
model was done by comparing with experimental real-scale data.
slopes in which the error magnitude can be greater than (10%. A
The study focuses on its hydraulic efficiency, which is the basis
similar conclusion was obtained by Martins et al. (2014) on the
for the comparison and validation. In total, forty combinations of
simulation of a gully under normal drainage. For small flow rates,
flow rates and slopes were tested. Two different meshes were gen-
the interFoam model misrepresented the flow dropping from the
erated with the Salome platform software to represent the geometry
surface into the gully by producing a falling jet attached to the side
of the grate inlet with different refinements. The refinement of the
wall, instead of producing a free-fall jet profile. It is likely that the
mesh near the channel bottom was shown to be preponderant to
same should occur in other drop structures as manholes, weirs, or
achieve good accuracy of the numerical model, especially for shal-
orifices.
low waters.
Eq. (12) can help clarify the reason for RDEs larger than 10%.
The relative deviations between the numerical and experimental
According to Eq. (12), a negative RDE value means that the
data were calculated in relation to the flow rates and the various
numerical efficiency is higher than the experimental (efficiency
slopes. The numerical model is much more efficient in medium-
is overpredicted) whereas positive RDE values represent the oppo-
high efficiencies range, which are mostly found in urban drainage
site (efficiency is underpredicted). For lower discharges the outflow
systems. A linear relation was found between the flow Froude
jet remains attached to the upstream wall of the gully box, misrep-
number and the efficiency of the grate. The R2 values found were
resenting the jet profile and the existing void fraction that should be
similar to the experimental relations achieved in Gómez and
observed between the upstream wall and the outflow jet. This phe-
Russo (2009).
nomenon decreases the velocity of the jet and the amount of flow
This study showed that the interFoam VOF solver can provide
intercepted by the gully, thus resulting in the underprediction of
results similar to the ones obtained using experimental facilities,
the gully efficiency and positive RDEs. For the highest discharges,
rendering the use of the numerical model a useful alternative to
the jet is detached from the wall and the numerical simulation
laboratory testing in the efficiency prediction of this and other types
improves; nonetheless, in this case a 3D vortex appears between
of gullies with grates.
the wall and the outflow jet that increases both jet velocity and the
amount of flow intercepted and overpredicts the efficiency of the
gully (negative RDEs). These discrepancies are mostly found in
Acknowledgments
the regions of extreme values (both inflows and slopes). The inter-
mediate inflow ranges and medium-small slopes (≤2%), which re- The authors would like to acknowledge the financing through FCT
present most of the drainage inlet in real urban systems, show a (Fundação para a Ciência e Tecnologia, IP—Portuguese Founda-
good numerical accuracy, and the interFoam gully model can be tion for Science and Technology) through projects PTDC/AAC-
employed for assessing its efficiency. AMB/101197/2008, PTDC/ECM/105446/2008 and UID/MAR/
Fig. 8 displays experimental efficiencies against the numerical 04292/2013. Pedro Lopes was financed by FCT, through the Ph.D.
efficiencies. The perfect agreement between those values is repre- scholarship Grant SFRH/BD/85783/2012, subfinanced by MEC
sented by the continuous line and relative deviations are repre- (Portuguese Ministry of Education and Science) and FSE (European
sented with dashed lines. Although the limits of good accuracy Social Fund), under the programs POPH/QREN (Human Potential
for the numerical model are not generally defined, because it de- Operational Programme from National Strategic Reference Frame-
pends on the detail we want to resolve, in CFD simulations of UDS work) and POCH (Human Capital Operational Programme) from
and for most engineering purposes it is usual to consider a good Portugal2020. The computational time spent on the Centaurus Cluster
numerical performance for simulations in which the relative devi- of the Laboratory for Advanced Computing at University of Coimbra
ations from the real data fall below (10% (e.g., Begum et al. 2011). is also acknowledged.
As such, Fig. 8 shows the (10% limits and the maximum and mini-
mum error lines. Sixty percent of the simulations fall within the
limits (10%. Thirty percent are found in the zone limited by Notation
the −10 and −28% lines, and 10% are found in the zone limited
with the lines RDE ¼ þ10 and þ19%. Therefore, it is suggested The following symbols are used in this paper:
that the interFoam model must be carefully applied for a continu- B = width of the channel (mm);
ous grate whenever the efficiency falls below 0.68 (limit marked Bg = width of the grate in the flow direction (mm);
with dot-slash vertical line). C1 , C2 = correction coefficients (-);
Fig. 9 shows the relations between the Froude number, depth, d = index of agreement (-);
and grate length with the efficiency both for experimental and E = efficiency (%);
numerical data. The NumR2 is very similar to the one obtained with F = Froude number (-);

© ASCE 04016039-7 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


fσ = volumetric surface tension force (kg=m2 =s2 ); FHWA. (2001). “Urban drainage design manual—HEC22.” Rep. HEC No.
g = gravitational acceleration (m=s2 ); 22, FHWA-NHI-01-021, Washington, DC.
h = flow depth (mm); Gómez, M., and Russo, B. (2009). “Hydraulic efficiency of continuous
ix = longitudinal slope of the channel (%); transverse grates for paved areas.” J. Irrig. Drain. Eng., 10.1061/
L = length of the channel (mm); (ASCE)0733-9437(2009)135:2(225), 225–230.
Lg = length of the grate (mm); Granata, F., de Marinis, G., and Gargano, R. (2014). “Flow-improving
elements in circular drop manholes.” J. Hydraul. Res., 1–9.
Ls = length of the grate slot (mm);
Gupta, H., Sorooshian, S., and Yapo, P. (1999). “Status of automatic cal-
Lw = length of the grate occupied by water (mm);
ibration for hydrologic models: Comparison with multilevel expert cal-
p = total pressure (kg=m=s2 );
ibration.” J. Hydrol. Eng., 10.1061/(ASCE)1084-0699(1999)4:2(135),
p% = modified pressure (total pressure minus 135–143.
hydrostatic pressure) (kg=m=s2 ); Hirt, C. W., and Nichols, B. D. (1981). “Volume of fluid (VOF) method for
Qi , Qif , Qout = inflow, intercepted flow and outflow for one slot the dynamics of free boundaries.” J. Comput. Phys., 39(1), 201–225.
(L=s); Issa, R. I. (1985). “Solution of the implicitly discretised fluid flow equa-
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

QI , QIF , QOut = inflow, intercepted flow and outflow for a group tions by operator-splitting.” J. Comput. Phys., 62(1), 40–65.
of slots (L=s); Jarman, D. S., et al. (2008). “Computational fluid dynamics as a tool for
q = unit discharges (L=s=m); urban drainage system analysis: A review of applications and best
u = vector of mean velocity (m=s); practice.” Proc., 11th Int. Conf. on Urban Drainage on CD-ROM,
uc = compressive velocity at air-water interface (m=s); Edinburgh, Scotland, U.K., 1–10.
α = ratio between water and air in each cell (-); Leandro, J., Chen, A., Djordjević, S., and Savić, D. A. (2009). “Compari-
β = correction coeficient (-); son of 1D/1D and 1D/2D coupled (sewer/surface) hydraulic models for
κ = surface curvature (1=m); urban flood simulation.” J. Hydraul. Eng., 10.1061/(ASCE)HY.1943
μ = dynamic viscosity (kg=m=s2 ); -7900.0000037, 495–504.
ρ = fluid density (kg=m3 ); Leandro, J., Lopes, P., Carvalho, R., Páscoa, P., Martins, R., and
σ = surface tension (kg=s2 ); and Romagnoli, M. 2014. “Numerical and experimental characterization
of the 2D vertical average-velocity plane at the centre-profile and quali-
τ = shear stress tensor (kg=m=s2 ).
tative air entrainment inside a gully for drainage and reverse flow.”
Comput. Fluids, 102, 52–61.
Leandro, J., Schumann, A., and Pfister, A. (2016). “A step towards con-
sidering the spatial heterogeneity of urban key features in urban hydrol-
ogy flood modelling.” J. Hydrol., 535, 356–365.
References Lopes, P., Leandro, J., Carvalho, R. F., Páscoa, P., and Martins, R. (2015).
“Numerical and experimental investigation of a gully under surcharge
Begum, S., Rasul, M. G., Brown, R. J., Subaschandar, N., and Thomas, P.
conditions.” Urban Water J., 12(6), 468–476.
(2011). “An experimental and computational investigation of perfor-
mance of green gully for reusing stormwater.” J. Water Resuse Desalin., Martins, R., Leandro, J., and de Carvalho, R. F. (2014). “Characterization
1 (2), 99–112. of the hydraulic performance of a gully under drainage conditions.”
Bennett, P. R., Stovin, V. R., and Guymer, I. (2011). “Improved CFD sim- Water Sci. Technol., 69(12), 2423–2430.
ulation approaches for manhole mixing investigations.” 12th Int. Conf. Moriasi, D. N., et al. (2007). “Model evaluation guidelines for systematic
on Urban Drainage (ICUD12), IWA Publishing, London, 1–8. quantification of accuracy in watershed simulations.” Am. Soc. Agric.
Berberović, E., van Hinsberg, N. P., Jakirlić, S., Roisman, I., and Tropea, C. Biol. Eng., 50(3), 885–900.
(2009). “Drop impact onto a liquid layer of finite thickness: Dynamics Nash, J. E., and Sutcliffe, J. V. (1970). “River flow forecasting through
of the cavity evolution.” Phys. Rev. E, 79(3), 1–15. conceptual models. Part I—A discussion of principles.” J. Hydrol.,
Brackbill, J. U., Kothe, D. B., and Zemach, C. (1991). “A continuum 10(3), 282–290.
method for modeling surface tension.” J. Comput. Phys., 100(2), NFCO. (1998). Catalog, Neenah Foundary Company, Neenah, WI.
335–354. OpenFOAM [Computer software]. ESI Group, France.
Carvalho, R. F., and Leandro, J. (2012). “Hydraulic characteristics of a drop OpenFOAMWiki. (2013). “Contrib/swak4Foam—OpenFOAMWiki [online].”
square manhole with a downstream control gate.” J. Irrig. Drain. Eng., 〈http://openfoamwiki.net/index.php/Contrib/swak4Foam〉 (Sept. 16, 2013).
10.1061/(ASCE)IR.1943-4774.0000437, 569–576. RGSPPDADAR. (1995). “Regulamento geral dos sistemas públicos e pre-
Carvalho, R. F., Leandro, J., David, L. M., Martins, R., and Melo, N. diais de distribuição de água e drenagem de águas residuais.” Portuguese
(2011). “Numerical research of the inflow into different gullies outlets.” By-Laws 23/95 from 23rd August—Republic Diary. I: Serie-B,, Lisbon,
Computing and control for the water industry, Exeter, U.K. Portugal, 5284–5319.
Carvalho, R. F., Leandro, J., Martins, R., and Lopes, P. (2012). “Numerical Romagnoli, M., Carvalho, R. F., and Leandro, J. (2013). “Turbulence char-
study of the flow behaviour in a gully.” 4th IAHR Int. Symp. on acterization in a gully with reverse flow.” J. Hydraul. Eng., 10.1061/
Hydraulic Structures, APRH Publishing, Lisbon, Portugal. (ASCE)HY.1943-7900.0000737, 736–744 .
Comport, B. C., Cox, A. L., and Thornton, C. I. (2012). “Performance
Rubinato, M., Shucksmith, J., Saul, A. J., and Shepherd, W. (2013). “Com-
assessment of grate inlets for highway median drainage.” Urban Drain-
parison between InfoWorks hydraulic results and a physical model of an
age and Flood Control District, Colorado State Univ., CO.
urban drainage system.” Water Sci. Technol., 68(2), 372–379.
Comport, B. C., and Thornton, C. I. (2009). “Hydraulic efficiency of grate
and curb inlets for urban storm drainage.” Urban Drainage and Flood Rusche, H. (2002). “Computational fluid dynamics of dispersed two-phase
Control District, Colorado State Univ., CO. flows at high phase fractions.” Ph.D. thesis, Imperial College of Science,
Djordjević, S., et al. (2013). “Experimental and numerical investigation of Technology and Medicine Dept. of Mechanical Engineering, London.
interactions between above and below ground drainage systems.” Water Russo, B., and Gómez, M. (2011). “Methodology to estimate hydraulic
Sci. Technol., 67(3), 535–542. efficiency of drain inlets.” Proc., ICE—Water Manage., 164(2), 81–90.
Djordjević, S., Prodanović, D., Maksimović, C., Ivetić, M., and Savić, Russo, B., Gómez, M., and Tellez, J. (2013). “Methodology to esti-
D. A. (2005). “SIPSON-simulation of interaction between pipe flow mate the hydraulic efficiency of nontested continuous transverse
and surface overland flow in networks.” Water Sci. Technol., 52(5), grates.” J. Irrig. Drain. Eng., 10.1061/(ASCE)IR.1943-4774.0000625,
275–283. 864–871.
FHWA. (1984). “Drainage of highway pavements—HEC12.” Rep. HEC Salome. (2011). “SALOME 6.4 platform download [online].” 〈http://www
No. 12, FHWA-TS-84-202, Washington, DC. .salome-platform.org/downloads/salome-v6.4.0〉.

© ASCE 04016039-8 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039


Spaliviero, F., and May, R. W. P. (1998). “Spacing of road gullies. WEF and ASCE. (1992). “Design and construction of urban stormwater
Hydraulic performance of BS EN 124 gully gratings and kerb inlets.” management systems.” ASCE Manuals and Rep. on Engineering Prac-
Rep. SR 533, HR Wallingford, U.K. tice No. 77, New York.
Tabor, G. (2010). “OpenFOAM: An Exeter perspective.” 5th European Weller, H. (2008). “A new approach to VOF-based interface capturing
Conf. on Computational Fluid Dynamics—ECCOMAS CFD, Pereira, methods for incompressible and compressible flows.” Rep. TR/HGW/
J. C. F., Sequeira, A., and Pereira, J. M. C., eds., Lisbon, Portugal. 04, OpenCFD, Berkshire, U.K.
Ubbink, O. (1997). “Numerical prediction of two fluid systems with sharp Willmott, C. J. (1981). “On the validation of models.” Phys. Geog., 2(2),
interfaces.” Ph.D. thesis, Imperial College of Science, U.K. 184–194.
Downloaded from ascelibrary.org by Chonbuk National University on 05/21/16. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04016039-9 J. Irrig. Drain Eng.

J. Irrig. Drain Eng., 04016039

You might also like