Sie sind auf Seite 1von 17

662 © IWA Publishing 2015 Journal of Hydroinformatics | 17.

4 | 2015

Numerical analysis of hydraulic jumps using OpenFOAM


Arnau Bayon-Barrachina and Petra Amparo Lopez-Jimenez

ABSTRACT
Arnau Bayon-Barrachina (corresponding author)
The present paper deals with a hydraulic jump study, characterization and numerical modeling.
Petra Amparo Lopez-Jimenez
Hydraulic jumps constitute a common phenomenon in the hydraulics of open channels that Department of Hydraulic and Environmental
Engineering,
increases the shear stress on streambeds, so promoting their erosion. A three-dimensional Universitat Politècnica de València,
Cami de Vera, s/n,
computational fluid dynamics model is proposed to analyze hydraulic jumps in horizontal smooth 46022 Valencia,
Spain
rectangular prismatic open-air channels (i.e., the so-called classical hydraulic jump). Turbulence is
E-mail: arbabar@iiama.upv.es
modeled using three widely used Reynolds-averaged Navier–Stokes (RANS) models, namely: Standard
k  ε, RNG k  ε, and SST k  ω. The coexistence of two fluids and the definition of an interface
between them are treated using a volume method in Cartesian grids of several element sizes. An
innovative way to deal with the outlet boundary condition that allows the size of the simulated domain
to be reduced is presented. A case study is conducted for validation purposes (FR1 ∼ 6.10, Re1 ∼
3.5·105): several variables of interest are computed (sequent depths, efficiency, roller length, free
surface profile, etc.) and compared to previous studies, achieving accuracies above 98% in all cases. In
the light of the results, the model can be applied to real-life cases of design of hydraulic structures.
Key words | hydraulic jump, open channel, OpenFOAM, RANS, k-epsilon, k-omega

NOMENCLATURE

ANN artificial neural network h water depth


SPH smoothed particle hydrodynamics H hydraulic head
CFD computational fluid dynamics ΔH energy drop
DNS direct numerical simulation η hydraulic jump efficiency
LES large eddy simulation Γ dimensionless water depth
RANS Reynolds-averaged Navier–Stokes k turbulent kinetic energy
RNG re-normalization group ε dissipation rate of turbulent kinetic energy
SST shear stress transport ω specific dissipation rate of turbulent kinetic
FVM finite volume method energy
VOF volume of fluid μ dynamic viscosity
u freestream velocity μt turbulent eddy dynamic viscosity
uc compression velocity ν kinematic viscosity
uτ shear velocity νt turbulent eddy kinematic viscosity
p pressure Pk production of turbulent kinetic energy
t time Pb effect of buoyancy
q inlet flow rate YM effect of dilatation
xi Cartesian reference system component Sk , Sε modulus of mean rate-of-strain tensor
fb body forces C1ε , C2ε , C3ε model parameters
ρ fluid density σk, σε model parameters
doi: 10.2166/hydro.2015.041
663 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

g acceleration of gravity Equation (1) indicates


Δx mesh element size
ui
Cμ model parameter Fri ¼ pffiffiffiffiffiffiffi (1)
ghi
α water fraction in mesh element
Fr Froude number
where ui is flow freestream velocity, g is acceleration of
Re water-height-based Reynolds number
gravity, and hi is water depth. Despite the fact that the
Ma Mach number
nature of hydraulic jumps is essentially chaotic, within a
Cr Courant number
certain range of approaching Froude numbers (Fr1 ), this
w channel width
phenomenon can, to a certain extent, become stable.
x1 position of hydraulic jump toe
According to Hager (), stabilized hydraulic jumps
x2 position of roller end
occur when Fr1 ∈ [4:5, 9:0]. Lower values produce tran-
x3 position of hydraulic jump end
sition jumps, characterized by low efficiencies and the
X dimensionless longitudinal coordinate
formation of long waves of irregular period, whereas
Lj length of hydraulic jump
higher Froude numbers produce choppy jumps, which
Lr length of roller
are unstable and prone to flow detachment and wave
ξi generic flow property of fluid “i”
and spray formation. For this reason, most stilling basin
Y ratio of sequent depths
design guidelines, such as that of the US Bureau of Recla-
yþ dimensionless wall distance
mation (Peterka ), recommend aiming for Froude
uþ dimensionless velocity
number values that produce stabilized hydraulic jumps.
Ei error in variable i results
Characterizing and analyzing this phenomenon is of
R2 coefficient of determination
paramount importance from both the technical and the
 superindex relative to classical hydraulic jump
environmental point of view. Hager () and Chanson
1 subindex relative to supercritical flow
() performed extensive reviews on the attempts to
2 subindex relative to subcritical flow
study this phenomenon throughout history. Some of these
works focused on a theoretical comprehension of the
characteristic features of the classical hydraulic jump. The
so-called classical hydraulic jump is defined by Hager
() as the hydraulic jump that occurs in smooth horizon-
INTRODUCTION tal prismatic rectangular channels. Resch & Leutheusser
() performed a thorough study on air entrapment and
Hydraulic jumps are the most used method to dissipate energy dissipation processes depending on the inlet flow
energy in hydraulic structures and they occur in water characteristics. Gualtieri & Chanson () extended this
flows suddenly changing from a supercritical regime to analysis to a wider range of inlet flow conditions.
subcritical. This virulent phenomenon is characterized Most of the studies performed up until now have focused
by large pressure and velocity fluctuations, air entrain- on the analysis of easily-measurable external macroscopic
ment and turbulent dissipation processes. It can variables using an experimental approach. This can be
therefore trigger erosion processes or scour on hydraulic explained partially by the difficulty in measuring certain vari-
structures with calamitous consequences. By definition, ables using non-intrusive acoustic and optical methods in
hydraulic jumps occur in gravity-driven flows when highly aerated flows (Murzyn et al. ). However, since
the Froude number (ratio of inertia to gravitational the 1970s, coinciding with the emergence of computational
forces) drops below unity. The Froude number is a dimen- fluid dynamics (CFD), more and more studies on the hydrau-
sionless number of fluid mechanics that, in horizontal lic jump are conducted by means of numerical methods. In
channels at a given section i, can be computed as this regard, computational techniques brought a brand new
664 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

approach to water engineering modeling. For example, some et al. () used ANN to successfully predict aquifer dis-
hard-to-model phenomena, such as heat transfer (Thomas charge processes. Farhoudi et al. () used fuzzy logic
et al. ) or coupled biological processes (Muttil & Chau methods to analyze the scour downstream of stilling
), could be for the first time implemented thanks to basins. Other authors used one-dimensional and two-dimen-
numerical methods. This implied a whole paradigm shift sional approaches to reproduce the flow in similar
and so the literature on this topic is vast: e.g., Ma et al. geometries (Dewals et al. ). However, in hydraulic
(), among others, modeled hydraulic jumps using different engineering, flows are generally strongly three-dimensional
CFD techniques; Caisley et al. () managed to reproduce (Ahmed & Rajaratnam ). Therefore, the use of a fully
accurately a hydraulic jump in a canoe chute using FLOW- three-dimensional deterministic model, such as proposed
3D; and Bombardelli et al. () used this commercial soft- here, allows its application to a wider range of cases. Com-
ware to successfully model a stepped spillway following a parisons among different numerical methods to model
similar approach to that proposed here. Other approaches hydrological and hydraulic phenomena can be found in
different from CFD have also been used in all kinds of the literature (Chen & Chau ; Wu et al. ).
water engineering applications, e.g., De Padova et al. () The main goal of this work is to propose a fully three-
successfully reproduced a hydraulic jump using smoothed dimensional CFD-based method to model classical hydraulic
particle hydrodynamics (SPH) techniques. jumps using the open source platform OpenFOAM. The use
However, as Murzyn & Chanson (a) state, math- of freely available open source codes allows a continuous com-
ematical models still have problems in reproducing the munity-based improvement of the model and avoids having to
physics of certain hydraulic phenomena, although they pay for costly software licenses. In this sense, other models can
can contribute to their better comprehension. As Romagnoli be found, also reproducing hydraulic jumps using OpenFOAM
et al. () remark, an entire comprehension of the hydrau- (Romagnoli et al. ; Witt et al. ). However, different
lic jump internal flow features and turbulence structures has outlet boundary conditions are herein presented. This
not been achieved so far. For Murzyn & Chanson (a), change constitutes an asset as it allows the model boundaries
the main features of hydraulic jumps that have not been to be brought closer to the hydraulic jump, which involves a
fully understood are the following: fluid mixing, bubble significant saving of computational resources.
break-up and coalescence, free surface turbulent inter- A case study is also conducted, where several variables
actions, and wave formation and breaking processes. of interest, such as hydraulic jump efficiency or roller
Other authors have used a CFD approach to analyze length, are computed and compared to previous analytical
hydraulic jumps in terms of shear stresses, potential erosion and experimental studies. The sensitivity of the model to cer-
on stream boundaries and other more practical applications tain parameters, such as mesh element size, turbulence
(Chanson ). Liu & Garcia () published a model model used or boundary conditions, is assessed.
combining the CFD code OpenFOAM and the Exner As discussed in later sections, given the accuracy of the
Equation to model erosion and sedimentation processes in results that have been achieved, this model is fully appli-
hydraulic structures using mesh deformation. cable to more complex geometries where hydraulic jumps
Nevertheless, despite the increasing number of publi- have to be investigated, such as dam spillways, stilling
cations in this area, in the numerical modeling of basins and river rapids to name but a few.
hydraulic structures and water engineering applications,
deterministic models (e.g., CFD) are overwhelmed in
number by their statistical counterparts, also known as METHODS
‘black box’ models (Chau et al. ). Thus, several authors
used artificial neural networks (ANNs) to successfully pre- Geometry and mesh
dict the scour that occurs at hydraulic structures, such as
bridge piers (Toth & Brandimarte ) or culvert outlets In this model of the hydraulic jump, the geometry to discre-
(Liriano & Day ). Taormina et al. () and Cheng tize is rather simple: the domain consists of a prismatic
665 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

rectangular channel. For this discretization, two big cat- structured rectangular hexahedral mesh is considered the
egories of approaches are normally used, namely: best choice.
unstructured and structured meshing. In some cases, meshes can be slightly refined in the
Unstructured meshes are generally better suited for a vicinity of solid boundaries for accurately resolving the
selective refinement, so preventing the over-refinement of flow features in boundary layers, where larger property gra-
regions where no large gradients of property are expected dients occur. This may result in the formation of highly
(Kim & Boysan ). Besides, these meshes fit better into skewed elements, although this is not a real issue as long
complex geometries, show less closure issues and their arbi- as orthogonality between the mesh axes and solid bound-
trary topology makes automatizing the meshing process aries is ensured (Hirsch ). However, in this case,
easier (Biswas & Strawn ). Nevertheless, none of these experience demonstrates that the mesh element size neces-
advantages applies to the case under study, as the geometry sary to capture the freestream flow features is not smaller
is extremely simple and no mesh refinement is required. than that necessary to resolve the boundary layer features.
Some authors state that mesh non-orthogonality does As a consequence, cubic mesh elements of uniform size Δx
not affect results as long as the skewness of its elements is are used throughout the entire domain (see Figure 1). The
kept low enough (Huang & Prosperetti ). Nevertheless, optimum mesh element size is highly case specific, so it is
structured meshes tend to be more accurate than their determined by means of a mesh sensitivity analysis.
unstructured counterparts ceteris paribus (Biswas &
Strawn ). Besides, structured meshing algorithms are Numerical model
generally more straightforward to implement and faster to
execute. According to Keyes et al. () structured meshing Water level of open channel flows can be obtained by shal-
algorithms present a more regular access to memory, which low wave approaches. However, they are not sufficient
significantly reduces its latency. Also, as discussed below, in when modeling complex geometries as only a water depth
multiphase flows, topologically orthogonal meshes with value is assigned to each point on the streambed. In cases
their axis aligned with the fluid interface tend to show where a full description of the flow characteristics is necess-
fewer numerical problems. For all these reasons, a static ary, resolving the Navier–Stokes Equations becomes a must.

Figure 1 | Example of a mesh used in the model with zoom on the jump toe region.
666 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

Equations (2) and (3) are the Navier–Stokes Equations for Hence, the PIMPLE algorithm is used here as a good
mass and momentum conservation in their incompressible compromise between computation requirements and stab-
form. Unfortunately, their complete analytical resolution ility. This algorithm is implemented in OpenFOAM, a
has not been achieved so far, so numerical models are freely available open source platform constituted by all
necessary to approximate a solution to every problem sorts of C þþ applications and libraries to solve all kinds
involving fluid motion. of continuum mechanics problems (Weller et al. ).
This code uses a tensorial approach and object-oriented pro-
¼0
∇u (2) graming techniques following the widely known finite
volume method (FVM), first used by McDonald (). An
in-depth explanation of the algorithm implementation can

@u 1
  ∇u
þu  ¼  ∇p þ υ∇2 u
 þ fb (3) be found in the PhD theses of Jasak (), Ubbink ()
@t ρ
and Rusche ().
where u is velocity, p is pressure, t is time, ρ is density, υ is
kinematic viscosity, and fb is body forces (gravity and sur- Water surface tracking
face tension). The flow is assumed to be incompressible in
order to save computational resources and so density vary- The coexistence and interaction of several fluids and the
ing terms have been canceled out. This assumption can be way that the interface among them is defined is of para-
done as Mach numbers (ratio of the flow velocity to the mount importance in numerical modeling of multiphase
sound velocity) are below the commonly accepted threshold flows. Complex algorithms must be developed to model
of Ma < 0:3 (Young et al. ). this phenomenon, whose stability and accuracy have a
A wealth of algorithms has been developed to approxi- strong influence on the model final results (Hyman ).
mate numerically the Navier–Stokes Equations during the Surface tracking methods fall into two families of
last decades. Nevertheless, none of them constitutes a per- approaches, namely: the surface methods and the volume
fect solution as their performance is highly case specific. methods. On the one hand, surface methods explicitly
Indeed, this is a topic extensively discussed in the literature define the free interface either using a Lagrangian approach,
( Jang et al. ; Barton ). It is important to remark that i.e., tracking a set of surface marker particles (Daly ), or
the algorithm performances are generally assessed in terms using a Eulerian approach, i.e., defining functions that deter-
of computation requirements and stability as, eventually, mine the free surface position (Osher & Sethian ). These
all algorithms should converge to a similar solution. The methods present topology issues when dealing with highly
most widely used algorithm to execute stationary simu- deformed flows and breaking surfaces. For this reason,
lations and normally the default option in all CFD codes is they are not considered appropriate to model hydraulic
the SIMPLE algorithm (Patankar & Spalding ). Several jumps.
improvements to its original implementation, such as On the other hand, volume methods adapt better to this
SIMPLER or SIMPLEC, have been made since the model kind of phenomena, but do not define a neat flow interface
was developed. One of the most used variations is the explicitly. Instead, a surface tracking method has to be
PISO algorithm (Issa ). However, problems may arise implemented in the model. Some models use a Eulerian–
when dealing with multiphase flows where phase changes Lagrangian approach (particles on fluid methods) combin-
are abrupt or the density difference is large (Brennan ). ing a Eulerian flow resolution with particle tracking
To overcome this issue, an algorithm was developed combin- (Harlow & Welch ). However, in three-dimensional
ing the best features of SIMPLE and PISO; the so-called models, the large number of necessary particles makes the
PIMPLE (OpenFOAM ). This algorithm merges the computational cost of this approach unaffordable. For this
outer-correction tools of SIMPLE with the inner-corrector reason, an entirely Eulerian approach is used in the present
loop of PISO in order to achieve a more robust and general- model. This kind of approach proved to be more computa-
izable pressure-velocity coupling (Rodrigues et al. ). tionally efficient as it only has to deal with a single
667 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

variable value per mesh element (Ubbink ). This vari- the momentum transfer. It also reduces the scour risk by
able is an indicator property (α) expressing the proportion cavitation (Bung & Schlenkhoff ) and the shear stres-
of one fluid or another that every mesh element contains. ses on the channel boundaries (Chanson ). Therefore,
Its distribution throughout the domain is modeled by this is a phenomenon of paramount importance in highly
approximating an additional convection transport equation aerated flows such as hydraulic jumps, bores, breaking
(Equation (4)). This implies considering both fluids, A and waves, etc. Unfortunately, surface tracking methods per
B, as a single multiphase fluid, whose properties are treated se cannot reproduce phenomena smaller than the mesh
as weighted averages according to the fraction occupied by element size, such as bubbles and droplets, or the entrap-
one fluid or another in each mesh element (see Equation ment of large amounts of air (Toge ). To overcome
(5)). This results in a set of α values between 0 and 1 this issue, additional air entrainment models are
throughout the entire modeled domain but no clear water- implemented. In low-aerated flows, a Eulerian–Lagrangian
air interface is defined a priori. approach is possible, where the Navier–Stokes Equations
are resolved and air is treated as a set of discrete particles.
@α With larger air fractions, this approach is no longer pos-
 α) ¼ 0
þ ∇  (u (4)
@t sible and an entirely Eulerian method is necessary.
Eulerian–Eulerian approaches yield better results than
ξ ¼ ξa α þ ξb (1  α) (5) their Eulerian–Lagrangian counterparts in the latter case.
Despite the fact that they require longer computation
where α is fluid fraction, u is velocity, t is time, and ξ rep- times, in entirely Eulerian approaches, buoyancy, drag
resents a flow generic property. As regards the method and lift forces are taken into account. For this reason, in
used to clean up the misty zones and so define a neat inter- the case of the model proposed here, a Eulerian–Eulerian
face, a wealth of approaches has been developed during the approach is implemented.
last decades. The traditional line techniques, such as SLIC
(Noh & Woodward ), PLIC (Youngs ) or FLAIR Turbulence
(Ashgriz & Poo ), provided the first viable solutions to
the surface definition issue in volume methods. However, Turbulence features can either be resolved down to their
they present problems of generalization to unstructured lowest scales (direct numerical simulation (DNS)), if the
meshes. The donor-acceptor methods, such as the original mesh is accordingly fine, or modeled under a wealth of
implementation of the VOF (Hirt & Nichols ), have different approaches. Despite the application of DNS
been widely used in the past, but they are prone to show models to multiphase flows having been reported
false interface deformation issues. (Borue et al. ), in most cases turbulence features are
In the present model, an interface compression algo- partial or completely modeled in common engineering
rithm is implemented in order to overcome the applications.
aforementioned issues. This method adds an extra term in Large eddy simulation (LES) approaches are also feas-
the left-hand side of Equation (4): ∇  (uc α[1  α]), where ible to model multiphase flows. Nevertheless, the most
uc is a compression velocity with normal direction to the used technique is the Reynolds-averaged Navier–Stokes
fluid interface. Multiplying uc by α[1  α] ensures that it (RANS). In these models, the so-called Reynolds stresses
will only affect those regions where the flow fraction is are averaged to find a closure to the Navier–Stokes
close to 0:5 (Rusche ). Equations. To do so, additional transport equations are
implemented in order to model the behavior of flow turbu-
Flow aeration lence. Among the available models, the performance of
three of the most used is here studied. The assessed
The aeration of a water flow modifies its volume, depth, models are the Standard k  ε (Launder & Sharma ),
density, and compressibility (Carvalho ), thus affecting the RNG k  ε (Yakhot et al. ) and the SST k  ω
668 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

(Menter ). The Standard k  ε model has been widely Boundary conditions
used in this kind of applications (López & Garcia ).
Its formulation is depicted in Equations (6) and (7) The boundary conditions imposed to force the hydraulic
jump to occur consist of a supercritical flow inlet, a subcriti-
  
@ @ @ μt @k cal flow outlet, smooth bottom and side walls and an upper
(ρk) þ (ρkui ) ¼ μþ þ Pk þ Pb
@t @xi @xj σ k @xj open air patch (see Figure 2). At the inlet, in order to fulfill
 ρε  YM þ Sk (6) the desired Froude number, a water depth (h1 ) and a poten-
tial velocity profile are imposed using a Dirichlet boundary
   condition. The pressure value is defined as a null von Neu-
@ @ @ μt @ε ε
(ρε) þ (ρεui ) ¼ μþ þ C1ε (Pk mann boundary condition, so forcing a hydrostatic profile.
@t @xi @xj σ ε @xj k
As regards the inlet variables of the RANS model, i.e., k, ε,
ε
2
þ C3ε Pb )  C2ε ρ þ Sε (7) and ω, they cannot be directly estimated from measure-
k
ments. Instead, they are set to an arbitrary low value and a
where k is turbulent kinetic energy, ε is the dissipation short initial stretch of channel is added in order for the
rate of k, t is time, ρ is density, xi is the coordinate in flow to develop while approaching the hydraulic jump.
the i axis, μ is dynamic viscosity, μt is turbulent dynamic As regards the outlet, the subcritical water height that
viscosity, Pk is production of turbulent kinetic energy, forces the hydraulic jump to occur within the simulated
Pb is the buoyancy effect, YM is the dilatation effect, and domain (h2 ) has to be imposed. This variable has to be
Sk and Sε are the moduli of mean rate-of-strain tensor. obtained by iteratively testing values until the resulting
The remaining terms, (Cμ , C1ε , C2ε , C3ε , σ k , and σ ε ) are hydraulic jump remains stable within the domain. Normally,
model parameters that, in the Standard k  ε model, are a subcritical water height and a hydrostatic profile should be
0:09, 1:44, 1:92, 0:33, 1:0, and 1:3, respectively. imposed at the outlet by means of a Dirichlet water level
The RNG k  ε model formulation differs from that of boundary condition. This, combined with a null Von Neu-
the Standard k  ε, essentially, in the values of the afore- mann boundary condition for velocity, would allow the
mentioned parameters. These changes seem to improve flow to leave the domain freely. However, the imposition
the model results to such an extent that, according to Brad- of a subcritical outlet by means of this approach in
shaw (), the RNG k  ε is the most used model in OpenFOAM appears to cause stability issues.
hydraulic applications. Indeed, to the knowledge of the authors, all cases of
Several authors claim that k  ε models are not suitable to hydraulic jump simulations using OpenFOAM reported in
model large adverse-pressure gradient flows (Menter ; the literature, such as Romagnoli et al. () or Witt et al.
Wilcox ). To overcome this issue, k  ω models were first (), have had to bypass this issue. To do so, they added
introduced by Wilcox (). Their implementation is signifi- an additional stretch of channel with an obstacle on the
cantly different from that of k  ε, as the dissipation rate of streambed, such as a step, a gate or a ramp, followed by a
turbulence kinetic energy (ε) is not modeled. Instead, a trans- conventional free outlet.
port equation for its relative value (ω ¼ ε=k) is implemented. In the present model, this problem is overcome by
Among them, the SST k  ω (Menter ) proved to perform imposing a velocity profile at the outlet and so letting the
better than the Standard and the baseline (BSL) k  ω. hydrostatic profile develop, as is done at the inlet boundary
The suitability of one model or another is highly case condition. Assuming mass conservation, this approach uni-
specific and differences from using one model or another vocally produces a given water height. This avoids having
are normally remarkable. Hence, in order to determine to model the aforementioned extra stretch of channel. As
which model performs best at a reasonable computational this implies bringing the boundary conditions closer to the
cost, a sensitivity analysis is conducted. To do so, simu- phenomenon under study, comparative simulations are run
lations are run using the three RANS models discussed in order to assess the model sensitivity to the boundary con-
above ceteris paribus. dition type.
669 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

Figure 2 | General scheme of a hydraulic jump and boundary conditions used in the model.

A no-slip condition is imposed at the walls and roughness The lowest yþ regions, the so-called viscous sublayer
is not considered (Hager ). An atmospheric boundary (Schlichting & Gersten ), are characterized by large gra-
condition is imposed at the top of the channel to allow dients of velocity and other properties and the
fluids to enter and leave the channel. This is achieved by predominance of viscous effects. To avoid having to resolve
imposing a null Von Neumann condition to all variables these regions, wall functions are often used in CFD models.
except for pressure, which is set to zero (atmospheric These functions are imposed as boundary conditions on
pressure). Figure 2 summarizes the model’s boundary con- solid patches to avoid the use of excessively fine meshes,
ditions and some of its most relevant variables. with the subsequent saving of computational resources. As
a consequence, the model mesh has to be refined so that
the yþ coordinate of the center of all mesh elements in
Wall treatment
touch with solid walls be somewhere between the buffer
and the logarithmic sublayers (yþ ∼ 30). It is important
The way the boundary layer is treated is of paramount
not to over-refine meshes when using wall functions. If
importance in fluid modeling. Von Karman () estab-
this happens, wall functions will be modeling the viscous
lished a universal law of the wall which defines the flow
sublayer, whereas the model itself would be resolving the
velocity profiles in the boundary layer. Velocity (u) and dis-
flow in this region. This controversy may cause finer
tance to wall (y) are adimensionalized using the shear
meshes to yield less accurate results.
velocity (uτ ) and the viscosity (υ), respectively
In terms of accuracy, the best choice would be to use a
low Reynolds number model with no wall function at all.

yþ ¼ y (8) However, this implies refining the mesh to such an extent
υ
that the computational cost may become unaffordable.
u There is a vast literature on improvements to the original
uþ ¼ (9)
uτ implementation of wall functions, such as Johnson &
670 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

Launder (), but most of the solutions proposed have not As regards the discretization of time derivatives, explicit
been adopted by most CFD codes. This is due to the fact schemes tend to be computationally lighter than their
that, despite these approaches being valid from a theoretical implicit homologues. However, they are also more unstable,
point of view, many of them may cause stability issues especially, in simulations with skewed meshes (Blazek )
(Blocken et al. ). or when solving RANS equations (Lafon & Yee ). There-
In this research, a high Reynolds number wall function fore, implicit time discretization schemes are preferred in
for RANS models and smooth solid surfaces is this model. This implies slightly longer computation times
implemented. The boundary layer in a case of these charac- and eventual accuracy problems due to phenomena such
teristics is likely to be slightly skewed (Taylor ). as wave damping (Casulli & Cattani ), but also higher
Nevertheless, as the flow mainstream direction is com- stability. Hence, a first-order accurate bounded implicit
pletely longitudinal, a bi-dimensional wall function is used Euler scheme is used to discretize time derivative terms.
for the sake of simplicity. The time step length is variable throughout the simu-
lation resolution process. Its value is automatically updated
Discretization schemes after every time step in order to ensure that the maximum
Courant number never overcomes a threshold of Cr < 0:75,
As regards the discretization schemes used to make the CFD ensuring convergence and stability of simulations.
model partial differential equations numerically approxim-
able, a good choice always must be a good compromise Postprocessing
between accuracy and stability. In spatial discretization,
upwind models are generally preferred to downwind The variables analyzed and compared to previous studies
approaches as the latter tend to show severe stability are now discussed. The reference values to which
issues. Compared to central differencing schemes, upwind model results are compared are denoted by a superscript
approaches are slower, but also less diffusive and so more asterisk (*).
accurate. The problem is that, when abrupt property gradi- The sequent depth (Y), i.e., the ratio of subcritical to
ents occur, the latter schemes may require limiters in supercritical flow depth (h1 and h2 , respectively), is a charac-
order to prevent spurious oscillations (Blazek ). Once teristic parameter of hydraulic jumps. According to Belanger
limited, upwind schemes, such as Van Leer’s (), are (), it can be estimated as a function of the approaching
very appealing to discretize abruptly varied properties. The Froude number using a series of simplifications of the
only drawback is that, when limited, these schemes Momentum Equation. Nevertheless, in channels of low
become first-order accurate. aspect ratio (h1 =w), side walls can play an important role
In the present model, the fluid fraction divergence and this equation is no longer valid. In this regard,
terms (α) are discretized using a limited Van Leer approach Murzyn & Chanson () claim that scale effects can
due to its abruptly variable nature. In the case of the RANS play an important role in channels of aspect ratio above
model variables, divergence terms are discretized using an 0:1. To overcome this issue, Hager & Bremen () pro-
unlimited upwind approach as they are less prone to cause posed the following approach introducing the Blasius
stability issues. The velocity divergence terms are discre- Equation to account for side wall friction effects, resulting
tized using a central differencing scheme in order to
avoid possible instabilities as well as to reduce compu- qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  h i
h2 1 2:5 Fr1 =8
Y ¼ ¼  1 þ 8  Fr21  1  1  0:7 [ log Re1 ] e
tation times. Also, all gradient and interpolation terms of h1 2
the model are discretized using this approach. Gaussian  
h1
 1  3:25 eFr1 =7 ( log Re1 )3 (10)
standard FVM is used to interpolate the variable values w
from cell centers to their faces. To save computational
resources, as this mesh is strictly Cartesian, no orthogonal- where h1 is supercritical water depth, w is channel width,
ity corrector is applied. Fr1 is approaching Froude number, and Re1 is the
671 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

supercritical height-based Reynolds number. Another rele- supercritical and subcritical water levels, respectively. The
vant variable of hydraulic jumps is the roller length (Lr ), variable X is the non-dimensional longitudinal coordinate
i.e., the stretch right downstream of the jump toe where (x), computed as a function of x1 (hydraulic jump toe pos-
water recirculation occurs and most of the air entrainment ition) and x2 (roller end position) as Equation (17) indicates
occurs. Some authors, such as Murzyn & Chanson
(b), define the roller length as the hydraulic jump hi  h2
ΓðxÞ ¼ (16)
region over which the water height increases monotonically. h1  h2
However, in this study the stagnation point is used as a cri-
terion to delimit the roller end. Hager () proposes the x  x1
X¼ (17)
following expression to estimate the roller length x2  x1

Fr1 The nature of hydraulic jumps is highly chaotic and


Lr  ¼ 12 h1 þ 100  h1  tan h (11) unstable, and so most of its characteristic variables show a
12:5
quasi-periodic behavior (i.e., patterns can eventually be
The efficiency of hydraulic jumps is defined as the ratio
observed, but their characteristic period is not constant).
of the energy drop to the upstream hydraulic head. These
For this reason, it is crucial to extend sufficiently the simu-
variables are obtained from Equation (12) as a function of
lation time to avoid bias in the results. The authors
water height (hi ), flow velocity (ui ) and acceleration of grav-
observe that stability of the solution can be assumed when
ity (g). Equation (13) represents how hydraulic jump
the residuals of all the variables drop below the 103
efficiency is computed. According to Hager & Sinniger
threshold and the water content of the whole modeled chan-
(), in classical hydraulic jumps, the latter variable can
nel stays stable during at least 10 s.
be estimated as a function of the approaching Froude
number using Equation (14)
Case study

u2i
Hi ¼ hi þ (12) A case study is conducted for validation purposes. The simu-
2g
lated case consists of a prismatic rectangular channel of
dimensions 6:00 × 0:50 × 0:75 m (length, width, and
ΔH H1  H2
η ¼ ¼ (13) height). The inlet flow is q ¼ 0:177 m =s and the supercriti-
3
H1 H1
cal depth is h1 ¼ 0:070 m: Hence, the inlet velocity is
u1 ¼ 5:057 m=s. The subcritical depth is obtained following
pffiffiffi!2
2 the procedure discussed above in this section. The density
η ¼ 1 (14)
Fr1 and the kinematic viscosity are ρ ¼ 1, 000 kg=m3 and
υ ¼ 106 m=s2 .
Water surface levels are a variable of paramount impor- The approaching Froude and Reynolds numbers are
tance in the design of hydraulic structures. Its accurate Fr1 ¼ 6:125 and Re1 ¼ 3:54  105 . A case of Fr1 between 6
estimation is crucial for a proper stilling basin design that and 7 is considered optimum for model validation as it cor-
avoids bank overflows. In the present work, the average responds to the middle of the range of Fr1 values
water surface levels are numerically computed and com- recommended by the US Bureau of Reclamation (Peterka
pared to the expression by Bakhmeteff & Matzke () ). This approaching Froude number is assumed to be
representative of the behavior of all stabilized hydraulic
Γ(x) ¼ tan h(1:5  X) (15) jumps, as described by Hager ().
As discussed above in this section, a mesh, turbulence
where Γ(x) is water level at x (hi ), non-dimensionalized fol- and boundary condition model sensitivity analysis is con-
lowing Equation (16), where h1 and h2 are the ducted. Each of the turbulence models mentioned above
672 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

(Standard k  ε, RNG k  ε, and SST k  ω) is tested in four throughout the region where streamlines cut the water free
different sized meshes. The mesh element sizes assessed are surface. Downstream of that, despite waves and small
7:00, 7:50, 7:78, and 8:75 mm, which means meshes of bubbles not disappearing completely, the characteristics of
6:511, 6:360, 4:737, and 3:467 million cells, respectively. developed flows can be observed again.
To fulfill the wall treatment function hypothesis, it is Figure 3 shows the wide span of bubble sizes that occur
checked in all cases that the yþ coordinate mostly remains in the turbulent shear and the recirculation region of hydrau-
in the range of values between 20 and 70. lic jumps. Chanson () found experimentally that the
range of bubble sizes in hydraulic jumps can extend over
several orders of magnitude. The average bubble size rapidly
decreases longitudinally. This is due to the fact that large
RESULTS AND DISCUSSION
bubbles cannot stay long in the recirculation region as
shear stresses break them up and buoyancy forces tend to
Graphic analysis
expel them (Babb & Aus ). Small bubbles are not deaer-
ated so quickly. Indeed, they can be dragged by advection
A visual analysis of the model results leads to the conclusion
forces throughout long distances until buoyancy finally
that a stabilized hydraulic jump is reached (see Figure 3). All
expels them.
the characteristic features of this kind of jump described by
Hager () can be observed, namely: compact and stable
appearance, low wave generation, gradual bubble deaera- Sensitivity analysis
tion, vortex formation within the roller, no flow separation
in the entering jet, etc. Figure 3 shows how, downstream As discussed above, a mesh, turbulence, and boundary con-
of the hydraulic jump where bubble deaeration occurs, dition model sensitivity analysis is conducted in order to
hydrostatic pressure and velocity profiles quickly reappear. determine the best combination to achieve accurate results
Also, the deaeration of large bubbles can be observed at an affordable computational cost.

Figure 3 | Instant representation of bubble formation and velocity and pressure fields.
673 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

As regards the outlet boundary condition used, Figure 4 models can be clearly observed at a mesh size of 7:50 mm,
shows examples of hydraulic jumps simulated using both thus the RNG k  ε model with 7:50 mm size mesh is the
approaches. A closer comparison between them shows no most accurate approach.
significant effect on the model outcome accuracy. Although As regards the estimation of roller lengths, the SST k  ω
an instant comparison, such as that in Figure 4, shows differ- model appears not to be able to capture accurately this vari-
ences in water level profile and flow aeration, these able. The Standard k  ε model shows a reasonable
differences completely disappear when results are averaged. accuracy (all errors are below 6%) and low sensitivity to
No undesirable effects, such as wave formation, occur mesh size, which is an asseet. However, RNG k  ε is
despite the new approach implying bringing the boundary even more accurate and shows a perfect monotonically
conditions significantly closer to the phenomenon under decreasing trend in errors, although the model is also
study. The domain reduction achieves computation times highly sensitive to mesh size variations. The RNG k  ε
up to 30% shorter in some cases. model with 7:00 mm size mesh appears to be the most accur-
The three tested turbulence models show small influ- ate approach in the roller length prediction.
ence on the sequent depth (Y) estimations. The most The prediction of the hydraulic jump efficiency achieves
accurate model is the RNG k  ε, followed by the SST the highest accuracy values, with the error of all models
k  ω and the Standard k  ε, although all errors are below 2%. The Standard k  ε with 7:00 mm size mesh is
below 4%. The inflexion point in the accuracy of the all the most accurate (0:1%) but, as observed in Figure 5, the

Figure 4 | Instant comparison of hydraulic jumps simulated using two different outlet boundary conditions: the traditional approach (top) and a new approach (bottom).

Figure 5 | Mesh and turbulence model sensitivity analysis – relative errors in the computation of sequent depths (Y), roller lengths (Lr ), and hydraulic jump efficiency (η) with respect to
bilbiography.
674 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

sensitivity of this variable to the model parameters is ex- streambed generally occur within this stretch. In stilling
tremely low. basin design cases, this can be very helpful in order to deter-
All further discussion on the results and model vali- mine the region of the structure that must be protected
dation is exclusively conducted using the results of the against scour.
RNG k  ε model with 7:00 mm size mesh. This is because Following Equation (13), the mean efficiency of the
this approach achieved the most accurate results in the hydraulic jump is η ¼ 58:2%. This agrees with the result
most sensitive variable (i.e., roller length) while being obtained from Equation (14) (Hager & Sinniger ),
reasonably accurate in the prediction of the less sensitive which estimates an efficiency of 59:0%, with only a 1:3%
variables. error. The good agreement in the computation of these
three variables compared to other works is depicted in
Quantitative analysis Figure 6.
The comparison of water surfaces to previous studies
The mean subcritical water depth obtained from the CFD (Bakhmeteff & Matzke ) proves the consistency of the
model is h2 ¼ 0:553 m, which leads to a mean ratio of model presented herein. Figure 7 shows the mean water
sequent depths of Y ¼ 7:916. Following Equation (10), levels computed using the three different turbulence
the reference value for this variable in classical hydraulic models. The most accurate RANS model is the Standard
jumps is Y  ¼ 7:951. This means that the model yields a kε (R2 ¼ 99:6%), followed by the RNG kε
value approximately 0:4% lower than that obtained using (R2 ¼ 99:2%) and the SST k  ω (R2 ¼ 98:5%). It can be
the expression proposed by Hager & Bremen (). deduced from the above that turbulence models exert very
Regarding the mean roller length, the model yields a little effect on the water free surface definition in average
mean value of Lr ¼ 2:320 m, being Lr  ¼ 2:330 m the value terms.
computed using Equation (11). This implies an underestima- Nevertheless, an instant observation of the evolution of
tion of only 0:4%. The accuracy in the prediction of this this variable shows that SST k  ω models produce a more
variable is crucial, as the largest shear stresses on the unstable and bursting water surface with high bubble and

Figure 6 | CFD model result comparison with previous analytical and experimental works. The analyzed variables are: sequent depth (Y), roller length (Lr ), and hydraulic jump efficiency (η).
675 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

addressed and solved. This problem involves the outlet


boundary condition, where an additional stretch of channel
with an obstacle attached to its bottom has to be added in
order to force the subcritical flow to occur. Using the
approach proposed in this paper, this additional stretch of
channel can be removed by modifying the outlet boundary
condition, with the subsequent saving of computational
resources (up to 30% in some cases) with no effects on the
model accuracy whatsoever.
Figure 7 | Water free surface level according to turbulence model used compared to
A case study of approaching Froude number Fr1 ¼ 6:125
previous studies. is simulated and the results are compared to previous studies
of similar characteristics, such as that of Hager () and
spray production. Both k  ε models produce smoother sur- Wu & Rajaratnam (). The roller length appears to be
faces, even though the Standard k  ε is also the model that the most sensitive variable to model parameters (the SST
yields a more uniform and less turbulent free surface. k  ω is not even able to capture this magnitude). Some
Table 1 shows the accuracy of the model according to hydraulic jump variables are better reproduced in compari-
the variable predicted using the most accurate approach son with other authors’ results, such as the sequent depth
(RNG k  ε turbulence model with 7:0 mm size mesh). It (error of only 0:4%), whereas others show lower accuracies,
can be deduced from such accurate results that the model e.g., the hydraulic efficiency (error of 1:3%). The water free
proposed herein is validated and so can be applied to real- surface is accurately reproduced by all turbulence models
life design cases. in average terms, the Standard k  ε being the most accurate
approach. An instant observation of the results shows that
the SST k  ω model surface looks more turbulent than its
CONCLUSIONS k  ε counterparts. In any case, the accuracy of all of the
variables analyzed is above 98% in all cases so the model
A three-dimensional CFD model for transient multiphase can be considered validated.
incompressible flow is developed in order to predict the be- In the light of the results, the model is ready to be
havior of classical hydraulic jumps. After analyzing the applied to real-life design cases, such as dam stilling
effects on the results of several model parameters, such as basins, stepped spillways, river rapids, meandering chan-
the mesh refinement degree, the turbulence model or the nels, etc. As discussed above, the most accurate turbulence
boundary conditions, a stable hydraulic jump and accurate model in this kind of application is the RNG k  ε, although
results are obtained. The model is built using exclusively very fine meshes are necessary to ensure good performance
free open source code, which implies avoiding expensive and this model proved to be slightly slower than the Stan-
software licenses. Also, a problem found in other cases dard k  ε. The latter turbulence model could be a better
where OpenFOAM is used to model hydraulic jumps is choice in cases where low computational requirements are
preferred without compromising accuracy excessively. The
Table 1 | RNG k  ε with 7:0 mm mesh model outcome and analytical/experimental data Standard k  ε also proved capable of reproducing the aver-
comparison age water free surface slightly better.
As for future work, the model is currently being used in
Variable Model output Reference result Accuracy
similar applications, both theoretical, such as triangular, cir-
Sequent depth (Y) 7:916 7:951 99:6%
cular and radial hydraulic jumps, and real-life cases. Also,
Roller length (Lr ) 2:320 m 2:330 m 99:6%
the air entrainment and concentration distribution in
Efficiency (η) 58:2% 59:0% 98:7%
hydraulic jumps is being studied using this model and com-
Surface (Γ)   99:2%
pared with experimental data.
676 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

ACKNOWLEDGEMENTS micro-roughness on energy dissipation and oxygen transfer.


In: 1st IAHR European Congress, May, Edinburgh, UK.
Caisley, M. E., Bombardelli, F. A. & Garcia, M. H.  Hydraulic
This research was conducted thanks to the funding provided model study of a Canoe Chute for low-head Dams in Illinois
by the VALi þ D R&D Program of the Generalitat (HES 63). Hydraulic Engineering Series 63. http:hdl.handle.
Valenciana (Spain). It would not have been possible net/2142/12214.
Carvalho, R.  Acçoes hidrodinamicas em estruturas hidraulicas:
without the contribution of Daniel Valero and Beatriz
modelacao computacional do ressalto hidráulico. Universidade
Nacher of the Hydraulics Laboratory of the School of de Coimbra, Portugal.
Civil Engineering (Universitat Politècnica de València). Casulli, V. & Cattani, E.  Stability, accuracy and efficiency of a
semi-implicit method for three-dimensional shallow water flow.
Computers & Mathematics with Applications 27, 99–112.
Chanson, H.  Drag reduction in open channel flow by aeration
REFERENCES
and suspended load. Journal of Hydraulic Research 32, 87–101.
Chanson, H.  Boundary shear stress measurements in undular
Ahmed, F. & Rajaratnam, N.  Three-dimensional turbulent flows: application to standing wave bed forms. Water
boundary layers: a review. Journal of Hydraulic Research 35, Resources Research 36, 3063–3076.
81–98. Chanson, H.  Hydraulics of aerated flows: qui pro quo?
Ashgriz, N. & Poo, J.  FLAIR: flux line-segment model for Journal of Hydraulic Research 51, 223–243.
advection and interface reconstruction. Journal of Chau, K., Wu, C. & Li, Y.  Comparison of several flood
Computational Physics 93, 449–468. forecasting models in Yangtze River. Journal of Hydrologic
Babb, A. & Aus, H.  Measurements of air in flowing water. Engineering ASCE 10, 485–491.
Journal of Hydraulics Division ASCE 107, 1615–1630. Chen, W. & Chau, K.  Intelligent manipulation and calibration
Bakhmeteff, B. A. & Matzke, A. E.  The hydraulic jump in of parameters for hydrological models. International Journal
terms of dynamic similarity. Transactions ASCE 100, of Environment and Pollution 28, 432–447.
630–680. Cheng, C., Chau, K., Sun, Y. & Lin, J.  Long-term prediction
Barton, I.  Comparison of SIMPLE-and PISO-type algorithms of discharges in Manwan Reservoir using artificial neural
for transient flows. International Journal for Numerical network models. Advances in Neural Networks 1040–1045.
Methods in Fluids 26, 459–483. Daly, B. J.  A technique for including surface tension effects in
Belanger, J.  Notes sur l’Hydraulique. Ecole Royale des Ponts hydrodynamic calculations. Journal of Computational
et Chaussees, Paris, France. Physics 4, 97–117.
Biswas, R. & Strawn, R. C.  Tetrahedral and hexahedral mesh De Padova, D., Mossa, M., Sibilla, S. & Torti, E.  3D SPH
adaptation for CFD problems. Applied Numerical modelling of hydraulic jump in a very large channel. Journal
Mathematics 26, 135–151. of Hydraulic Research 51, 158–173.
Blazek, J.  Computational Fluid Dynamics: Principles and Dewals, B., André, S., Schleiss, A. & Pirotton, M.  Validation
Applications. Elsevier, Amsterdam. of a quasi-2D model for aerated flows over stepped spillways
Blocken, B., Stathopoulos, T. & Carmeliet, J.  CFD simulation for mild and steep slopes. In: Proceedings of the 6th
of the atmospheric boundary layer: wall function problems. International Conference of Hydroinformatics, pp. 63–70.
Atmospheric Environment 41, 238–252. Farhoudi, J., Hosseini, S. & Sedghi-Asl, M.  Application of
Bombardelli, F. A., Meireles, I. & Matos, J.  Laboratory neuro-fuzzy model to estimate the characteristics of local
measurements and multi-block numerical simulations of the scour downstream of stilling basins. Journal of
mean flow and turbulence in the non-aerated skimming flow Hydroinformatics 12, 201–211.
region of steep stepped spillways. Environmental Fluid Gharangik, A. M. & Chaudhry, M. H.  Numerical simulation
Mechanics 11, 263–288. of hydraulic jump. Journal of Hydraulic Engineering ASCE
Borue, V., Orszag, S. & Staroslesky, I.  Interaction of surface 117, 1195–1211.
waves with turbulence: direct numerical simulations of Gualtieri, C. & Chanson, H.  Experimental analysis of Froude
turbulent open channel flow. Journal of Fluid Mechanics number effect on air entrainment in the hydraulic jump.
286, 1–23. Environmental Fluid Mechanics 7, 217–238.
Bradshaw, P.  Understanding and prediction of turbulent flow. Hager, W. H.  Energy Dissipators and Hydraulic Jump.
International Journal of Heat and Fluid Flow 18, 45–54. Springer, Berlin/New York.
Brennan, D.  The Numerical Simulation of Two-Phase Flows Hager, W. H. & Bremen, R.  Classical hydraulic jump:
in Settling Tanks. Imperial College of Science, Technology sequent depths. Journal of Hydraulic Research 27, 565–583.
and Medicine, London, UK. Hager, W. H. & Sinniger, R.  Flow characteristics of the
Bung, D. & Schlenkhoff, A.  Self-aerated skimming flow on hydraulic jump in a stilling basin with an abrupt bottom rise.
embankment stepped spillways: the effect of additional Journal of Hydraulic Research 23, 101–113.
677 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

Harlow, F. H. & Welch, J. E.  Numerical calculation of time- Ma, J., Oberai, A. A., Lahey Jr, R. T. & Drew, D. A.  Modeling
dependent viscous incompressible flow of fluid with free air entrainment and transport in a hydraulic jump using two-
surface. Physics of Fluids 8, 2182. fluid RANS and DES turbulence models. Heat and Mass
Hirsch, C.  Numerical Computation of Internal and External Transfer 47, 911–919.
Flows: The Fundamentals of Computational Fluid Dynamics. McDonald, P. W.  The Computation of Transonic Flow
Butterworth-Heinemann, London, p. 1. Through Two-Dimensional Gas Turbine Cascades. Paper
Hirt, C. & Nichols, B.  Volume of fluid (VOF) method for the 71-GT-89. American Society of Mechanical Engineers.
dynamics of free boundaries. Journal of Computational Menter, F. R.  Zonal two equation k-omega turbulence models
Physics 39, 201–225. for aerodynamic flows. AIAA paper 2906, 1993.
Huang, H. & Prosperetti, A.  Effect of grid orthogonality on Murzyn, F. & Chanson, H.  Experimental assessment of scale
the solution accuracy of the two-dimensional convection- effects affecting two-phase flow properties in hydraulic
diffusion equation. Numerical Heat Transfer 26 (1), 1–20. jumps. Experiments in Fluids 45, 513–521.
Hyman, J. M.  Numerical methods for tracking interfaces. Murzyn, F. & Chanson, H. a Two-Phase Gas-Liquid Flow
Physica D: Nonlinear Phenomena 12, 396–407. Properties in The Hydraulic Jump: Review and Perspectives.
Issa, R. I.  Solution of the implicitly discretized fluid flow Nova Science Publishers, Hauppage, NY.
equations by operator-splitting. Journal of Computational Murzyn, F. & Chanson, H. b Experimental investigation of
Physics 62, 40–65. bubbly flow and turbulence in hydraulic jumps.
Jang, D., Jetli, R. & Acharya, S.  Comparison of the PISO, Environmental Fluid Mechanics 2, 143–159.
SIMPLER, and SIMPLEC algorithms for the treatment of the Murzyn, F., Mouaze, D. & Chaplin, J.  Optical fibre probe
pressure-velocity coupling in steady flow problems. measurements of bubbly flow in hydraulic jumps.
Numerical Heat Transfer, Part A: Applications 10, 209–228. International Journal of Multiphase Flow 31, 141–154.
Jasak, H.  Error Analysis and Estimation for the Finite Volume Muttil, N. & Chau, K.-W.  Neural network and genetic
Method with Applications to Fluid Flows. PhD thesis, programming for modelling coastal algal blooms. International
Imperial College of Science, Technology and Medicine, Journal of Environment and Pollution 28, 223–238.
London, UK. Noh, W. F. & Woodward, P.  SLIC (simple line interface
Johnson, R. & Launder, B.  Discussion of ‘on the calculation calculation). In: Proceedings of the Fifth International
of turbulent heat transport downstream from an abrupt pipe Conference on Numerical Methods in Fluid Dynamics,
expansion’. Numerical Heat Transfer, Part A Applications 5, June 28–July 2, 1976. Twente University, Enschede,
493–496. The Netherlands, pp. 330–340.
Keyes, D., Ecer, A., Satofuka, N., Fox, P. & Periaux, J.  Parallel OpenFOAM  OpenFOAM: The Open Source CFD Toolbox
Computational Fluid Dynamics ’99: Towards Teraflops, User Guide. The Free Software Foundation Inc. http://foam.
Optimization and Novel Formulations. Elsevier, Amsterdam. sourceforge.net/docs/Guides-a4/UserGuide.pdf (accessed
Kim, S.-E. & Boysan, F.  Application of CFD to environmental 1 December 2013).
flows. Journal of Wind Engineering and Industrial Osher, S. & Sethian, J. A.  Fronts propagating with curvature-
Aerodynamics 81, 145–158. dependent speed: algorithms based on Hamilton-Jacobi
Kucukali, S. & Chanson, H.  Turbulence measurements in the formulations. Journal of Computational Physics 79, 12–49.
bubbly flow region of hydraulic jumps. Experimental Thermal Patankar, S. & Spalding, D.  A calculation procedure for
and Fluid Science 33, 41–53. heat, mass and momentum transfer in three-dimensional
Lafon, A. & Yee, H.  On the numerical treatment of nonlinear parabolic flows. Journal of Heat and Mass Transfer 15,
source terms in reaction-convection equations. In: AIAA 30th 1787–1806.
Aerospace Sciences Meeting and Exhibition, Reno, Nevada. Peterka, A. J.  Hydraulic design of stilling basins and energy
Launder, B. E. & Sharma, B. I.  Application of the energy- dissipaters. Engineering Monograph no. 25. US Bureau of
dissipation model of turbulence to the calculation of flow Reclamation.
near a spinning disc. Letters in Heat and Mass Transfer 1, Resch, F. & Leutheusser, H.  Reynolds stress measurements in
131–138. hydraulic jumps. Journal of Hydraulic Research 10, 409–430.
Liriano, S. & Day, R.  Prediction of scour depth at culvert Rodrigues, M. A., Padrela, L., Geraldes, V., Santos, J., Matos, H. A.
outlets using neural networks. Journal of Hydroinformatics 3, & Azevedo, E. G.  Theophylline polymorphs by
231–238. atomization of supercritical antisolvent induced suspensions.
Liu, X. & Garcia, M. H.  Three-dimensional numerical model The Journal of Supercritical Fluids 58, 303–312.
with free water surface and mesh deformation for local Romagnoli, M., Portapila, M. & Morvan, H.  Computational
sediment scour. Journal of Waterway, Port, Coastal, and simulation of a hydraulic jump. Mecanica Computacional.
Ocean Engineering. ASCE 134, 203–217. XXVIII, 1661–1672.
Lopez, F. & Garcia, M. H.  Mean flow and turbulence structure Rusche, H.  Computational Fluid Dynamics of Dispersed Two-
of open-channel flow through non-emergent vegetation. Phase Flows at High Phase Fractions. PhD thesis, Imperial
Journal of Hydraulic Engineering ASCE 127, 392–402. College of Science, Technology and Medicine, London, UK.
678 A. Bayon-Barrachina & P. A. Lopez-Jimenez | Numerical analysis of hydraulic jumps using OpenFOAM Journal of Hydroinformatics | 17.4 | 2015

Schlichting, H. & Gersten, K.  Boundary-Layer Theory. Von Karman, T.  Mechanische Ahnlichkeit und Turbulenz.
Springer, Berlin. Nachrichten von der Gesellschaft der Wissenschaften zu
Taormina, R., Chau, K.-W. & Sethi, R.  Artificial neural Göttingen, Fachgruppe 1 (Mathematik) 5, 58–76.
network simulation of hourly groundwater levels in a coastal Weller, H., Tabor, G., Jasak, H. & Fureby, C.  A tensorial
aquifer system of the Venice lagoon. Engineering approach to computational continuum mechanics using
Applications of Artificial Intelligence 25, 1670–1676. object-oriented techniques. Computers in Physics 12, 620–631.
Taylor, E. S.  The skewed boundary layer. Journal of Wilcox, D.  Turbulence modelling for CFD. DCW Industries,
Basic Engineering, Transactions ASME, Series D 81, La Canada, CA, USA.
297–304. Witt, A., Gulliver, J. & Shen, L.  Bubble Visualization in a
Thomas, S., Hankey, W., Faghri, A. & Swanson, T.  One- Simulated Hydraulic Jump. The Smithsonian/NASA
dimensional analysis of the hydrodynamic and thermal Astrophysics Data System. http://arxiv.org/pdf/1310.
characteristics of thin film flows including the hydraulic 2635v1.pdf (accessed on 14 February 2014).
jump and rotation. Journal of Heat Transfer ASME 112, Wu, C., Chau, K. & Li, Y.  Predicting monthly streamflow
728–735. using data-driven models coupled with data-preprocessing
Toge, G. E.  The Significance of Froude Number in Vertical techniques. Water Resources Research 45, 1–23.
Pipes: a CFD study. University of Stavanger, Norway. doi: 10.1029/2007WR006737.
Toth, E. & Brandimarte, L.  Prediction of local scour depth at Wu, S. & Rajaratnam, N.  Transition from hydraulic jump to
bridge piers under clear-water and live-bed conditions: open channel flow. Journal of Hydraulic Engineering ASCE
comparison of literature formulae and artificial neural 122, 526–528.
networks. Journal of Hydroinformatics 13, 812–824. doi: 10. Yakhot, V., Orszag, S., Thangam, S., Gatski, T. & Speziale, C. 
2166/hydro.2011.065. Development of turbulence models for shear flows by a
Ubbink, O.  Numerical Prediction of Two Fluid Systems with double expansion technique. Physics of Fluids A: Fluid
Sharp Interfaces. PhD thesis, Imperial College of Science, Dynamics (1989–1993) 4, 1510–1520.
Technology and Medicine, London, UK. Young, D. F., Bruce, R. M., Okiishi, Th. H. & Huebsch, W. W. 
Van Leer, B.  Flux-vector splitting for the Euler equations. In: A Brief Introduction to Fluid Mechanics. John Wiley & Sons,
Eighth International Conference on Numerical Methods in London.
Fluid Dynamics, 28 June–2 July 1982, Aachen. Springer, Youngs, D. L.  An interface tracking method for a 3D Eulerian
pp. 507–512. hydrodynamics code. Technical Report 44, 35–35.

First received 20 March 2014; accepted in revised form 1 February 2015. Available online 13 March 2015

Das könnte Ihnen auch gefallen