You are on page 1of 110

Analysis of Hydrodynamic Forces on

High-Head Slide Gates Using


Computational Fluid Dynamics

Juan Camilo Arango Escobar

Universidad Nacional de Colombia


Facultad de minas, Departamento de Ingeniería Mecánica
Medellín, Colombia
2018
Analysis of hydrodynamic forces on
high-head slide gates using
Computational Fluid Dynamics

Juan Camilo Arango Escobar

Tesis o trabajo de investigación presentada(o) como requisito parcial para optar al título
de (A dissertation presented in partial fulfilment on the requirements for the degree of):

Magister en Ingeniería Mecánica


(MEng Mechanical Engineering)

Director (Supervisor):
Ph.D. Alejandro Molina

Línea de Investigación (Research Field):


Hidráulica e Hidrodinámica (Hydraulics and Hydrodynamics)
Modalidad investigativa (Research Thesis)

Universidad Nacional de Colombia


Facultad de minas, Departamento de Ingeniería Mecánica
Medellín, Colombia
2018
(A mis padres y hermanos)

El mundo quiere vivir en la cima de la montaña, sin


saber que la verdadera felicidad está en la forma de
subir la escarpada.

Gabriel García Márquez


Acknowledgements
First of all, I am grateful to my parents for the support during my education, career and
professional life, and for their unceasing encouragement and care.

To the professor Alejandro Molina, for his valuable guidance and advice as director and
teacher, and for helping me grow the knowledge and skills for the development of this
Project.

To the engineers David Calderón and José David Vera for their interest and support during
the progress of this research.

I also place on record, my sense of gratitude to one and all who, directly or indirectly, have
lent their helping hand in this research.
Resumen y Abstract IX

Abstract
Coefficients that can be applied to current analytical expressions that predict the
hydrodynamic forces and, particularly, the downpull phenomenon on an inverted 30° lip
gate at a bottom outlet of a dam, were determined from Computational Fluid Dynamics
(CFD) simulations. Two slide gates were simulated: A typical 45° lip used for model
validation and the inverted 30° lip. These simulations made use of the VOF (Volume Of
Fluid) multiphase model, embedded in the commercial software ANSYS FLUENT, to
simulate the air and water interaction in the free surface flow downstream of the gate. The
CFD predictions were successfully validated against those by analytical expressions that
related the water discharge through the gate and the gate opening at a fixed reservoir head.
The predicted downpull was in agreement with that obtained from analytical expression for
the well-studied 45° lip gate. Through the validated CFD simulation it was possible to
determine the coefficients for the analytical calculation of downpull on the inverted 30° gate,
which were not available in the state-of-the-art analytical expressions. The methodology
here described can easily be applied to different gate geometries for which design
coefficients are not available.

Keywords: Slide gates, downpull, CFD, bottom outlet, VOF


Contenido XVI

Resumen

Los coeficientes aplicables a las expresiones analíticas actuales que predicen las fuerzas
hidrodinámicas y, en particular, el fenómeno de “downpull” en una compuerta de labio
invertida de 30° en una descarga de fondo de una gran presa, se determinaron a partir de
simulaciones de dinámica de fluidos computacional (CFD). Se simularon dos compuertas
deslizantes: una con labio típico de 45° utilizado para la validación del modelo y una con
labio invertido de 30°. Estas simulaciones hicieron uso del modelo multifase VOF (Volumen
de Fluido), disponible en el software comercial ANSYS FLUENT, para simular la
interacción de aire y agua en el flujo de superficie libre aguas abajo de la compuerta. Las
predicciones de CFD fueron validadas con las expresiones analíticas que relacionan la
descarga de agua a través de la compuerta dada la posición y la presión hidrostática. El
downpull encontrado estaba de acuerdo con el obtenido a partir de la expresión analítica
para la compuerta de labio inferior de 45°. A través de la simulación CFD validada fue
posible determinar los coeficientes para el cálculo analítico de downpull en la compuerta
con labio invertido de 30°, que no estaban disponibles en las expresiones analíticas
disponibles. La metodología aquí descrita se puede aplicar fácilmente a diferentes
geometrías de compuertas para las cuales los coeficientes de diseño no están disponibles.

Palabras clave: Compuertas deslizantes, Fuerzas hidrodinámicas, “downpull”,


CFD, Descarga de fondo, VOF.
1-1

Content
Pág.

Abstract .......................................................................................................................... IX

Resumen ......................................................................................................................... X

List of figures ................................................................................................................1-3

List of tables..................................................................................................................1-6

1. Introduction ............................................................................................................1-7
1.1 Description of high-head slide gates and the downpull Phenomenon............ 1-8
1.1.1 High-head slide gates ........................................................................ 1-8
1.1.2 Description of hydrodynamic forces and the downpull phenomenon .. 1-9
1.2 Thesis outline ............................................................................................. 1-10
1.3 Fluid dynamics governing equations ........................................................... 1-10
1.3.1 Continuity equation .......................................................................... 1-11
1.3.2 Momentum equation ........................................................................ 1-11
1.3.3 Navier-Stokes equation .................................................................... 1-13
1.4 Turbulence.................................................................................................. 1-13
1.5 RANS Equations ......................................................................................... 1-15
1.5.1 RANS-equations based turbulence models ...................................... 1-15
1.5.2 k-ε turbulence model ........................................................................ 1-16
1.6 ANSYS Fluent Volume of fluid (VOF) model ............................................... 1-16
1.7 Description of the problem .......................................................................... 1-17
1.8 Scope and aim of the research ................................................................... 1-19
1.8.1 General objective ............................................................................. 1-19
1.8.2 Specific objectives............................................................................ 1-19
1.9 Theoretical framework ................................................................................ 1-20
1.9.1 Water discharge through bottom outlets ........................................... 1-20
1.9.2 Hydrodynamic forces on flat gates ................................................... 1-21
1.9.3 Factors Influencing downpull ............................................................ 1-22
1.9.4 Analytical calculation of downpull ..................................................... 1-29
1.9.5 CFD and physical model approaches for the design of hydraulic gates1-35

2. Analysis of hydrodynamic forces on high-head slide gates using CFD ..........2-38

Abstract .......................................................................................................................2-38
2.1 Introduction ................................................................................................. 2-39
2.2 Methodology ............................................................................................... 2-42
1-2

2.2.1 Geometry and mesh ........................................................................ 2-42


2.2.2 Boundary Conditions........................................................................ 2-54
2.2.3 Results ............................................................................................ 2-58
2.3 Validation ................................................................................................... 2-66
2.3.1 Water discharge............................................................................... 2-66
2.3.2 Downpull .......................................................................................... 2-67
2.3.3 Bottom downpull coefficients for inverted lip gate............................. 2-71
2.4 Conclusions ................................................................................................ 2-74
2.5 Future work ................................................................................................ 2-76

Appendixes ................................................................................................................. 2-81

Appendix A ................................................................................................................. 2-82

Appendix B ................................................................................................................. 2-84

Appendix C ................................................................................................................. 2-86

Appendix D ................................................................................................................. 2-87

Appendix E.................................................................................................................. 2-89

Appendix F .................................................................................................................... 2-1

Appendix F .................................................................................................................... 2-1


1-3

List of figures
Pág.

Figure 1. High-head slide gate - Components ......................................................................... 1-8


Figure 2. Types of gates with known downpull coefficients (Henriques Da Silva, 2011)......... 1-18
Figure 3. Types of gates with unknown downpull coefficients ................................................ 1-18
Figure 4. Upstream gate seals configuration ......................................................................... 1-23
Figure 5. Downstream gate seals configuration ..................................................................... 1-24
Figure 6. Aeration pipes at a bottom outlet ............................................................................ 1-25
Figure 7. Different gate bottom geometries. A) Flat bottom gate. B) Curved bottom gate. C)
Inclined bottom gate .............................................................................................................. 1-26
Figure 8. Gate shaft geometry on a typical intake gate .......................................................... 1-27
Figure 9. Gate shaft geometry on a bottom outlet high-head gate ......................................... 1-28
Figure 10. Face-type gate ..................................................................................................... 1-29
Figure 11. Vertical force decomposition on a flat hydraulic gate ............................................ 1-30
Figure 12. Variation of Bottom downpull Coefficient 𝐾𝑏 with relative gate opening, for gates with
inclined bottom surface (Naudascher E. , 1964) .................................................................... 1-34
Figure 13. Variation of Bottom downpull Coefficient 𝐾𝑏 with relative gate opening, for gates with
projected skin plate (Naudascher E. , 1964) .......................................................................... 1-34
Figure 14. Variation of Bottom downpull Coefficient 𝐾𝑏 with relative gate opening, for gates with
rounded bottom lip (Naudascher E. , 1964) ........................................................................... 1-35
Figure 15. Scale model of a bottom outlet with high-head segment gates. (Calderón Villegas,
2016) ..................................................................................................................................... 1-36
Figure 16. CFD simulation of the same bottom outlet – Velocity contours (IMPSA, 2016) ..... 1-37
Figure 17. High-head slide gate – Components ..................................................................... 2-39
Figure 18. Types of gates with known downpull coefficients (Henriques Da Silva, 2011) ....... 2-40
1-4

Figure 19. Types of gates with unknown downpull coefficients............................................... 2-41


Figure 20. 45° lip gate geometry ............................................................................................ 2-42
Figure 21. Inverted 30° Gate geometry .................................................................................. 2-43
Figure 22. Side and top view of the bottom outlet for the simulation ....................................... 2-44
Figure 23. 3D Geometry of the fluid contained in the bottom outlet ........................................ 2-46
Figure 24. Aeration phenomenon at the bottom outlet ............................................................ 2-46
Figure 25. Symmetry plane for the simulation of the bottom outlet ......................................... 2-47
Figure 26. Final geometry for the CFD simulation of the bottom outlet ................................... 2-48
Figure 27. Side view of the mesh for the 45° lip gate ............................................................. 2-50
Figure 28. Detail of the mesh at the bottom surfaces of the 45° lip gate ................................. 2-52
Figure 29. Detail of the mesh at the top surfaces of the 45° lip gate ....................................... 2-52
Figure 30. Boundary conditions for the CFD models .............................................................. 2-55
Figure 31. Volume fraction initialization for the simulation – 45° lip gate – 60% opening (Phase 1
corresponds to water) ............................................................................................................ 2-56
Figure 32. Pseudo steady state analysis of the downpull force – 45° lip gate – 40% gate opening
.............................................................................................................................................. 2-57
Figure 33. Pseudo steady state analysis of the mass balance – 45° lip gate – 50% opening . 2-58
Figure 34. Volume fraction of the 45° lip gate – side view – Mid plane ................................... 2-59
Figure 35. Pressure contours for different gate openings ....................................................... 2-60
Figure 36. Velocity vectors for the mixture – 45° lip gate – 50% opening ............................... 2-62
Figure 37. Streamlines colored by velocity – 45° lip gate – 80% and 50% gate openings ...... 2-63
Figure 38. Static pressure isocontours comparison – 50% gate opening – 45° and inverted 30°
gates ...................................................................................................................................... 2-65
Figure 39. Analytical vs CFD water discharge validation ........................................................ 2-67
Figure 40. Top downpull force – Analytical vs CFD 45° lip gate ............................................. 2-68
Figure 41. Bottom downpull force – Analytical vs CFD 45° lip gate ........................................ 2-69
Figure 42. Resulting downpull force – Analytical vs CFD 45° lip gate ..................................... 2-70
Figure 43. Bottom downpull coefficients comparison – Experimental vs CFD ........................ 2-73
Figure 44. Line for the recompilation of the data for the mesh Independence study ............... 2-82
Figure 45. Results of the velocity measured along the line for the four different meshes ........ 2-83
Figure 46. Pressure along Line for the different time step cases. ........................................... 2-85
Figure 47. Residuals for the 45° lip gate – 30% opening ........................................................ 2-86
Figure 48. Residuals for the Inverted 30° lip gate – 60% opening .......................................... 2-86
1-5

Figure 49. Detail of mesh at bottom Surface of the gate for a Y+ of 300 ................................ 2-87
Figure 50. Detail of mesh at bottom Surface of the gate for a Y+ of 500 ................................ 2-88
Figure 51. Quality report (Skewness) of the meshes for the CFD simulation of the bottom outlet
(45° lip gate and Inverted 30° lip gate)................................................................................... 2-89
Figure 52. Velocity streamline – Inverted 45° lip gate – 50% gate opening .............................. 2-1
Figure 53. Streamlines colored by velocity – Inverted 30° lip gate – 50% opening................... 2-2
Figure 54. Streamlines colored by velocity – Inverted 30° lip gate – 50% opening................... 2-2
Figure 55. Volume fraction of the Inverted 30° lip gate – side view - 30% gate opening .......... 2-3
Figure 56. Volume fraction of the Inverted 30° lip gate – side view - 10% gate opening .......... 2-4
Figure 57.Static Pressure distribution on the lower lip of the slide gate – Inverted 30° lip gate –
90% opening ........................................................................................................................... 2-5
Figure 58.Static Pressure distribution at the domain– Inverted 30° lip gate – 30% opening ..... 2-6
1-6

List of tables
Pág.

Table 1. Discharge coefficients for flat gates .......................................................................... 1-21


Table 2. Model tests for the study of the downpull force (Erbisti, 2004) .................................. 1-22
Table 3. Dimensions of the 45° lip gate .................................................................................. 2-42
Table 4. Dimensions of the inverted 30° lip gate .................................................................... 2-43
Table 5. Parameters of the bottom outlet ............................................................................... 2-44
Table 6. Gate positions for the simulation of the bottom outlet ............................................... 2-45
Table 7. Parameters of the meshes for the 45° lip gate .......................................................... 2-48
Table 8. Parameters of the meshes for the Inverted 30° lip gate ............................................ 2-49
Table 9. Models setup for the simulation ................................................................................ 2-53
Table 10. Boundary conditions for the Gate 1 and Inverted 30° lip gates ............................... 2-54
Table 11. Bottom downpull coefficients for inverted lip slide gates ......................................... 2-72
Table 12. Different configurations for the mesh Independence study ..................................... 2-83
Table 13. Different configurations for the time step independence study ................................ 2-84
Table 14. Main parameters of the simulations for the Y+ Independence study ....................... 2-87
Table 15. Water discharge validation ....................................................................................... 2-1
Table 16. Hydraulic parameters for the analytical calculation of downpull for the 45° lip gate .. 2-2
Table 17. Results of downpull for the 45° lip gate..................................................................... 2-2
Table 18. CFD data obtained from the Inverted 30° lip gate simulations .................................. 2-3
Table 19. Coefficients for the calculation of downpull – Theoretical and CFD .......................... 2-4
1-7

1. Introduction
High-head slide gates are moving hydraulic structures used as flow control devices that operate
under high pressure (over 25 meters of head) and high-flow conditions. This type of gates consists
on a flat metallic structure that slides on side guides embedded in concrete and it’s operated by a
high capacity hydraulic servo-motor. High-head slide gates are mostly used as flow control
devices in irrigation canals, sewage works, spillways and bottom outlets for large dams.

During the opening or closure of the gate against flow, high magnitude operating forces on the
structure of the gate appear, these are hydrodynamic forces caused by the high-flow velocities
under the gate, which create a non-uniform pressure distribution on the surfaces of the gate,
resulting in a downward vertical force; this phenomenon is known as downpull.

The determination of the downpull force is necessary for the design of the gate and the capacity
of the hydraulic servo-motor and related equipment; however, since the magnitude of this force
depends on several variables, the existing analytical methods for its calculation rely on
experimental coefficients. These coefficients depend on the geometry of the gate; but, for certain
exceptional gate geometries these coefficients are unknown, making the existing analytical
methods inadequate for the calculation of the downpull force.

With the late computational development, Computational Fluid Dynamics (CFD) has become a
useful tool and provides a consistent prediction of fluid flow by means of mathematical modeling,
numerical methods and software tools, this allows the analysis and understanding of
hydrodynamic phenomena such as the downpull force on slide gates and represents a practical
aid for the design of hydro-mechanical equipment.

The aim of this research is to study the downpull force on high-head slide gates at a bottom outlet
with different gate lip geometries, using CFD to simulate the closure of the gate against flow. The
data from the simulations is then analyzed to understand the effect of the gate geometry in the
magnitude of the vertical force for different gate positions.

This investigation is a contribution for hydro-mechanical equipment designers and experts that
are interested in the design of hydraulic gates, the downpull phenomenon and the influence of
gate geometry in the magnitude of hydrodynamic forces. It also represents a guide for the
1-8

simulation of a bottom outlet of a large dam using CFD, the ANSYS Fluent software and a Volume
Of Fluid (VOF) multiphase model.

1.1 Description of high-head slide gates and the downpull


Phenomenon
Details of the slide gate, its components, a description of the hydrodynamic forces and downpull
phenomenon can be found below.

1.1.1 High-head slide gates


High-head slide gates are vertical lift hydraulic gates that consist on a flat metallic structure that
slides on metallic side guides and operate against flow with pressures that exceed the 25 meters
of head. This type of hydraulic gate retracts into a bonnet, which is a metallic housing that stores
de gate as it opens. The gate is operated by a hydraulic cylinder called a Servo-motor. This type
of hydraulic gate is mainly used as a flow control device or emergency closure gate in bottom
outlets of large dams. Figure 1 shows a high-head slide gate and its components:

Figure 1. High-head slide gate - Components


1-9

Unlike other types of flat gates, a remarkable characteristic of the slide gates is that less vibrations
are observed in partial openings against flow; this is due to the significant friction forces between
the side guides and the gate seals. This feature is highly desired for bottom outlet gates (Erbisti,
2004) where the water discharge has to be controlled by partially closing the gate as the water
level in the reservoir increases.

The most common type of high-head slide gates are the welded construction gates, in this case
the gate body consists on a flat platen or skin plate which is reinforced by welding vertical and
horizontal beams in the upstream side of the plate. These gates have bronze or brass machined
bars attached to the downstream part of the skin plate, this bars act as seals and are permanently
in contact with the surfaces of the side guides.

1.1.2 Description of hydrodynamic forces and the downpull


phenomenon
Hydrodynamic forces are the loads that result from water flowing against and around a rigid
structural element or system (Quimby, 2007). In this case the structure that is subjected to the
hydrodynamic forces is the gate body.

When the slide gate operates against flow, the pressure at the bottom surface of the gate is
reduced due to the high efflux velocities, while the pressure on the top surface of the gate is only
slightly changed from static condition. This pressure difference generates an unbalanced vertical
force which considerably exceeds the weight of the gate (Naudascher E. , 1964), this
phenomenon is commonly known as downpull. Although downpull force is predominant in the
operation of the gate, in some cases an Uplift force has been observed in gate model tests, this
Uplift force is due to a higher dynamic pressure created by the water flow against the bottom
surface of the gate, however, this phenomenon has only been observed in certain gate geometries
when they are almost completely open (80%-90% opening) for the rest of the gate positions the
downpull force is predominant.

The magnitude of the downpull force on slide gates is particularly high considering the high
pressure operating conditions of the bottom outlets where the slide gate is used as flow control
device. Because of the number of variables involved in determining the downpull force for each
position of the gate, the existing analytical methods used for its calculation must be considered
1-10

approximate (Lewin, 1995). The only reliable methods for the determination of the downpull force
are gate model tests (Erbisti, 2004), or validated numerical simulations using Computational Fluid
Dynamics.

1.2 Thesis outline


This document is divided in 2 chapters, the first chapter is an introduction that includes a general
description of the gate which is the object of study, the scope and aim of the research, a general
description of the hydrodynamic forces and downpull phenomenon, the governing equations of
fluid flow, basic concepts of computational fluid dynamics, the models that will be used for the
simulation, a brief description of the problem, the objectives of the thesis and state of the art.
Finally, the 2nd chapter is an article titled “Analysis of hydrodynamic forces on high-head slide
gates using CFD”, this article summarizes the development, methodology, results and conclusions
of this research.

1.3 Fluid dynamics governing equations


The determination of downpull on high-head slide gates is an engineering problem that is related
to fluid (gas and liquid) dynamics. In general, all the problems related to fluid dynamics are about
solving the governing partial differential equations that represent the mass, momentum and
energy conservation.

These governing equations that describe the fluid dynamics only have analytical solution for
special problems with simplifying assumptions, but for applied engineering problems with
turbulent flows and more complex geometries these governing equations have to be solved
numerically using Computational Fluid Dynamics (CFD).

Computational Fluid Dynamics is the science of predicting fluid flow, heat transfer, mass transfer,
chemical reactions, and related phenomena by solving the mathematical equations described
above which govern these processes using numerical methods (Bakker, 2006).

For applied engineering problems, the domain containing the fluid has to be discretized in order
to approximately solve the governing equations of fluid flow. These equations can be numerically
solved using finite differences, finite volumes or finite elements. In this case the software used for
the simulation of the high-head slide gate, which is ANSYS Fluent, uses the finite volumes
method.
1-11

In this section, governing equations of the fluid flow, numerical methods and the existing analytical
method for the downpull calculation are described.

1.3.1 Continuity equation


Also known as the law of conservation of mass, the continuity equation states that the rate at
which the mass enters, is equivalent to the rate at which it exits the system, plus the accumulation
of mass within the system. This requires:

1 𝐷𝜌 (1)
+ ∇. 𝑢
⃗ =0
𝜌 𝐷𝑡

This equation expresses the general differential form of the principle of conservation of mass;
however, the first term of the equation can be assumed zero for incompressible fluids. A fluid can
be defined as incompressible if its density does not change with pressure, this assumption is valid
for liquids and gases with speeds under 100 m/s (although gases are compressible, for Mach
numbers under 0.3 the change of absolute pressure in flow is small, making the changes in density
negligible). According to the above, Equation (1) can be reduced to Equation (2) for
incompressible flows:

∇. 𝑢
⃗ =0 (2)

1.3.2 Momentum equation


The law of conservation of momentum is obtained from applying Newton’s law of motion to an
infinitesimal element. The net force on the infinitesimal element is a product of the 3 forces
governing the fluid flow, which are the gravitational forces, the viscous forces and the pressure
forces. Considering that the motion of the infinitesimal element requires that the net force on the
element must equal mass times the acceleration of the element, that the only body force
corresponds to gravity and that on Newtonian fluids the deformation rate is proportional to the
shear stress, Equation (3) is obtained:
1-12

𝐷𝑢𝑖 ∂𝜏𝑖𝑗 (3)


𝜌 = + 𝜌𝑔𝑖
𝐷𝑡 ∂𝑥𝑗

The first term of this equation represents the momentum per unit volume, the second term
represents the volume forces per unit volume and the third term represents the surface forces per
unit volume. This equation is called the “Cauchy’s equation of motion” and is valid for fluids and
solids, the difference appears when the stress is related to the deformation (Gómez, 2017). The
term 𝜏𝑖𝑗 in Equation (3) represents the stress tensor. This stress tensor for an incompressible fluid
at motion takes the simple form

𝜏𝑖𝑗 = −𝑝𝛿𝑖𝑗 + 2𝜇𝑒𝑖𝑗 (4)

Equation (4) is known as the generalized law of Stokes for an incompressible fluid. Where 𝑝 is
defined as the mechanical pressure and 𝜇 is defined as the viscosity of the fluid. According to this,
the first term of Equation (4) corresponds to the normal stresses and the second term corresponds
to viscous shear stresses. Additionally, the coefficient of bulk viscosity is defined as

2 (5)
𝜅 = 𝜇 + 𝜇′
3

In principle, the coefficient of bulk viscosity 𝜅 is a measurable quantity; however, extremely large
values of Dp/Dt are necessary in order to make any measurement, such as within shock waves.
Moreover, measurements are inconclusive about the nature of K (Kundu, 2002). According to this,
the Stokes hypothesis considers that this term can be assumed 0, and under this statement the
constitutive equation reduces to Equation (6) for Newtonian fluids

2 (6)
𝜏𝑖𝑗 = − (𝑝 + 𝜇∇. 𝑢𝑖 ) 𝛿𝑖𝑗 + 2𝜇𝑒𝑖𝑗
3
1-13

1.3.3 Navier-Stokes equation


The Navier Stokes equation describes the motion of a viscous fluid. This equation is obtained by
replacing the stress tensor described in Equation (6) in the “Cauchy’s equation of motion”
described in Equation (4) and assuming conservation of mass described in Equation (1). The
partial differential Equation (7) is obtained:

𝐷𝑢𝑖 𝜕𝑝 (7)
𝜌 =− + 𝜌𝑔𝑖 + 𝜇∇2 𝑢𝑖
𝐷𝑡 𝜕𝑥𝑖

where

𝜕 2 𝑢𝑖 𝜕 2 𝜇𝑖 𝜕 2 𝜇𝑖 𝜕 2 𝜇𝑖 (8)
∇2 𝑢𝑖 = = + +
𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥1 2 𝜕𝑥2 2 𝜕𝑥3 2

is the laplacian for 𝑢𝑖 . Equation (7) is only valid for incompressible fluids. The Navier Stokes
equations described above are nonlinear partial differential equations. For the solution of such
equations a numerical approach on a computer is almost the only existing method. For problems
related to the fluid flow such as the hydrodynamic forces on high-head slide gates, Computational
Fluid Dynamics is the method for solving the governing nonlinear partial differential equations
described above.

1.4 Turbulence
In fluid mechanics, turbulence is the most common type of flow found in applied engineering
problems, in a turbulent flow the fluid motion is highly random, unsteady and three dimensional
(Rodi, 1993). Although there is no precise definition for turbulent flow, the following aspects can
be observed:

 Large Reynolds numbers: The turbulent flow only occurs at higher Reynolds numbers, for
example, the transition from laminar to turbulent flow in a pipe occurs at 𝑅𝑒𝐷 ≥ 2300.
1-14

 Randomness: A turbulent flow is random, irregular and chaotic. However, it is deterministic


and is described by the Navier Stokes Equations. (Davidson, 2018)
 Vorticity: It is characterized by the formation of eddies with a wide range of sizes. (Gómez,
2017)
 High Diffusivity: Turbulent flows are characterized for being highly diffusive, this means
that the rate of some physical processes such as thermal conduction, chemical diffusion,
transport of mass and momentum increase in a turbulent flow.
 High Dissipation: Turbulent flows dissipate mechanical energy due to shear stresses at a
large scale. This dissipation creates much larger flow losses which generate internal
energy increasing the temperature of the fluid.
 Fluctuations: In a turbulent flow, it has been observed that the measure of any variable
fluctuates over time.

As mentioned before, the turbulent flow is described by the Navier Stokes Equations, according
to this, for a precise prediction of fluid flow these equations can be used directly to simulate
turbulent flow. This type of flow simulation is known as Direct Numerical Simulation (DNS).
However, the Direct Numerical Simulation requires that the maximum grid spacing is smaller than
the smallest turbulent length scale (Kolmogorov length scale), and the time interval has to be
smaller than the smallest time interval (Kolmogorov time scale), which are the smallest scales in
turbulent flow. This limitation requires a discretization so refined and a time step so small that the
length of the domain and the time for the simulation have to be very small (such as a laboratory
experiment), this makes the Direct Numerical Simulation not applicable for engineering problems,
where the domain can be of considerable size and a transient simulation requires several seconds
of flow to capture the phenomenon of interest.

An alternative method is named a Large-Eddy Simulation (LES). In LES the large eddies of
turbulent flow are computed directly and only small scale motions are modelled, resulting in a
significant reduction of computational cost compared to DNS (Zhiyin, 2018). However, the LES
approach is still computationally expensive for the simulation of a bottom outlet, where the domain
of the simulation is approximately 80 meters long.

A less expensive computational strategy are the RANS-based turbulence models (Reynolds
Averaged Navier Stokes), these turbulence models average the equations of motion over time.
The RANS approach is the most commonly used method in hydraulic engineering applications,
1-15

where resolving the turbulence flow detail is as important, as key flow quantities of engineering
interest (Averaged velocity, averaged pressure distribution, wall shear stress, etc.).

1.5 RANS Equations


As described above, The RANS equations (Reynolds averaged Navier stokes) are time-averaged
equations of motion for turbulent fluid flow. The basis for these equations is the Reynolds
decomposition which proposes the decomposition of variables into their mean part and a
fluctuation from the mean as is shown in Equation (9):

𝑢̃𝑖 = 𝑈𝑖 + 𝑢𝑖 , (9)

𝑝̃ = 𝑃 + 𝑝,

𝑇̃ = 𝑇̅ + 𝑇 ′ .

Where 𝑈𝑖 , 𝑃, 𝑇̅ are the variables of the mean flow, and 𝑢𝑖 , 𝑝, 𝑇′ are the fluctuations from the
variables of the mean flow. These decompositions are later replaced in the equations of motion
to obtain the Reynolds Averaged Navier Stokes Equations (RANS).

1.5.1 RANS-equations based turbulence models


The purpose of the turbulence models based on the RANS equations is to compute the Reynolds
Stresses. This can be achieved by three main categories of RANS based turbulence models:

 Linear eddy viscosity models: In these models the Reynolds stresses are modelled by a
linear relationship with the mean flow straining field as obtained from the RANS equation.

 Nonlinear eddy viscosity models: In these turbulence models an eddy viscosity coefficient
is used to relate the mean turbulence field to the mean velocity field in a nonlinear
relationship.

 Reynolds stress model: These RANS based-turbulence models are based on the exact
Reynolds stress transport equation.
1-16

In this case, for the simulation of the operation of the high-head slide gate, the turbulence model
that will be applied is the K-epsilon model, which is a linear eddy viscosity model.

1.5.2 k-ε turbulence model


The most common RANS-based turbulence model used in applied engineering simulations is the
K-epsilon model. It is a two equation linear eddy viscosity model. This indicates that the model
uses two extra transport equations to characterize the turbulent properties of the flow

The first transported variable is turbulent kinetic energy K and the second transported variable is
the turbulent dissipation 𝜀.

For the turbulent kinetic energy K, which is the variable that determines the energy in the
turbulence:

𝜕 𝜕 𝜕 𝜇𝑡 𝜕𝑘 (10)
(𝜌𝑘) + (𝜌𝑘𝑢𝑖 ) = [(𝜇 + ) ] + 𝑃𝑘 + 𝑃𝑏 − 𝜌𝜖 − 𝑌𝑀 + 𝑆𝑘
𝜕𝑡 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜎𝑘 𝜕𝑥𝑗

and for the turbulent dissipation 𝜖, which is the variable that determines the energy
in the turbulence:

𝜕 𝜕 𝜕 𝜇𝑡 𝜕𝜖 𝜖 𝜖2 (11)
(𝜌𝜖) + (𝜌𝜖𝑢𝑖 ) = [(𝜇 + ) ] + 𝐶1𝜖 (𝑃𝑘 + 𝐶3𝜖 + 𝑃𝑏 ) − 𝐶2𝜖 𝜌 + 𝑆𝜖
𝜕𝑡 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜎𝜖 𝜕𝑥𝑗 𝐾 𝑘

where 𝑢𝑖 is defined as the velocity in the given direction and 𝑢𝑡 is defined as the eddy viscosity.
As indicated in Equation (10) and Equation (11), there are a set of constants defined as 𝜎𝜖 , 𝐶1𝜖 ,
𝐶2𝜖 and 𝜎𝑘 . The values of these constants have been determined by a series of iterations for a
wide range of turbulent flows and are set as default by ANSYS-Fluent.

1.6 ANSYS Fluent Volume of fluid (VOF) model


The VOF (Volume Of Fluid) model is the multiphase model used by ANSYS Fluent for the
simulation of the fluid flow of two or more fluids, based on the assumption that these are
1-17

immiscible, such as the air and water flow downstream of the slide gate. The model is based on
tracking the interface between the two phases of the simulation (Air and water), this is obtained
by solving a single set of momentum equations for the two phases and tracking the volume fraction
of each phase all over the domain.

According to the above, the VOF model is appropriate for the simulation of the operation of the
bottom outlet with a high-head slide gate, since a free surface flow of water is observed
downstream of the gate.

Since the VOF model solves a single set of momentum equations, the velocity field is shared
among the two phases of the simulation. The momentum equation depends on the volume
fractions of all phases which is represented by the density (ρ) and the viscosity (μ) (Belduque,
2015). According to the equation below, any property (𝑥) of each phase can be calculated using
the volume fraction for all phases 𝛼𝑞 at each cell.

𝑥 = ∑ 𝛼𝑞 𝑥𝑞 (12)

1.7 Description of the problem


The magnitude of hydrodynamic forces on high-head slide gates is important due to the high
pressure operating conditions which creates high efflux velocities in the penstock or bottom outlet
where the gate is used as flow control device. In gate model tests, hydrodynamic downpull forces
with magnitudes of 3 times the gate weight has been observed, making hydrodynamic downpull
the main operational force during the opening or closure of the gate against flow.

According to the above, to guarantee a safe and satisfactory operation of the gate, an appropriate
determination of hydrodynamic forces on high-head slide gates is fundamental for the
dimensioning of the gate body and the capacity of the servo-motor and related equipment.

The main existing analytical method for the calculation of hydrodynamic downpull forces is
described by (Naudascher E. , 1964), this method allows the calculation of the hydrodynamic
downpull force for all the gate positions, based on geometric parameters and operating conditions.
1-18

The method relies on experimental coefficients that depend on geometric parameters of the gate.
Figure 2 shows the types of gates that were covered in the research developed by (Naudascher
E. , 1964), therefore, experimental coefficients are known for the analytical calculation of downpull
on these types of gates:

Figure 2. Types of gates with known downpull coefficients (Henriques Da Silva, 2011)

Figure 2 shows the most common bottom geometries on slide gates, however, these geometries
do not cover all the existing gates that are implemented in hydropower projects nowadays. The
Figure 3 shows some possible gate bottom geometries that have been designed lately by gate
manufacturers, which have been applied in several hydroelectric power plants:

Figure 3. Types of gates with unknown downpull coefficients


1-19

The downpull coefficients for the gates shown in Figure 3 are unknown, making the existing
analytical method invalid for the calculation of the hydrodynamic downpull force; therefore, gate
model tests or numerical simulations have to be performed to determine an approximate
magnitude of the hydrodynamic downpull force on these types of geometries.

1.8 Scope and aim of the research


This research is an attempt to determine, for all the gate positions, the hydrodynamic downpull
force on the gate “a” shown in Figure 3, which is not covered in the existing analytical method.
This will be achieved by using Computational Fluid Dynamics to simulate the flow in a bottom
outlet with this type of gate. Finally, the dimensionless downpull coefficients will be estimated for
further analytical calculations.

This will be accomplished by validating a Computational Fluid Dynamics simulation on a typical


slide gate (Case “b” from Figure 2) by comparing the theoretical discharge coefficients and the
ones obtained with the simulation. Also, the downpull obtained with the simulation will be
compared with the one calculated with the existing analytical method described by (Naudascher
E. , 1964). Once validated, the gate of the bottom outlet will be replaced with the gate that is object
of this research (Case “a” from Figure 3).

1.8.1 General objective


To determine the parameters for the analytical calculation of the hydrodynamic forces on a high-
head sliding gate with an inverted gate lip ( = - 30°) using Computational Fluid Dynamics.

1.8.2 Specific objectives


 To develop a numerical flow simulation using Computational Fluid Dynamics of a bottom
outlet with a high-head slide gate.
 To validate the results obtained from Computational Fluid Dynamics simulation using the
existing analytical method for a typical slide gate.
 To determine the downpull coefficients to calculate hydrodynamic downpull from the data
obtained with the Computational Fluid Dynamics simulation of a slide gate with an inverted
gate lip.
1-20

1.9 Theoretical framework


In this chapter, a review of the current investigation in the subject of hydrodynamic forces and
downpull phenomenon on slide gates is described. It includes a detailed description of the
downpull phenomenon on hydraulic gates and the existing analytical calculation method, it also
includes a review of the Computational Fluid approach of several researches about the simulation
of hydraulic gates and bottom outlets.

1.9.1 Water discharge through bottom outlets


The water discharge through a bottom outlet of a large dam is calculated under the consideration
that the gate corresponds to the hydraulic control of the structure. This means that the flow through
the bottom outlet is controlled by the clear opening area of the gate. According to this, the flow
through the bottom outlet is given by Equation (13):

𝑄𝑤𝑎𝑡𝑒𝑟 = 𝐶𝑐 𝐴√2𝑔(𝐻 − 𝐻𝑒 − ℎ) (13)

Where:

𝐶𝑐 : Discharge coefficient (obtained by model tests by the U.S Army Corps of Engineers).

𝐴: Clear opening area of the gate.

𝐻: Total head in the reservoir.

𝐻𝑒 : Entrance head loss.

ℎ: Piezo-metric head at the contracted jet.

The discharge coefficient depends on the gate opening and geometry; it can be determined
through measurements in gate model test, however, in the absence of specific values the following
coefficients may be adopted (Erbisti, 2004), suggested by the U.S Army Corps of Engineers:
1-21

Table 1. Discharge coefficients for flat gates


% of gate opening Discharge coefficient
Cc
10 0,73
20 0,73
30 0,74
40 0,74
50 0,75
60 0,77
70 0,78
80 0,80
90 0,80

For the particular case of this bottom outlet, the head in the reservoir is a constant value of 70
meters of head, the entrance head loss is calculated with the Darcy-Weisbach head loss equation
and the Piezo-metric head at the contracted jet is calculated with Equation (14):

ℎ = 𝐶𝑐 𝑦 + 𝐻𝑑 (14)

Where:

𝐶𝑐 : Discharge coefficient (See Table 1. Discharge coefficients for flat gates).

𝑦: Clear opening height of the gate.

𝐻𝑑 : Depression downstream of the gate.

1.9.2 Hydrodynamic forces on flat gates


When a flat gate is partially open and there is water flow in the tunnel, the hydrostatic balance is
broken (Erbisti, 2004), this results in a non-uniform distribution of the piezo-metric head on the
surfaces of the gate. This distribution of the piezo-metric head is particularly low at the bottom of
the gate where the flow velocities are high, resulting in pressures even below the atmospheric.
1-22

Also, the top surface of the gate is subjected to the net head. This low pressure zone at the bottom
of the gate and the high pressure zone at the top of the gate generates a vertical downward force
known as downpull.

In high-head gates, the downpull is affected not only by the geometry of the gate bottom, but also
by the geometry of the gate bonnet which can induce a higher flow of water passing over the top
of the gate, which can exert a mayor effect on the magnitude of downpull (Lewin, 1995). In a
bottom outlet of a large dam the state of flow downstream the gate is a free surface flow, in which
the space is partially filled with air and water and the pressure is atmospheric or below depending
on the velocity of the flow.

1.9.3 Factors Influencing downpull


It has been proven by experimental data obtained from several hydraulic model investigations that
downpull is a function of the total head upstream and the gate position, Table 1 indicates some of
the scale model tests developed for the analysis of this phenomenon:

Table 2. Model tests for the study of the downpull force (Erbisti, 2004)

From the results obtained in this model tests, it has been observed that the magnitude of downpull
is affected by a series of factors, such as:

1.9.3.1 Location of the gate seals

The location of the gate seals has great effect on the resulting downpull force during the operation
of the flat gates, two arrangements of gate seals can be distinguished:

 Upstream gate seals.


 Downstream gate seals.
1-23

Figure 4 shows the configuration of a flat gate with upstream gate seals

Figure 4. Upstream gate seals configuration

Since the gate seals are located on the upstream side of the gate, the pressure on the top
surface of the gate, during the opening or closure operation, is negligible; subsequently the
resulting top force, which is the main component of downpull, is insignificant. Due to this feature,
upstream seal gates are not subjected to considerable downpull forces.

The other arrangement are the gates with downstream gate seals, Figure 5 shows this
configuration of flat gates:
1-24

Figure 5. Downstream gate seals configuration

In this type of gates, the seals are located on the downstream side of the gate; this allows the
generation of a column of water at the gate shaft generating a piezo metric head on the top surface
of the gate, as shown Figure 5. The force generated by this water column on the area of the top
surface of the gate is the main portion of the resulting downpull force.

1.9.3.2 Aeration

During the operation of the gate, the high velocity flow downstream of the gate generates
pressures below the atmospheric; this drop of pressure can be as low as the vapor pressure of
the water resulting in cavitation and vibration (Erbisti, 2004), which can affect the gate and the
civil works structural integrity.

To solve this problem, aeration vents and pipes are located on the downstream side of the gate;
these pipes are connected to the atmosphere to allow the air demand and the pressure balance
1-25

at the downstream side of the gate. Figure 6 shows the configuration of a bottom outlet of a large
dam and the location of the aeration pipes:

Figure 6. Aeration pipes at a bottom outlet

The lack of aeration pipes creates sub-atmospheric pressures downstream of the gate and
increases the net head acting on the gate (Sagar B. , 1977). This results in larger downpull
forces. The presence of aeration pipe on the downstream side of the gate, as show in Figure 6,
balances this sub atmospheric pressure and reduces the downpull on the gate.

1.9.3.3 Gate bottom geometry

The gate bottom geometry has important influence on the magnitude of the downpull force on the
gate; it is one of the main factors when estimating the hydrodynamic forces and the downpull
effect during the gate closure. Figure 7 shows different configurations of the gate bottom geometry
in flat gates:
1-26

Figure 7. Different gate bottom geometries. A) Flat bottom gate. B) Curved bottom gate. C)
Inclined bottom gate

The experimental data shows that an increase of the projection of the skin plate (parameter “e”
from the gate “A” in Figure 7) reduces the magnitude of downpull due to a reduction of the velocity
near the bottom beam, which increases the local pressure (Erbisti, 2004).

In gates with curved bottom geometry (gate “B” from Figure 7), it is proven by the analysis carried
out by (Naudascher E. , 1964), that a larger radius “r” increases the local pressure at the bottom
surface of the gate. This generates a significant reduction of the downpull force.

For gates with an inclined bottom surface (gate “C” from Figure 7) the sloping angle "𝜃" has an
important influence on the pressure at the bottom surface, it is known that as the sloping angle
increase the downpull is reduced (Erbisti, 2004). The maximum adopted angle "𝜃" is 60° due to
structural and fabrication limitations (Erbisti, 2004).

1.9.3.4 Gate installations

The civil works where the gate is installed have great influence on the magnitude of the downpull
force. Three arrangements of gate installations can be distinguished:

1. Shaft-type gate
2. Bonnet-type gate
3. Face-type gate
1-27

The shaft-type gates are typically used as intake gates for hydroelectric power plants. In this
arrangement the gate is operated in a shaft or a well. In this type of gates, the dimensions and
geometry of the shaft have effect on the column of water during the operation of the gate. This
column of water generates a pressure on the top surface of the gate which is a substantial part of
the downpull force. The Figure 8 shows the typical configuration of an intake gate:

Figure 8. Gate shaft geometry on a typical intake gate

It is known that the downpull force on the gate is affected by the distances between the gate and
the walls of the gate shaft (distances “a” and “b” from the figure above) and on the gate thickness
and the top seal width (Distances “d” and “d’” from the figure above).

The bonnet-type gates are usually used for high-head slide gates at bottom outlets of large dams,
in this type of arrangement the gate contracts into a bonnet and is directly operated by hydraulic
servo-motor. The Figure 9 shows the configuration of a bonnet-type gate:
1-28

Figure 9. Gate shaft geometry on a bottom outlet high-head gate

The gate bonnet is pressurized as the gate closes, equivalently with the water column in the shaft
of the shaft-type gate. This pressure at the gate bonnet generates the vertical downward force on
the top surface of the gate, which is a substantial part of the downpull force.

The face-type gates installed at the intake of the tunnel, are usually used for maintenance
purposes. On this type of gate, the pressure at the top surface is equal to the pressure at the
reservoir. This configuration generates larger downpull forces compared to the two types of gates
indicated in Figures 8 and 9. The configuration of the Face-type gate is shown in Figure 10.
1-29

Figure 10. Face-type gate

It is noted that for the downpull calculation of this type of gate, the downpull coefficient
corresponding to the top geometry of the gate installations is considered maximum (Top downpull
coefficient = 1).

1.9.4 Analytical calculation of downpull


The analytical method used for the calculation of downpull on hydraulic gates was developed by
(Naudascher E. , 1964). This method is based on the geometric parameters of the gate and the
installations to determine a non-dimensional term 𝜅 and the most significant flow and gate
parameters to estimate a magnitude of the downpull force for each gate position. In their article,
the downpull force is determined by Equation (13):

𝑉𝑗2 (15)
𝐷𝑝 = 𝐾𝐵 𝑑 𝛾
2𝑔
1-30

where:

𝐾: Downpull coefficient

𝐵: Width of the gate (between side seals).

𝑑: Gate thickness (See figure 8).

𝛾: Specific weight of water.

𝑉𝑗 : Velocity of the contracted jet under the gate.

𝑔: Gravity.

The downpull force, described by Equation (15), can be decomposed in two forces, the top
downpull force acting on the top surface of the gate and the bottom downpull force acting at the
bottom surface of the gate, Figure 11 shows the decomposition of the downpull force on the
hydraulic gate:

Figure 11. Vertical force decomposition on a flat hydraulic gate


1-31

The sum of the forces 𝐹𝑡 and 𝐹𝑏 shown Figure 11 results in the downpull force acting on the gate.
These forces are determined by Equation (16) and Equation (17) respectively:

Top downpull force:

𝑉𝑗2 (16)
𝐹𝑡 = (𝐾𝑡 )𝐵 𝑑 𝛾
2𝑔

Bottom downpull force:

𝑉𝑗2 (17)
𝐹𝑏 = −(𝐾𝑏 )𝐵 𝑑 𝛾
2𝑔

The resulting downpull force is then given by the Equation (16):

𝑉𝑗2 𝑉𝑗2 𝑉𝑗2 (18)


𝐷𝑝 = (𝐾𝑡 )𝐵 𝑑 𝛾 − (𝐾𝑏 )𝐵 𝑑 𝛾 = (𝐾𝑡 − 𝐾𝑏 )𝐵 𝑑 𝛾
2𝑔 2𝑔 2𝑔

The coefficients 𝐾𝑡 and 𝐾𝑏 are respectively known as top and bottom downpull coefficients. It is
𝑗 𝑉2
noted that the expressions indicated in Equation (14) and Equation (15) include the term 2𝑔

which is equivalent to the net head acting on the surface of dimensions B x d, and the top and
bottom downpull coefficients corresponds to a percentage of the net head acting on the top and
bottom surfaces of the gate.

1.9.4.1 Top downpull coefficient

The top downpull coefficient has values between 0 and 1 and depends on the top geometry of
the gate installations (Distances “a” and “b” from Figure 9). Since the pressure distribution at the
top surface of the gate is generally constant (Naudascher E. , 1964), it can be determined by the
Equation (19) for all gate openings:
1-32

1 (19)
𝐾𝑡 = 2
𝐶 𝑏
1 + (𝐶2 𝑎)
1

Where:

𝐶2 : Discharge coefficient of the top flow at the upstream side of the gate.

𝐶1 : Discharge coefficient of the top flow at the downstream side of the gate.

𝑏: Distance between the top seal of the gate and the surface of the gate shaft.

𝑎: Distance between the upstream side of the gate and the surface of the gate shaft.

It is noted that the expressions 𝐶2 𝑏 and 𝐶1 𝑎 correspond to the discharge coefficient multiplied by
the dimensions “a” and “b” indicated in Figure 9, it can be interpreted by this relation that as the
distance decreases between the gate and the downstream wall of the shaft (Distance “b” from
Figure 9) the top downpull coefficient increases. The analysis of the Equation (19) shows that the
relation between the distances “a” and “b” has a much greater influence on 𝐾𝑡 than the discharge
coefficients 𝐶1 and 𝐶2 , because, while “a” and “b” may take on any values, depending on the
design, practice demonstrates that the variation range of the discharge coefficients is small
(Erbisti, 2004). According to this, for practical purposes, the upstream and downstream top flow
coefficients, respectively known as 𝐶1 and𝐶2 , may be considered equal.

1.9.4.2 Bottom downpull coefficient

One of the main factors that affect the magnitude of the downpull force on flat hydraulic gates is
the bottom geometry; Figure 7 shows the different configurations of the bottom geometry on a flat
gate. Depending on the shape and geometry of this lower lip the pressure distribution at the bottom
surface changes, varying the resulting bottom force 𝐹𝑏 (Figure 10.). The flow along the bottom
surface of the gate is continuously accelerated. Therefore, the velocity at the contracted jet is
designated as a relevant variable (Naudascher E. , 1964) for the calculation of the downpull force.
The Bottom downpull coefficient at a point at the bottom surface of the gate is given by Equation
(20):
1-33

ℎ𝑖 − ℎ (20)
𝐾𝑏 =
𝑉𝑗2
2𝑔

Where:

ℎ𝑖 − ℎ: Difference of the piezo metric heads at a point at the bottom surface of the gate.

𝑉𝑗 : Velocity of the contracted jet

This expression indicates the difference between the piezo-metric heads at a point which is the
𝑗 𝑉2
net head, divided by the velocity head given by2𝑔 . The integration of the Bottom downpull

coefficient described in Equation (18) results in the coefficient along the width and thickness of
the bottom surface of the gate. This is given by:

𝑑 𝐵
1 ℎ𝑖 − ℎ (21)
𝐾𝑏 = ∫ ∫ 𝑑𝐵 𝑑𝑑
𝐵 𝑑 0 0 𝑉𝑗2
2𝑔

Where:

𝐵: Width of the gate.

𝑑: Gate thickness

Since the piezo metric heads ℎ𝑖 − ℎ and velocity at the contracted jet 𝑉𝑗 is variable with the gate
opening position, the Bottom downpull coefficient changes with the gate opening. Figures 12, 13
and 14 show the variation of the Bottom downpull coefficient for different types of lower lip
geometries (Y-Axis) and gate openings (X-Axis) where 𝑦 is the gate opening and 𝑦0 is the
completely open gate height:
1-34

Figure 12. Variation of Bottom downpull Coefficient 𝐾𝑏 with relative gate opening, for gates with
inclined bottom surface (Naudascher E. , 1964)

Figure 13. Variation of Bottom downpull Coefficient 𝐾𝑏 with relative gate opening, for gates with
projected skin plate (Naudascher E. , 1964)
1-35

Figure 14. Variation of Bottom downpull Coefficient 𝐾𝑏 with relative gate opening, for gates with
rounded bottom lip (Naudascher E. , 1964)

The graphs indicating the variation of the downpull coefficient with the gate opening (Figures 12-
14) were obtained by physical model tests developed by (Naudascher E. , 1964), for this test a
two dimensional tunnel-type gate was tested under submerged-flow conditions with a test section
of 3’ by 3’ cross section and a gate thickness of 2, 5”. The distribution of piezo-metric head along
the gate lip was measured at eleven points with piezo-metric holes of 0.033-in. diameter, located
in the bottom surface of the gate, near the center line of the tunnel.

1.9.5 CFD and physical model approaches for the design of hydraulic
gates
Over the past few decades, the CFD approach have progressed to a point where precise solutions
can be found for complex fluid flow problems (Kevin W. Linfield, 2008), which reduces the
incertitude in certain hydraulic phenomena related to the operation and design of hydraulic
structures such as hydraulic gates. On the other side, it is difficult to determine how long physical
models have been used in engineering applications for the optimization of designs with complex
flows. Some of the advantages of the CFD modeling vs the physical modeling are:

- Time.
- Modelling cost.
1-36

- Scale (The CFD approach for gates works with a 1:1 scale vs the physical model which
required a 1:12 or 1:16 scale).

However, for certain problems such as solid particle drop-out or some time dependent problems,
the physical model may be more accurate. Some of the fluid flow phenomena that have been
studied using both computational and physical approach, for the design of hydraulic gates such
as:

- Hydraulic transient analysis during the closure of hydraulic gates.


- Downpull force on high-head slide gates and intake fixed-wheel gates.
- Vortex shedding in the free surface flow adjacent to the gate.
- Hydrodynamic torque on large butterfly valves.
- Operation of spillway segment gates.
- Operation of bottom outlet segment gates.

For the operation of hydraulic gates, both approaches are valid and have been applied for the
design. For example, Figure 15 shows a scale model for a high-head segment gate, developed
by the Universidad Nacional de Colombia:

Figure 15. Scale model of a bottom outlet with high-head segment gates. (Calderón Villegas,
2016)
1-37

For the exact same bottom outlet, a CFD model was developed and validated with the results
obtained from the physical model.

Figure 16. CFD simulation of the same bottom outlet – Velocity contours (IMPSA, 2016)

Other examples of researches using the CFD approach for the analysis of fluid flow problems in
hydraulic gates are:

- (Henriques Da Silva, 2011) Develops an analytical method for the calculation of downpull
using CFD as the validation method.
- (Almaini, 2010) Determines the downpull coefficients on intake gates using CFD and
compares the CFD results with experimental data.
- (Calderón Villegas, 2016) Determines the air demand during the operation of a high-head
segment gate at a bottom outlet using the CFD approach and the VOF multiphase model.
The results obtained with CFD are validated using analytical methods and experimental
data from the scale model shown in Figure 20.
- (Uysal, 2014) Determines the bottom downpull coefficients using a CFD model and
validates the results with the experimental coefficients determined by (Naudascher E. ,
1964).
2-38

2. Analysis of hydrodynamic forces on high-head


slide gates using CFD

Abstract
Coefficients that can be applied to current analytical expressions that predict the hydrodynamic
forces and, particularly, the downpull phenomenon on an inverted 30° lip gate at a bottom outlet
of a dam, were determined from Computational Fluid Dynamics (CFD) simulations. Two slide
gates were simulated: A typical 45° lip gate used for model validation and the inverted 30° lip gate.
These simulations made use of the VOF (Volume of Fluid) multiphase model, embedded in the
commercial software ANSYS FLUENT, to simulate the air and water interaction in the free surface
flow downstream of the gate. The CFD predictions were successfully validated against those by
analytical expressions that related the water discharge through the gate and the gate opening at
a fixed reservoir head. The predicted downpull was in agreement with that obtained from analytical
expression for the well-studied 45° lip gate. Through the validated CFD simulation it was possible
to determine the coefficients for the analytical calculation of downpull on the inverted 30° gate,
which were not available in the state-of-the-art analytical expressions. The methodology here
described can easily be applied to different gate geometries for which design coefficients are not
available.
2-39

2.1 Introduction
High-head slide gates are vertical lift hydraulic gates that consist on a flat metallic structure that
slides on metallic side guides and operate against flow with pressures that exceed the 25 meters
of head. This type of hydraulic gate retracts into a bonnet, which is a metallic housing that stores
the gate as it opens. The gate is operated by a hydraulic cylinder called a Servo-motor. This type
of hydraulic gate is mainly used as a flow control device or emergency closure gate in bottom
outlets of large dams. Figure 17 shows a high-head slide gate and its components:

Figure 17. High-head slide gate – Components

During the opening or closure of the gate against flow, high magnitude operating forces on the
structure of the gate appear. These are hydrodynamic forces caused by the high-flow velocities
2-40

under the gate, which creates a non-uniform pressure distribution on the surfaces of the gate,
resulting in a downward vertical force; this phenomenon is known as downpull.

In gate model tests, hydrodynamic downpull forces with magnitudes of 3 times the gate weight
has been observed (Erbisti, 2004), making hydrodynamic downpull the main operational force
during the opening or closure of the gate against flow. According to this, to guarantee a safe and
satisfactory operation of the gate, an appropriate determination of hydrodynamic forces on high-
head slide gates is fundamental for the dimensioning of the gate body, the capacity of the servo-
motor and related equipment.

The main existing analytical method for the calculation of hydrodynamic downpull forces is
described by (Naudascher E. , 1964). This method allows the calculation of the hydrodynamic
downpull force for all the gate positions, based on geometric parameters and operating conditions.
The method relies on experimental coefficients that depend on geometric parameters of the gate.
Figure 18 shows the types of gates that were covered in the research developed by (Naudascher
E. , 1964), therefore, experimental coefficients are known for the analytical calculation of downpull
on these types of gates:

Figure 18. Types of gates with known downpull coefficients (Henriques Da Silva, 2011)

However, these geometries do not cover all the existing gates that are implemented in hydropower
projects nowadays. Figure 19 shows some possible gate bottom geometries that have been
designed lately by gate manufacturers, which have been applied in several hydroelectric power
plants in the world:
2-41

Figure 19. Types of gates with unknown downpull coefficients

The downpull coefficients for the gates shown in figure 19 are unknown, making the existing
analytical method invalid for the calculation of the hydrodynamic downpull force; therefore, gate
model tests or numerical simulations have to be performed to determine an approximate
magnitude of the hydrodynamic downpull force on these types of geometries.

With the late computational development, Computational Fluid Dynamics (CFD) has become a
useful tool and provides a consistent prediction of fluid flow by means of mathematical modeling,
numerical methods and software tools, this allows the analysis and understanding of
hydrodynamic phenomena such as the downpull force on slide gates and represents a practical
aid for the design of hydro-mechanical equipment.

This research is an attempt to determine, for all the gate positions, the hydrodynamic downpull
force on the gate with the geometry “a” shown in Figure 19, which is not covered in the existing
analytical method developed by (Naudascher E. , 1964). This will be achieved by using
Computational Fluid Dynamics to simulate the flow in a bottom outlet with this type of gate. Finally,
the dimensionless downpull coefficients will be estimated on this particular gate so that simple
analytical calculations become available to the gate designer that does not have the time or
experience that a CFD analysis demands.
2-42

2.2 Methodology

2.2.1 Geometry and mesh


The CFD simulations were carried out using the commercial solving ANSYS-FLUENT software.
This solves the Reynolds averaged Navier-Stokes equations in 3D. The VOF (Volume of Fluid)
model represented the two-phase flow in the simulation.
The CFD model was validated by comparing the results of the simulation with those of calculations
following the analytical method described in (Naudascher E. , 1964) for the prediction of the
downpull forces. Table 3 presents the characteristics of the gate corresponding to the 45° lip gate:

Table 3. Dimensions of the 45° lip gate

Figure 20. 45° lip gate geometry

The CFD results were validated, the analysis of a second gate geometry with a similar
configuration as that used in the validation was undertaken. The gate subject of analysis had a
different geometry. Table 4 presents the geometry of the gate used for analysis.
2-43

Table 4. Dimensions of the inverted 30° lip gate

Figure 21. Inverted 30° Gate


geometry

The hydraulic structure simulated using CFD is a bottom outlet for a large dam that involves a
tunnel that derives in two square-section conduits, each conduit is equipped with an emergency
slide gate and a main high-head segment gate. This configuration is commonly used for bottom
outlets, where the water discharge has to be regulated with a segment gate, and the slide gate is
used for emergency and maintenance of the segment gate. Figure 22 shows the configuration of
the bottom outlet:
2-44

Figure 22. Side and top view of the bottom outlet for the simulation

As indicated in Figure 22, the bottom outlet is composed by the following elements:

 Upstream steel penstock.


 Slide gate.
 Segment gate.
 Downstream steel penstock.

The upstream steel penstock consists on a metallic pressure vessel embedded in concrete,
designed for standing the maximum reservoir pressure when the gates are closed. The upstream
gate is known as a high-head slide gate and the downstream gate is known as a high-head
segment gate. The downstream steel penstock is designed for open-channel flow. In Table 5, the
main parameters of the bottom outlet are indicated:

Table 5. Parameters of the bottom outlet


Variable Value
Upstream Penstock length 23 meters
2-45

Variable Value
Downstream Penstock length 12 meters
Gate guides and steel liner length 5 meters
Slide gate clear opening 3,9 meters high x 3,0 meters wide
Segment gate clear opening 3,9 meters high x 3,0 meters wide
Max. flow 650 m³/s
Max. pressure 70 meters of head

This configuration of bottom outlet is typically used for high-flow and high-head hydraulic
structures, where the flow has to be regulated by modulating the position of the segment gate
depending on the level of water at the reservoir. However, since the focus of this research are the
hydrodynamic forces on the slide gate, the segment gate will remain completely open (as shown
in Figure 22) while the slide gate will move from completely open (100% opening) to almost
completely closed (10% opening). In Table 6, the different gate positions for the simulation of the
bottom outlet are described:

Table 6. Gate positions for the simulation of the bottom outlet


Simulation # Slide gate position Segment gate
(% opening) position (% opening)
Simulation 1 90 %
Simulation 2 80 %
Simulation 3 70 %
Simulation 4 60 %
Simulation 5 50 % 100 %
Simulation 6 40 %
Simulation 7 30 %
Simulation 8 20 %
Simulation 9 10 %

These gate configurations are implemented for both the validation and Inverted 30° lip gates.

The geometry was modeled using the Space Claim CAD software, available with ANSYS Fluent.
This geometry represents the fluids (air and water) contained in the bottom outlet.
2-46

Figure 23. 3D Geometry of the fluid contained in the bottom outlet

As shown Figure 23, the slot represents the slide gate; this slot will be vertically displaced to
simulate the different positions of the slide gate to achieve the 9 simulations described in Table 6.
The air intake indicated downstream of the slide gate corresponds to the space between the
segment gate arms, this allows the aeration of the free surface flow downstream of the gate. This
air enters through the access tunnel which is connected to atmosphere. This phenomenon is
represented in Figure 24:

Figure 24. Aeration phenomenon at the bottom outlet


2-47

Figure 23. 3D Geometry of the fluid contained in the bottom outlet, makes evidence the existence
of a symmetry plane. To reduce the computational cost of the CFD model, that symmetry plane
was implemented in the simulation of the bottom outlet, as Figure 25 illustrates:

Figure 25. Symmetry plane for the simulation of the bottom outlet

According to this, the final geometry for the CFD simulation of the bottom outlet is indicated in
Figure 26:
2-48

Figure 26. Final geometry for the CFD simulation of the bottom outlet

The mesh was developed using the ANSYS Fluent meshing software. The geometry was
partitioned in blocks to allow a better refinement and control of the meshing process. For each of
the simulations of the bottom outlet with a high-head slide gate there is a different mesh, in Table
7 and Table 8 the number of elements, average cell size and quality report (Max. Skewness) of
the 45° lip gate and the 30°inverted lip gate geometries:

Table 7. Parameters of the meshes for the 45° lip gate


Gate opening Number of Average Cell
Simulation # Max. Skewness
(%) elements size (m)
1 100 1´812.650 0,10 0,48
2 90 1´956.321 0,10 0,47
3 80 2´050.547 0,10 0,44
4 70 2´221.892 0,10 0,45
5 60 2´446.276 0,10 0,48
6 40 2’641.587 0,10 0,47
2-49

7 30 2’859.241 0,10 0,48


8 20 3’045.215 0,10 0,46
9 10 3’214.357 0,10 0,48

Table 8. Parameters of the meshes for the Inverted 30° lip gate
Gate opening Number of Average Cell
Simulation # Max. Skewness
(%) elements size (m)
10 100 1´812.650 0,10 0,46
11 90 1´956.321 0,10 0,45
12 80 2´050.547 0,10 0,43
13 70 2´221.892 0,10 0,48
14 60 2´446.276 0,10 0,46
15 40 2’641.587 0,10 0,44
16 30 2’859.241 0,10 0,45
17 20 3’045.215 0,10 0,47
18 10 3’214.357 0,10 0,46

One of the most important quality parameters for meshing is the skewness, it is defined as the
difference between the shape of the actual cell and the shape of a rectangular cell of equivalent
volume. The value of the skewness is obtained using the normalized angle deviation method,
which compares the largest angle in the cell with the 90° desired angle, being 0 the ideal skewness
value. Skewness values over 0, 95 indicate deformed cells and may lead to convergence
difficulties in the simulation. Although skewness quality was the main parameter for the meshing,
the orthogonal quality and aspect ratio quality parameters were also considered and kept in the
recommended ranges values indicated in (Seo, 2014).

Appendix E indicates the skewness quality report of the meshes shown in the tables above.

Figure 27 presents an example of the mesh for the simulation 5 for the 45° lip gate. Similar meshes
were obtained for the other 18 simulations. While a coarse mesh was implemented towards the
exit of the canal, the most refined mesh was constructed near the slide gate. A more complete
description of the mesh around the gate is presented below:
2-50

Figure 27. Side view of the mesh for the 45° lip gate

2.2.1.1 Near wall modeling strategy

In this research a wall function approach was implemented to represent the interaction of the fluid
with a wall. It is known that this method is a good approximation for flows with high pressure
gradients (Gerasimov, 2006), such as those present on the surfaces of the slide gate, and allows
the use of a relatively coarse mesh in the near wall region, which reduces the computational cost
of the CFD model.

To estimate the height of the first cell at the near-wall regions, the Reynolds number was
calculated according to Equation (22):

𝜌𝑈. 𝐿 (22)
𝑅𝑒 =
𝜇

Where 𝑅𝑒 is the Reynolds number, 𝜌 is the density of water, 𝐿 is the characteristic length
(distance between the bottom surface of the tunnel and the gate), 𝜇 is the dynamic viscosity of
water and 𝑈 is the freestream velocity
2-51

The next step is to determine an appropriate Y+ value. For standard wall functions, each wall-
adjacent cell centroid may be located within a Y+ value of 300 (Gerasimov, 2006) . According
to this, the first cell height is determined using Equation (21):

𝑌+𝜇 (23)
∆s =
𝑈𝑓 𝜌

Where the 𝑌 + is the target value indicated above and 𝑈𝑓 corresponds to the friction velocity. The
friction velocity is determined by Equation (22):

𝜏𝑤 (24)
𝑈𝑓 = √
𝜌

Where 𝜏𝑤 is the wall shear stress and 𝜌 is the density of the fluid. The wall shear stress is
determined using Equation (23):

1 (25)
𝜏𝑤 = . 𝐶𝑓 . 𝜌. 𝑈 2
2

Where 𝐶𝑓 is defined as the skin friction coefficient, 𝜌 is the density of the fluid and 𝑈 is the mean
flow velocity. To calculate 𝐶𝑓 empirical results have been used to estimate its value, using
Equation (24), which can be applied for internal flows:

(26)
−0,25
𝐶𝑓 = 0,079. 𝑅𝑒

Adopting this procedure and Equations (20) to (24) the value of the first cell height can be
determined. For this case, a first cell height of at least 1 cm has to be adopted at the walls of
the gate to obtain the desired Y+ value of 300. In Figure 28, the detail of the mesh at the walls
of the slide gate, is shown:
2-52

Figure 28. Detail of the mesh at the bottom surfaces of the 45° lip gate

Figure 29. Detail of the mesh at the top surfaces of the 45° lip gate
2-53

However, since the downpull force is predominantly affected by the pressure gradients instead of
the viscous shear stresses at the surfaces of the gate, an increase of the Y+ values to 500 does
not show a significant difference at the resulting downpull force for a given position of the gate. In
the Appendix D a Y+ independence study is described.

2.2.1.2 Models

For the free surface flow downstream of the slide gate of the bottom outlet, where the fluids are
immiscible (Air and water), and the length scale is larger than the computational mesh, the most
appropriate multiphase model is the VOF (Volume of Fluid) model. In Table 9, the parameters for
the setup of the models for the simulations are defined:

Table 9. Models setup for the simulation

Parameter Value
Multiphase Model VOF
Turbulence Model K-epsilon standard
Scheme Explicit
Phase 1 Water
Phase 2 Air

The selection of K-epsilon is motivated, among other things, by the fact that it is more
computationally efficient than other turbulence models such as the K-omega (Wasserman, 2016).
Furthermore, the results obtained using second-order schemes for all the variables are fairly
accurate (Calderón Villegas, 2016) and that previous studies (Calderón Villegas, 2016) found
negligible differences between the results obtained with the K-epsilon model and those obtained
with the K-epsilon RNG and K-omega SST models on a bottom outlet with similar geometry and
operating conditions as those of the present study.

The Volume of Fluid (VOF) model was chosen to represent the water-air interface. Particularly,
the VOF explicit scheme was selected considering that it has a sharper interface and a more
accurate solution than the implicit scheme (Seo, 2014). This scheme is applicable for this
simulation considering that the skewness values (under 0,5) and that the phases are not
2-54

compressible (Mach numbers for the air are estimated under 0,35). It is also noted that although
the CFD simulations are aimed at understanding the steady-state behavior of the gate. However,
the explicit VOF scheme only applies for transient simulations; therefore, a transient formulation
will be performed until a pseudo-steady state of flow is achieved.

Given that the interest of the simulation is the resulting downpull force on the gate, the “Implicit
body force” option for the VOF model was activated, this option allows the handling of
hydrodynamic forces in a robust numerical manner (Seo, 2014), such as the ones presented at
the surfaces of the gate.

2.2.2 Boundary Conditions


Table 10 presents the boundary conditions for both CFD models:

Table 10. Boundary conditions for the Gate 1 and Inverted 30° lip gates

Boundary Type Value Volume fraction


condition
Water inlet Pressure inlet 70 meters of head Water = 1
Air = 0
Air inlet Pressure inlet 0 meters of head Water = 0
(atmospheric pressure) Air = 1
Outlet Pressure outlet 0 meters of head From neighboring cell
(atmospheric pressure)
Symmetry plane Symmetry - -

It is important to annotate that although there is an expected air velocity at the Air Inlet, the area
of this inlet is almost the same as the access tunnel and the pressure at this point is estimated as
atmospheric. Figure 30 depicts in the existing geometry, the configuration of the boundary
conditions listed in table 10.
2-55

Figure 30. Boundary conditions for the CFD models

The reason to determine a pressure inlet – pressure outlet combination of boundary conditions is
that the flow through the bottom outlet depends on the position of the gate, according to this, a
velocity inlet – pressure outlet combination is more difficult to determine, considering that the
calculation of water flow through the bottom outlet (for the definition of a velocity inlet) would
require the use of experimental discharge coefficients for each position of the gate.

According to this, the same boundary conditions of pressure inlet – pressure outlet were adopted
as a strategy to validate the simulations. For a defined gate opening, the simulation would yield a
specific water flow that can be compared with those estimated from empirical correlations. The
turbulence intensity or turbulence level at the inlets and outlets of the model was defined as 5%
given the turbulent character of the flow.

2.2.2.1 Methods and controls

For the solution of the governing fluid flow equations with the models indicated above, a series of
numerical methods and solver settings were adopted for the Pressure-Velocity coupling, spatial
discretization and for the transient formulation, this setup was defined in accordance to (Seo,
2014), as the best practices for the explicit VOF scheme.
2-56

The simulations were conducted with second-order discretization for all the balance equations.
The pressure-velocity coupling scheme was PISO. The pressure interpolation scheme was
PRESTO! and the gradients were computed based on a Least Squares Cell Based approach. All
these settings are standard for CFD use and did not have a significant influence in the solution.

To guarantee convergence, the global Courant number was 2 and the time step was originally set
at 1x10-3 seconds, but an option of variable time step based on Courant number was activated.
The number of iterations per time-step was set to 50, however, when a pseudo-steady state was
obtained (residuals below 1x10-5) 15 iterations per time step were enough to guarantee
convergence.

The initial conditions, regarding volume fraction, where those in Figure 31. This is air was only
present downstream from the gate and at a height equal or higher than the gate opening. In Figure
31 illustrates the case for a 60% gate opening but similar initial conditions were established for
other gate openings.

Figure 31. Volume fraction initialization for the simulation – 45° lip gate – 60% opening (Phase 1
corresponds to water)

In relation to the method for the reconstruction of the free surface flow downstream of the slide
gate, the applied method is the Geo-Reconstruct scheme, this was determined considering that it
is the most accurate method for the reconstruction of the air-water interface, compared to the
compressive and CICSAM methods (Seo, 2014). Although the VOF explicit scheme combined
2-57

with a compressive method may result in a sharp interface reconstruction, even comparable to
the Geo-Reconstruct method, the computational cost is not particularly affected by this setup (5%
more computational time with the Geo- Reconstruct).

Given the dynamic character of the simulations, it is necessary to guarantee that the results are
obtained under pseudo-steady state conditions. This is, that the variation of important variables
such as the downpull force and mass balance remain stable. Figure 32 indicates the variation of
the resulting downpull with time (sum of the reports at all surfaces of the gate) and the mass
balance between the inlets and outlets of the system. In the first second the CFD simulation is
clearly unstable as the value of the downpull varies. However, after 2.9 seconds the value remains
stable (For both downpull and mass balance report). To avoid the unrealistic behavior at low times,
caused by the fact that the solver needs time to obtain convergence, in this research all values
are reported once pseudo-steady state was achieved (i.e. after 2.9 s in Figure 32. Pseudo steady
state analysis) and 0.5 seconds of simulation under pseudo-steady state.

Figure 32. Pseudo steady state analysis of the downpull force – 45° lip gate – 40% gate opening
2-58

Figure 33. Pseudo steady state analysis of the mass balance – 45° lip gate – 50% opening

2.2.3 Results

2.2.3.1 Characterization of the hydrodynamic field

Figure 34 shows the isocontours for the different phases in the simulation in a plane along the
center section of the bottom outlet. Red is related to water, blue to air and green to the interphase.
The figure shows results for different gate openings: 80%, 40% and 20%. The CFD simulation
captures the main characteristics of the flow such as the existence of a contracted jet downstream
of the gate as well as the air intake that is responsible for a clear two-phase flow.
2-59

Figure 34. Volume fraction of the 45° lip gate – side view – Mid plane

Figure 35 indicates the static pressure contours of the 45° gate lip for four different gate openings,
it can be observed that as the gate closes the pressure at the bonnet of the gate (see Figure
17Figure 17. High-head slide gate – Components) increases. However, it is interesting to point
out that the pressure at the bonnet remains stable from the 60% opening position to the completely
closed gate, This confirms the hypothesis indicated by (Sagar B. , 1977) Where he points out that
the pressure at the bonnet or gate shaft is slightly changed from static condition in most of the
gate closing operation.
2-60

Figure 35. Pressure contours for different gate openings

Another interesting annotation is that downstream of the slide gate, there is a very low pressure
zone (dark blue contours downstream) which decreases as the gate closes. this is caused by the
2-61

high efflux velocities and the section change in the geometry at this point. To avoid this
phenomenon, some bottom outlets are provided with lower aeration pipes. It is important to point
out that the tendency of the vacuum or low pressure zone downstream of the gate, which
decreases as the gate closes, confirms the theory indicated in (Levin, 1968) where the maximum
air demand of the gate occurs at higher gate openings (100%-60%). It is noted that although the
lack of aeration may increase the downpull force on the gate, in this particular case, the low
pressure zone is located far from the gate (8 meters downstream), and does not affect the
downpull results.

Figure 36 illustrates, as velocity vectors, the flows of air and water for two different gate openings
of the 45° lip gate. The simulation indicates that the high velocity free surface flow downstream of
the gate, generates an air demand, which is satisfied through the upper section of the segment
gate, located downstream of the slide gate (see Figure 24. Aeration phenomenon at the bottom
outlet). (Calderón Villegas, 2016) Studied the phenomena of air entrainment in a bottom outlet
with a high head segment gate of similar dimensions and operating conditions, using a CFD
model, which is validated with both analytical and experimental approaches. According to this
study, the velocity of the air flow demanded by the bottom outlet may exceed 50 m/s, which is
considerably higher than the velocity of water, this phenomenon is clearly observed in Figure 36.
It can be noted that the air demand for the 70% opening is higher than the 90%, which is expected
since the Froude number that affects the air-water flow relation is higher in the 70% gate opening,
due to the higher water flow velocities.
2-62

Figure 36. Velocity vectors for the mixture – 45° lip gate – 50% opening

It is interesting to annotate that the velocity in the upper section of the gate, which is known as
the gate bonnet, is close to 0 m/s (dark blue vectors in the upper part of the gate), this lower
velocity confirms the higher static pressure at the gate bonnet which is shown in Figure 35.
Pressure contours for different gate openings.
2-63

In Figure 37, the streamlines colored by the velocity of flow, are indicated. This streamlines
describe the direction in which a particle of water will travel through the domain, 25 points at the
water inlet and 10 points at the air inlet were defined as starting points for the streamlines. It can
be observed that the velocity at the upstream part of the bottom outlet is reduced as the gate
closes, this is due to the lower water flow, however, the velocity downstream of the gate does not
change significantly, this phenomenon is presented since the water flow downstream is a free
surface flow regardless of the position of the gate.

Figure 37. Streamlines colored by velocity – 45° lip gate – 80% and 50% gate openings
2-64

The rest of this section is devoted to the evaluation of the hydrodynamics for inverted 30° lip gate.
Appendix F shows the phase contours and other results that are very similar to those obtained for
45° lip gate and, therefore, are not commented in this section. The discussion below focuses on
differences between both gates.

As discussed above, the downpull force applied on high-head slide gates is the result of an
unbalanced static pressure distribution on the surfaces of the gate, due to the high flow velocity
at the gate lip, according to this, the geometry of the gate lip has a very important influence on the
pressure distribution at the structure of the gate. Figure 38. Static pressure isocontours
comparison – 50% gate opening – 45° and inverted 30° gates compares the static pressure
distribution on both gates which are object of study.
2-65

Figure 38. Static pressure isocontours comparison – 50% gate opening – 45° and inverted 30°
gates

It can be observed that the pressure at the gate bonnet is not affected by the geometry and
orientation of the gate lip, however, the pressure distribution at the bottom surface of the inverted
30° lip gate is considerably lower compared to the 45° lip gate. This pressure difference between
the top and bottom surfaces generates downpull, and in accordance to this, the inverted 30° lip
gate will be subjected to larger downpull forces than the 45° lip gate. Another interesting
2-66

annotation is that the vacuum downstream (dark blue contours) appear to be higher for the
inverted 30° lip gate, this will produce a higher air demand during the operation of the gate.

2.3 Validation
Validation of the simulations for both gates was conducted by comparing analytical calculations
based on state-of-the-art correlations and the CFD predictions. Section Water discharge2.3.1
Water discharge describes the analytical expressions that allow the calculation of water discharge
through both gates. This variable, therefore, was selected to validate the results in both cases. A
correct prediction of the discharge should validate the CFD simulation in general.

A more specific validation of the downpull calculation was also carried out by comparing
predictions of downpull for the inverted 30° lip gate, for which a theoretical expression exists. A
successful validation of the downpull calculation for a 45° lip gate should give confidence on the
CFD simulation for the inverted 30° lip gate, that is the subject of this research.

2.3.1 Water discharge


The analytical results of the water discharge described in Section 2.3.1 were compared to those
obtained with CFD for the 45° lip and the inverted 30° lip gate. A pressure inlet condition, defined
as 70 meters of head was applied for all the simulations in order to validate the discharge.
validation compares the analytical and CFD predictions of water discharge for all positions of the
45° and inverted 30° gate (see Table 6. Gate positions for the simulation of the bottom outlet) .
Agreement is good as the highest difference is only 3% and the trend is correctly matched.
Appendix F presents some of the results of the CFD simulations that were used to obtain the data
in Figure 39 as well as the calculation details for the analytical results. This validation gives
confidence on the ability of the software to predict the flow through the gate.
2-67

350

300

250
Water flow (m³/s)

200
Analitycal
CFD 45° lip gate
150 CFD Inverted 30° lip gate

100

50

0
10% 20% 30% 40% 50% 60% 70% 80% 90%
Gate opening %

Figure 39. Analytical vs CFD water discharge validation

As indicated in Figure 39. Analytical vs CFD water discharge validation, the difference between
the values of the theoretical discharge and the CFD models is acceptable (1%-3% difference),
however, it is noted that the water discharge for the Inverted 30° lip gate is lower, indicating that
the geometry of the gate for this particular case may have lower discharge coefficients compared
to the ones indicated by the U.S Army Corps of Engineers.

2.3.2 Downpull
The top, bottom and resulting downpull forces were analytically computed and compared to the
results obtained with the CFD 45° lip gate, which is the typical gate geometry, to validate the
downpull results from the computational model.
2-68

Figure 40 indicates the top downpull force which is applied on the top surface of the gate (see
Figure 11. Vertical force decomposition on a flat hydraulic gate) for the analytical and CFD models,
it is important to annotate that the pressure at the gate bonnet is responsible for the top downpull
force, and since the gate bonnet pressurizes as the gate closes (Sagar B. , 1977), higher
pressures are expected for lower gate openings (60% - 10%). In this gate opening range the
pressure at the bonnet remains constant and is equivalent to the pressure upstream. This
phenomenon is particularly evident in Figure 35 and Figure 40.

1400

1200

1000
Force (kN)

800

600

400

200

0
10% 20% 30% 40% 50% 60% 70% 80% 90%
Gate opening (%)

Top Downpull force CFD Top Downpull force Analytical

Figure 40. Top downpull force – Analytical vs CFD 45° lip gate

Figure 41 indicates the bottom downpull force which is applied at the lower gate lip or bottom
surface, according to the graph, the force at the bottom surface is variable in the 400 – 1100 kN
range, while the force at the top surface is variable in the 800 – 1250 kN range, this magnitude
difference between the two forces makes evident that the flow of water generates a downward
vertical force, which is called downpull.
2-69

1200

1000

800
Force (kN)

600

400

200

0
10% 20% 30% 40% 50% 60% 70% 80% 90%
Gate opening (%)

Bottom Downpull force CFD Bottom Downpull force Analytical

Figure 41. Bottom downpull force – Analytical vs CFD 45° lip gate

Figure 42. indicates the resulting downpull force acting on the gate for different gate openings, for
both analytical and CFD models of the 45° lip gate. This result is obtained from the subtraction of
the bottom downpull force from the top downpull force, which is the resulting vertical hydrodynamic
force.
2-70

600

500

400
Force (kN)

300

200

100

0
10% 20% 30% 40% 50% 60% 70% 80% 90%
Gate opening (%)

Resulting Downpull force CFD Resulting Downpull force Analytical

Figure 42. Resulting downpull force – Analytical vs CFD 45° lip gate

According to Figure 42, the downpull force obtained from the CFD 45° lip gate is lower compared
to that calculated from the analytical method, however, it can be observed that the tendency of
the curve is similar. Understanding why in Figure 42 the downpull predicted by CFD is lower than
that estimated by the analytical correlation required explaining the differences in both top and
bottom forces calculated by both models. While in Figure 40 CFD predicts a higher Ft, which
should result in a higher downpull prediction, the value of Fb predicted by CFD in Figure 41 is
significantly higher than that predicted by the analytical model. This difference between the two
models may be due to the following factors

 Scale factors inherent to the physical model test, from which the downpull coefficients
were obtained.
 The piezometer distribution of the model test does not cover all the bottom surface of the
gate model test, and the measurements are assumed equal along the gate width.
 Numerical approximations and simplifications from the CFD solvers and models.
2-71

Although it is complicated to explain with certainty the reason for the difference between the
downpull predictions by the analytical and CFD models, the results in Figure 42 give confidence
of the ability of CFD to estimate the downpull of gate geometries for which experimental studies
are not yet available. The discussion above suggests that both models, the analytical and the
CFD, have certain shortcomes that may be responsible for the difference. Determine which is
more correct can probably only be accomplished by comparison with careful experimental data,
something that is out of the scope of this thesis. The main factor for this difference is that the
forces at the bottom surface of the gate are larger for the CFD compared to the analytical method.

2.3.3 Bottom downpull coefficients for inverted lip gate


According to the description of the problem, described in 2.1, the purpose of this research was to
determine the bottom downpull coefficients for a particular gate geometry, which was not covered
in the existing analytical method developed by (Naudascher E. , 1964). This was accomplished
by carrying out a CFD simulation of the inverted 30° lip gate. From the data in the CFD simulation
the coefficients can be extracted using Equation (15),

𝐹𝑏 (27)
𝐾𝑏 =
𝑉𝑗2
𝐵 𝑑 𝛾 2𝑔

The data corresponding to the velocity of the contracted jet under the gate “𝑉𝑗 ” was obtained from
the area weighted average of the flow of water divided by the height of the contracted jet under
the gate which is determined in the Volume of Fluid reports. The force “𝐹𝑏 ” was obtained from a
surface force report at the bottom gate lip. The other variables 𝐵, 𝑑, 𝛾 indicated in the Equation
(27) correspond to the width of the gate, the thickness of the gate and the specific weight of water
respectively. The results for the bottom downpull coefficients on inverted lip slide are indicated in
Table 11. Bottom downpull coefficients for inverted lip slide gates:
2-72

Table 11. Bottom downpull coefficients for inverted lip slide gates
Gate opening (%) Bottom downpull coefficient 𝑲𝒃 (m/m)

100 0,00
90 -0,19
80 -0,26
70 -0,30
60 -0,29
50 -0,27
40 -0,49
30 -0,49
20 -0,44
10 -0,20
0 0,00

It is noted that the obtained Bottom downpull coefficients for an inverted lip gate are negative; this
indicates that on this gate the flow generates suction at the bottom surface which substantially
increases the resulting downpull force. This increase in the downpull force is a desirable feature
for emergency high-head gates, where the gate must have a tendency to the closure operation to
guarantee the emergency closure. However, this large downpull forces require a higher capacity
servo-motor (see Figure 17. High-head slide gate – Components). In relation to the error
associated to the coefficients indicated in Table 12, it is noted that, from the mesh, time step and
Y+ independence study the difference in the resulting downpull force was negligible (±1%
difference), according to this, for practical purposes, the magnitude of the dimensionless bottom
downpull coefficients indicated in the table above may be applied for analytical predictions of the
downpull force on an inverted 30° lip gate.

Figure 43. Bottom downpull coefficients comparison – Experimental vs CFD compares the bottom
downpull coefficient (𝐾𝑏 ) as obtained from the scale model developed by (Naudascher E. , 1964)
2-73

and the CFD model. For ϴ = 45° a direct comparison between CFD and the analytical model is
possible. As discussed above, CFD tends to under estimate the value of Kb.

Interestingly, the values of Kb predicted by CFD are slightly higher than those predicted by the
analytical model. The reason for this difference is that the forces at the bottom surface of the gate
were somewhat higher for the CFD model compared to the analytical model.

Figure 43. Bottom downpull coefficients comparison – Experimental vs CFD


2-74

2.4 Conclusions
 A numerical flow simulation using Computational Fluid Dynamics of a bottom outlet with
two different gate geometries and for several gate positions was successfully achieved
using the commercial CFD software ANSYS Fluent, the obtained results were analytically
validated.

 The downpull force results obtained from the gate 1 CFD model are analytically validated
with the method developed by (Naudascher E. , 1964). This can be confirmed since the
downpull force values have similar magnitudes (5% - 10% difference) and the two curves
have the same tendency.

 The bottom downpull coefficients for the inverted lip slide gate (θ = -30°) for further
analytical calculations were obtained in a validated CFD model of a bottom outlet. These
values are indicated in Table 12. “Bottom downpull coefficients for inverted lip slide gates”.

 The water discharge obtained from the gate 1 and gate 2 CFD models were validated in
accordance to the experimental water discharge coefficients recommended by the U.S
Army Corps of Engineers Hydraulic Design Criteria.

 It is confirmed by the CFD results, that the existing experimental discharge coefficients,
indicated by the U.S Army Corps of Engineers Hydraulic Design Criteria, may be applied
for high-head slide gates, and that the geometry of the gate slightly affects the water
discharge through the bottom outlet (see Figure 49. Analytical vs CFD water discharge
validation). However, for practical purposes, the existing experimental water discharge
coefficients can be applied for analytical calculations, regardless of the geometry of the
slide gate. This can be concluded since the water discharge of the two gates simulated is
very similar (1%-3%) difference.

 According to the CFD results, the downpull force obtained from the CFD 45° lip gate is
5%-10% lower for all the gate positions (see Figure 53. Resulting downpull force –
Theoretical vs CFD 45° lip gate), compared to the one achieved from the analytical
method, which is based on experimental coefficients obtained from scale models.
2-75

however, it is noted that the tendency of the curve is similar. This difference in the obtaining
results is due to the following factors:

o Scale factors inherent to the model test developed by (Naudascher E. , 1964) and
the CFD model. The scale model tested was a flat gate with a 0,9 m x 0,9 m cross
section, while the CFD model considered a 3,0 m x 3,9 m cross section. The
thickness of the scale gate was 63,5 mm, while the thickness of the CFD gate was
650 mm.
o The piezometer distribution of the scale gate does not cover all the bottom surface
of the gate, and the measurements are assumed equal along the gate width (see
Figure 53. Piezometer tube distribution gate model test lip).
o The installations of the scale model and the CFD model are different. While the
scale model considered a flat gate which operates in a gate well, the CFD model
considered a high-head slide gate that operates in a gate bonnet. Although the
gate well and gate bonnet pressurize equivalently as the gate closes, it’s been
observed that the geometry of these installations have great effect on the resulting
downpull force (Erbisti, 2004).
o The CFD models considered a series of simplifications such as the RANS
equations (Reynolds Averaged Navier Stokes), space discretization and numerical
methods to obtain an approximate solution of the fluid flow governing equations.
However, the calibration process (See Appendixes A, B and D) of the models
indicated that the downpull results obtained with CFD were not very sensible to
changes in the simulation setup.

 According to the results obtained in this research, the bottom downpull coefficients for the
inverted lip slide gate (θ = -30°) and the typical flat bottom (θ = 0°) slide gate studied by
(Naudascher E. , 1964) are very similar, in accordance to this, for practical purposes, the
analytical calculation of the downpull force in this particular gate geometry may be
estimated using the existing experimental coefficients for the flat bottom slide gate (θ =
0°).
2-76

2.5 Future work


 It is noted that the results obtained in this research were obtained under pseudo-steady
state simulations, as well as the physical models studied by (Naudascher E. , 1964) for
the determination of the downpull coefficients. However, the hydraulic transients during an
opening or closure event of a high-head slide gate may have effects on the hydrodynamic
forces and the downpull phenomenon. This hydraulic transient simulation will be studied
using a dynamic mesh to represent the movement of the slide gate and measure the
change in the downpull force with time.

 During the operation of the bottom outlet, which is object of this research, the pressure in
the lower chamber of the servomotor will be measured. This pressure measurement can
be translated to a resulting operational force of the slide gate, which can be used to
determine the real downpull force during the closure operation. The data obtained from
this measurements can be applied for an additional validation of the results obtained in
this CFD research.

 The focus of this research is the high-head slide gate of the bottom outlet, which is used
as emergency or maintenance gate. The main gate which will control the water flow in this
bottom outlet is the high-head segment gate (see Figure 22. Side view of the bottom outlet
for the simulation). These segment gates are not subjected to downpull forces, however,
the water flow through the bottom outlet generates a hydrodynamic torque which increases
the operational forces on this gate. This is a very interesting fluid flow problem, object of
another CFD research.
2-77

REFERENCES

[1]. Abdolahpour, M. &. (2014). Flow Aeration after Gate in Bottom Outlet Tunnels. Arabian
Journal for Science and Engineering, 39(5), 3441-3448. http://doi.org/10.1007/s13369-
014-0954-5.

[2]. Akoz, M. S., Kirkgoz, M. S., & Ahmet, A. O. (2010). Experimental and Numerical Modeling
of a Sluice Gate Flow. Journal of Hydraulic Research, 167-176.

[3]. Amorim, J. C., & Lima de Andrade, J. (1999). Numerical Analysis of the Hydraulic
Downpull on Vertical Lift Gates. Waterpower & Dam Construction.

[4]. Arias Belduque, J. J., & Molina, A. (2015). Improvement of the performance of a
hydrocracking reacton using a computational fluid dynamics perspective. Universidad
Nacional de Colombia - Sede Medellín, Antioquia.

[5]. Aydin, I., Telci, I., & Dundar, O. (2006). Prediction of downpull on closing high head gates.
Journal of Hydraulic Research Vol. 44 No. 6, 11.

[6]. Bakker, A. (2006). Solution Methods - Applied Computational Fluid Dynamics. Obtenido
de Computational Fluid Dynamics: http://www.bakker.org/dartmouth06/engs150/

[7]. Calderón Villegas, D. (2016). Estudio computacional de la hidrodinámica y del sistema de


aireación en descargas de fondo con compuertas de alta presión. 102. Universidad
Nacional de Colombia - Sede Medellín.

[8]. Davidson, L. (2018). Fluid mechanics, turbulent flow and turbulence modeling. SE-412 96
Göteborg, Sweden: Chalmers.
2-78

[9]. Dr. Rassul, M. Almanini, Dr. Mustafa T. Al-Kifae, & Dr. Shaymaa A. M. Alhashimi. (2010).
Prediction of Downpull Force on Tunnel Gate with Different Gate Lip Gometry. Journal of
Kerbala University, Vol. 8 No.4 Scientific. 2010, 16.

[10]. Erbisti, P. (2004). Design of Hydraulic Gates. Lisse: Balkema Publishers.

[11]. Gerasimov. (December de 2006). Modeling turbulent flows. Obtenido de


Introductory FLUENT Training:
http://www.southampton.ac.uk/~nwb/lectures/GoodPracticeCFD/Articles/Turbulence_Not
es_Fluent-v6.3.06.pdf

[12]. Gómez, A. (2017). Curso de Hidrodinámica, Posgrado en aprovechamiento de


recursos hidráulicos, Universidad Nacional de Colombia Sede Medellín. Medellín.

[13]. Hargreaves, D., Morvan, H., & Wright, N. (29 de March de 2007). Validation of the
Volume of Fluid Method for Free Surface Calculation: The Broad-Crested Weir.
Engineering Applications of Computational Fluid Mechanics, Volume 1, 146.

[14]. Henriques da Silva, J. (September de 2011). Esforços Hidrodinâmicos em


Comportas Verticais (Downpull). Instituto Superior de Engenharia de Lisboa.

[15]. Kalinske, A. A. (1943). Closed Conduit Flow. Journal of Hydraulic Engineering,


108(Paper No. 2205), 1435–1447.

[16]. Kundu, P. K., & Cohen, I. (2002). Fluid Mechanics, Second Edition. San Diego:
Elsevier Science.

[17]. Kuzmin, D. (Introduction to Computational Fluid Dynamics). Technische Universität


Dortmund. Retrieved from Introduction to Computational Fluid Dynamics:
http://www.mathematik.uni-dortmund.de/~kuzmin/cfdintro/lecture1.pdf

[18]. Levin. (1968). Formulaire des conduites forcées oléoducs et conduits d' aeration.
Paris: Dunod.

[19]. Lewin, J. (1995). Hydraulic Gates and Valves. London: Thomas Telford
Publications.

[20]. Linfield, K., & Mudry, R. (2008). The Pros and Cons of CFD and Physical Flow
Modeling. Air Flow Sciences Corporation, 8.
2-79

[21]. Molina, A. (18 de Febrero de 2017). Curso Dinámica de Fluidos Computacional.


Universidad Nacional de Colombia - Sede Medellín, Antioquia, Colombia.

[22]. Naudascher, E. (1964). Hydrodynamic Analysis for High Head Leaf Gates. Journal
of the Hydraulics Division.

[23]. Naudascher, E. (1991). Hydrodynamic forces. Hydraulic Structures design manual:


Hydraulic Design considerations. The International Association for Hydraulic Research
(IAHR).

[24]. Prosperetti, A., & Tryggvason, G. (2007). Computational Methods for Multiphase
Flow. Cambridge: Cambridge University Press.

[25]. Quimby, B. (2007). Hydrodynamic loads. Retrieved from BG Structural


Engineering:
http://www.bgstructuralengineering.com/BGASCE7/BGASCE7006/BGASCE70604.htm

[26]. Roache, P. (1998). Verification and Validation in Computational Science and


Engineering. Hermosa Publishers.

[27]. Rodi, W. (1993). Turbulence Models and Their Application in Hydraulics. New York:
Taylor & Francis.

[28]. Rodríguez, J. D. (2013). Metodología para el control de riesgos en compuertas


radiales de descargas de fondo para grandes presas. Universidad Nacional de Colombia.

[29]. Roshan, M. A. (2014). Flow aeration after gate in bottom outlet tunnels. Arabian
Journal for Science and Engineering.

[30]. Sagar, B. (1977). Downpull in high-head gate installations. Water Power and Dam
Construction.

[31]. Sagar, B. (1995). Hydrogates task committee design guidelines for high head
gates. Journal of Hydraulics Engineering, ASCE 121.

[32]. Seo, J. (2014). Multiphase Flow Modeling with Free Surfaces Flow.

[33]. Uysal, M. A. (2014). Prediction of Downpull on high head gates using computational
fluid dynamics.
2-80

[34]. Versteeg, H., & Malalasekera, W. (2007). An Introduction to Computational Fluid


Dynamics, Second Edition. Harlow, England: Pearson Education Limited.

[35]. Vyzikas, T. (2014). Application of numerical models and codes. Plymouth, AZ,
USA: University of Plymouth.

[36]. Wasserman, S. (2016). Choosing the Right Turbulence Model for Your CFD
Simulation. Retrieved from engineering.com:
https://www.engineering.com/DesignSoftware/DesignSoftwareArticles/ArticleID/13743/C
hoosing-the-Right-Turbulence-Model-for-Your-CFD-Simulation.aspx

[37]. Zhiyin, Y. (2015). Large-eddy simulation: Past, present and the future. Chinese
Journal of Aeronautics, 28 (1). pp. 11-24. ISSN 1000-9361, 15.
2-81

Appendixes
2-82

Appendix A
Mesh independence study

Fort the mesh Independence study of the computational model of the bottom outlet, the magnitude
of the flow velocity was analyzed on the following line along the geometry

Figure 44. Line for the recompilation of the data for the mesh Independence study

Four different meshes were simulated to compare the data of the velocity measured in the line
shown in the figure above, the method used for the mesh independence study was using half the
number of computational cells for each simulation to demonstrate the influence of the mesh
refinement on four different levels, the parameter “n” was defined as the mean cell size for the
meshing, being the mesh “n” the most refined.
2-83

Table 12. Different configurations for the mesh Independence study


Mesh quality (max. Type of mesh
Mesh Mean cell size (m) Number of elements (U)
skewness)
n 0,10 2’613.459 0,47 Structured
2n 0,20 1’124.576 0,51 Structured
4n 0,40 457.584 0,53 Structured
8n 0,80 178.023 0,62 Structured
The four simulations were carried out with the same boundary conditions and the same simulation
time. The following figure shows the results of the velocity at the line shown in Figure 56 for the
four different meshes described in the table above:

Figure 45. Results of the velocity measured along the line for the four different meshes
As reported by the chart shown in the figure above, the data obtained from the simulation with
the meshes “4n” and “8n” are different from the one obtained with the meshes “n” and “2n”.
2-84

According to this the meshes “n” and “2n” are considered acceptable for the simulation of the
bottom outlet.

Appendix B
Time step independence study

To demonstrate that the results obtained from the simulations are not function of the time step
used, four simulations with different time steps were carried out. In these simulations the
parameter “t”, known as the time step was set in four different values, being the simulation “t” the
original time step, which was increased to the double at each simulation until the “8t” time step
was reached. The increased time step required a larger number of iterations to reach the
acceptable convergence residuals at each step. The following table shows the different
configurations for the simulations of the time step independence study.

Table 13. Different configurations for the time step independence study
Simulation Time step size (sec) Time step setting
t 0,001 Fixed
2t 0,002 Fixed
4t 0,004 Fixed
8t 0,008 Fixed

The comparison between the results of the four simulations indicated in Table 14 was obtained
by measuring the pressure along the model in the line indicated in Figure 56. The graphic
indicating this comparison is shown in Figure 58.
2-85

Figure 46. Pressure along Line for the different time step cases.

As indicated in Figure 58, the time step does not considerably affect the results of the simulation,
according to this, a time step independence is obtained for the models developed during this
research.
2-86

Appendix C
Residuals

The residuals obtained during the simulations developed during this research reached the
convergence criteria configured to 1e-5 for all the different residuals. Figures 60 and 61 illustrate
the residuals obtained for the 45° lip gate at 30% opening and the Inverted 30° lip gate at 60%
opening.

Figure 47. Residuals for the 45° lip gate – 30% opening

Figure 48. Residuals for the Inverted 30° lip gate – 60% opening
2-87

Appendix D
Y+ independence study

The purpose of this analysis is to determine the effect of the Y+ values on the resulting downpull
force on the surfaces of the slide gate. For this analysis, two different meshes were simulated for
the 30% opening (See Table 4. Gate positions for the simulation of the bottom outlet). The
simulation designated as Y considered a Y+ value of 300, while the simulation designated as 2Y
considered a Y+ value of 500. Table 15 indicates the parameters of the 2 simulations developed
for the Y+ independence analysis

Table 14. Main parameters of the simulations for the Y+ Independence study
First cell height (cm) Number of elements
Simulation Y+ Value
(U)
Y 300 1 2’613.459
2Y 500 1,6 1’598.156

The computed data for the simulations corresponds to the hydrodynamic force on the bottom
surface of the gate, which is subjected to the largest pressure gradients and viscous stresses.
Additionally, the forces on the bottom surface or gate lip are the main portion of the downpull
force. Figures 60 and 61 shows the detail of the two compared meshes at the bottom surface of
the gate:

Figure 49. Detail of mesh at bottom Surface of the gate for a Y+ of 300
2-88

Figure 50. Detail of mesh at bottom Surface of the gate for a Y+ of 500

The two meshes were simulated until a pseudo steady state on the bottom surface of the gate
was achieved after a time of 2,9 seconds of flow. The force applied on this surface is composed
of a pressure applied on the area of the surface and a viscous component due to the shear stress
generated by the water flow below the gate.

Force result – Direction vector (0 1 0)


Pressure force Viscous force (N)
Simulation Total force (N)
(N)
Y 886791 -1203 885.588
2Y 872648 -780 871.868

It can be concluded, based on the results obtained from the different Y+ values, that larger Y+
values does affect the resulting viscous force applied on the bottom surface of the gate, however,
since 99% of the force corresponds to pressure forces and 1% corresponds to viscous forces, this
difference does not affect the downpull force of the gate.
2-89

Appendix E
Mesh quality reports

Figure 51 illustrates one of the obtaining mesh quality reports for the different meshes applied for
the simulations of the bottom outlet.

Figure 51. Quality report (Skewness) of the meshes for the CFD simulation of the bottom outlet
(45° lip gate and Inverted 30° lip gate).
2-1

Appendix F
Hydraulic data

Tables 17 to 21 indicate the data for the analytical calculations developed and the averaged CFD data of the downpull force and
coefficients.

Table 15. Water discharge validation

Water discharge - bottom outlet - CFD Validation

Piezometric Water Water


Clear Clear Water
Gate Head in the Entrance head at the Discharge discharge discharge
opening opening discharge
opening reservoir Head loss contracted coefficient CFD - Gate CFD - Gate
height width theoretical
jet 1 model 2 model

[%] [m] [m] [m] [m] [m] [m/m] [m3/s] [m3/s] [m3/s]

10% 0,39 3 70 3,550754889 0,2847 0,73 30,77307455 32,548 23,552


20% 0,78 3 70 3,550754889 0,5694 0,73 61,41359261 65,291 52,216
30% 1,17 3 70 3,599395367 0,8658 0,74 93,13744979 94,382 82,84
40% 1,56 3 70 3,599395367 1,1544 0,74 123,9095276 126,84 115,526
50% 1,95 3 70 3,648035845 1,4625 0,75 156,5502354 159,513 144,648
60% 2,34 3 70 3,745316801 1,8018 0,77 192,2199738 193,662 188,852
70% 2,73 3 70 3,793957279 2,1294 0,78 226,5050454 228,721 225,549
80% 3,12 3 70 3,891238235 2,496 0,8 264,5376327 270,556 263,578
90% 3,51 3 70 3,891238235 2,808 0,8 296,874112 300,853 295,077472
2-2

Table 16. Hydraulic parameters for the analytical calculation of downpull for the 45° lip gate
Hydraulic downpull on high head slide gate - Analytical 45° lip gate
Bottom Top Bottom Resulting
Clear Top Velocity
Gate Clear Gate downpull downpull downpull downpull
openin Distanc Distanc downpull of the
openin openin thicknes coefficient force force force
g e A1 e A2 coefficien contracte
g g width s theoretica theoretica theoretica Theoretica
height t d jet
l l l l
[%] [m] [m] [m] [m] [m] [m/m] [m/m] [m/s] [N] [N] [N]
10% 0,39 3 0,65 0,25 0,0625 0,94 0,79 35,95 1186116 995596 190520
20% 0,78 3 0,65 0,25 0,0625 0,94 0,65 35,86 1179904 814871 365033
30% 1,17 3 0,65 0,25 0,0625 0,94 0,58 35,78 1174708 723914 450794
40% 1,56 3 0,65 0,25 0,0625 0,94 0,55 35,68 1168285 682717 485569
50% 1,95 3 0,65 0,25 0,0625 0,94 0,53 35,56 1160425 653464 506961
60% 2,34 3 0,65 0,25 0,0625 0,94 0,55 35,46 1153651 674165 479486
70% 2,73 3 0,65 0,25 0,0625 0,94 0,56 34,27 1077612 641179 436433
80% 3,12 3 0,65 0,25 0,0625 0,94 0,58 31,72 923142 568886 354256
90% 3,51 3 0,65 0,25 0,0625 0,94 0,54 28,19 729396 418491 310905

Table 17. Results of downpull for the 45° lip gate


Hydraulic downpull on slide gate - 45° lip gate

Top downpull force Bottom downpull Resulting downpull CFD top downpull CFD Bottom
Gate opening
CFD force CFD force CFD coefficient downpull coefficient

[%] [N] [N] [N] [m/m] [m/m]


10% 1240879 1095460 145419 0,984630414 0,86924127
20% 1239945 949730 290215 0,989069464 0,75757307
30% 1263321 885587 377734 1,01217322 0,70953261
2-3

Hydraulic downpull on slide gate - 45° lip gate

Top downpull force Bottom downpull Resulting downpull CFD top downpull CFD Bottom
Gate opening
CFD force CFD force CFD coefficient downpull coefficient

40% 1248740 804424 444316 1,005991281 0,64804806


50% 1236709 787248 449461 1,003047623 0,6385069
60% 1212840 782156 430684 0,989464333 0,63810186
70% 1183391 775557 407834 1,033562928 0,67736442
80% 1091001 775557 315444 1,112314439 0,79070803
90% 876037 675557 200480 1,130394199 0,87170486

Table 18. CFD data obtained from the Inverted 30° lip gate simulations
Hydraulic downpull on slide gate - Inverted 30° lip gate

Top downpull force Bottom downpull Resulting downpull CFD top downpull CFD Bottom
Gate opening
CFD force CFD force CFD coefficient downpull coefficient

[%] [N] [N] [N] [m/m] [m/m]


10% 1269863 -245418 1515281 1,007629053 -0,19473778
20% 1246595 -327224 1573819 0,994373983 -0,26101744
30% 1264084 -377350 1641434 1,012784536 -0,30233295
40% 1256497 -35987 1292484 1,01224036 -0,02899131
50% 1249952 -338914,4 1588866,4 1,013788516 -0,27488058
60% 1224579 -598648 1823227 0,999041295 -0,48839158
70% 1218884 -495295 1714179 1,064562192 -0,43258614
80% 1156497 -429870 1586367 1,179089947 -0,4382678
90% 1076552 -79149 1155701 1,389128696 -0,1021299
2-4

Table 19. Coefficients for the calculation of downpull – Theoretical and CFD
Bottom downpull coefficients for hydraulic gates - CFD and experimental

θ = 45 θ = 30 θ = 20 θ=0 Gate opening (%)


(Experimental) (Experimental) (Experimental) (Experimental) θ = -30 (CFD) θ = 45 (CFD)
0,000 0,000 0,000 0,000 0,000 0,000 0%
0,790 0,590 0,430 -0,200 -0,195 0,869 10%
0,650 0,390 0,210 -0,360 -0,261 0,758 20%
0,580 0,280 0,090 -0,280 -0,302 0,710 30%
0,550 0,210 0,000 -0,290 -0,290 0,648 40%
0,530 0,200 -0,020 -0,410 -0,275 0,639 50%
0,550 0,250 0,000 -0,630 -0,488 0,638 60%
0,560 0,290 0,050 -0,630 -0,493 0,677 70%
0,580 0,310 0,140 -0,500 -0,438 0,791 80%
0,540 0,400 0,300 -0,290 -0,202 0,872 90%
0,000 0,000 0,000 0,000 0,000 0,000 100%
2-1

Appendix F
CFD results for the inverted 30° lip gate

Figure 52. Velocity streamline – Inverted 45° lip gate – 50% gate opening
2-2

Figure 53. Streamlines colored by velocity – Inverted 30° lip gate – 50% opening

Figure 54. Streamlines colored by velocity – Inverted 30° lip gate – 50% opening
2-3

Figure 55. Volume fraction of the Inverted 30° lip gate – side view - 30% gate opening
2-4

Figure 56. Volume fraction of the Inverted 30° lip gate – side view - 10% gate opening
2-5

Figure 57.Static Pressure distribution on the lower lip of the slide gate – Inverted 30° lip gate –
90% opening
2-6

Figure 58.Static Pressure distribution at the domain– Inverted 30° lip gate – 30% opening

You might also like