Sei sulla pagina 1di 58

EXPERIMENTAL ANALYSIS AND NUMERICAL SIMULATION

OF HYDRAULIC JUMP

A thesis submitted toward partial fulfillment of the requirements


for the degree of

MASTER OF ENGINEERING
In Water Resources and Hydraulic Engineering
Course affiliated to Faculty of Engineering & Technology
Jadavpur University

Submitted by
TIRTHA ROY BISWAS
Exam Roll No: M4WRE1612

Under the guidance of


Dr. SUBHASISH DAS
Assistant Professor
School of Water Resources Engineering, Jadavpur University
&
Prof. (Dr.) ASIS MAZUMDAR
Director
School of Water Resources Engineering, Jadavpur University

School of Water Resources Engineering


M.E. (Water Resources & Hydraulic Engineering) course affiliated to
Faculty of Engineering & Technology
Jadavpur University
Kolkata 700032, India

2016

i
Declaration of Originality and Compliance of Academic Ethics

I hereby declare that this thesis contains literature survey and original research
work by the undersigned candidate, as part of my Master of Engineering in
Water Resources & Hydraulic Engineering in the Faculty of Interdisciplinary
Studies, Jadavpur University during academic session 2015-16.

All information in this document have been obtained and presented in accordance
with academic rules and ethical conduct.

I also declare that, as required by these rules and conduct, I have fully cited and
referenced all material and results that are not original to this work.

Name : Tirtha Roy Biswas

Exam Roll Number : M4WRE1612

Thesis Title : Experimental Analysis and Numerical


Simulation of Hydraulic Jump

Signature with Date :

iii
M.E. (Water Resources & Hydraulic Engg.) course affiliated to
Faculty of Engineering & Technology
Jadavpur University
Kolkata, India

Certificate of Recommendation

This is to certify that the thesis entitled “Experimental Analysis and


Numerical Simulation of Hydraulic Jump” is a bonafide work carried out
by Mr. Tirtha Roy Biswas under our supervision and guidance for partial
fulfillment of the requirement for the Post Graduate Degree of Master of
Engineering in Water Resources & Hydraulic Engineering during the
academic session 2015-2016.

THESIS ADVISOR THESIS ADVISOR


Dr. Subhasish Das Prof. (Dr.) Asis Mazumdar
Assistant Professor Director
School of Water Resources Engineering School of Water Resources Engineering
Jadavpur University Jadavpur University

DIRECTOR
School of Water Resources Engineering
Jadavpur University

DEAN
Faculty of Interdisciplinary Studies, Law & Management
Jadavpur University

v
M.E. (Water Resources & Hydraulic Engg.) course affiliated to
Faculty of Engineering & Technology
Jadavpur University
Kolkata, India

CERTIFICATE OF APPROVAL **

This foregoing thesis is hereby approved as a credible study of an engineering

subject carried out and presented in a manner satisfactorily to warranty its

acceptance as a prerequisite to the degree for which it has been submitted. It is

understood that by this approval the undersigned do not endorse or approve any

statement made or opinion expressed or conclusion drawn therein but approve the

thesis only for purpose for which it has been submitted.

Committee of
Final Examination
For the evaluation
Of the thesis

** Only in case the thesis is approved.

vii
ACKNOWLEDGEMENT
_____________________________________________________

I express my sincere gratitude to my supervisors Prof. (Dr.) Asis Mazumdar and Dr.
Subhasish Das under whose supervision and guidance this work has been carried out.
It would have been impossible to carry out this thesis work with confidence without
their wholehearted involvement, advice, support and constant encouragement
throughout. They have not only helped me in my thesis work but also have given
valuable advice to proceed further in my life.

I also express my sincere gratitude to Dr. Debasri Roy, Associate Professor; Prof. (Dr.)
Arunabha Majumder, Professor Emeritus; Dr. Pankaj Kr. Roy, Associate Professor
and Dr. Rajib Das, Assistant Professor of School of Water Resources Engineering,
Jadavpur University for their valuable suggestions.

I would also express my sincere thanks to Mr. Biprodip Mukherjee, Research Scholar
and Mr. Asim Kuila, Research Scholar of School of Water Resources Engineering,
Jadavpur University for their unconditional support and affection during my work.

Thanks are also due to all staff of School of Water Resources Engineering and the
Regional Centre, NAEB, Jadavpur University for their help and support.

I am also grateful to my parents who have been my strength and inspiration


throughout my M.E. course.

Date : May, 2016


Place : Jadavpur University TIRTHA ROY BISWAS
(Exam Roll No: M4WRE1612)

ix
ABSTRACT

Saint Venant equations are numerically solved to simulate the formation of hydraulic
jump in a rectangular channel having a small bed slope. The MacCormack’s scheme is
used for their solution by applying specified initial and boundary conditions until a
steady state flow is reached. The location of the hydraulic jump is determined as a part
of these computations. The artificial viscosity technique should be used in the
computations to dampen the superior oscillations near the steep gradient of the
simulated hydraulic jump. Twenty laboratory experiments were carried out for
verification of the numerical model. Upstream Froude number for these experiments
ranged from 2.17 to 7.0 in three different bed slopes 0, 0.02174, 0.0475. The simulated
hydraulic jump profiles using the MacCormack’s scheme shows a good agreement with
the experimental data. An empirical equation was developed to determine the location of
hydraulic jump using regression analysis based on simulated data. A software based on
Computational Fluid Dynamics (CFD) was also used to simulate two of these
experiments. The results obtained from CFD analysis matched fairly with the
experimental results.

xi
CONTENTS
Declaration iii

Certificate of Recommendation v

Certificate of Approval vii

Acknowledgement ix

Abstract xi

1.0 INTRODUCTION 1
2.0 LITERATURE REVIEW 1
3.0 OBJECTIVE 2
4.0 METHODOLOGY 3
4.1 Governing Equations 3
4.2 Numerical Scheme 5
4.2.1 MacCormack’s Method 5
4.2.2 Initial and Boundary condition 7
4.2.3 Stability criteria and Artificial viscosity 7
4.3 Experimental investigations 8
5.0 NUMERICAL MODEL 9
5.1 Simulation procedure 12
5.2 Spatial grid size 12
6.0 COMPARISON BETWEEN EXPERIMENTAL AND NUMERICAL
RESULTS 12
6.1 Discussion 33
7.0 GENERAL EQUATION USING REGRESSION ANALYSIS 34
8.0 SIMUALTION BY A CFD SOFTWARE 36
8.1 Volume of Fluid method (VOF) 37
8.2 Simulation procedure 37
8.3 Result 38
9.0 SUMMARY AND CONCLUSION 40
10.0 FUTURE SCOPE 41
11.0 REFERENCES 41

xiii
1.0 INTRODUCTION
Hydraulic jump is a sudden transition to subcritical flow from a supercritical flow. The
phenomenon is observed in canals below sluice gates, at the foot of spillways or when
there is an abrupt change in slope from steep to flat. Hydraulic jumps are practically
used as energy dissipaters below spillways to avoid scouring in the downstream, for
mixing chemicals and aerate water for city water supplies or for removing air pockets
from water supply lines for preventing air locking.

It is essential to determine the various parameters like length and location of the jump,
amount of energy dissipated etc. to design a hydraulic structure. The theory of jump was
developed in earlier days mainly on the basis of extensive experimental and empirical
work. The first experimental investigation was made by Bidone in 1818. Safranez
(1929), Bakhmeteff and Matzke (1936), Bradley and Peterka (1957), Rouse et al.
(1959), Silvester (1964), A.S.C.E. Task Force (1964), Rajaratnam (1967, 1968), Garg
and Sharma (1971), Leutheusser and Kartha (1972), Sarma and Newnham (1973), and
Hager (1987) conducted extensive laboratory experiments and developed design charts
and empirical relationships between various flow parameters. The advent of high speed
digital computers and developments in Computational fluid dynamics gave the
opportunity to solve the governing equations of hydraulic jump numerically.

2.0 LITERATURE REVIEW


An extensive amount of data has been reported in the literature on the hydraulic jump.
To determine the jump location, Chow (1959) computed the water surface profiles for
supercritical flow starting from the upstream end and the subcritical flow starting from
the downstream end, and the jump is formed at a location where the specific forces on
both sides of the jump are equal. McCorquodale and Khalifa (1983) used the strip-
integral method to compute the jump length, water surface profile, and pressures at the
bed. Abbott et al. (1969) used the finite-difference method and Katopodes (1984) used
the finite-element method to solve the St. Venant equations numerically until a steady
state was reached. The location of the hydraulic jump is automatically computed as part
of the solution.

Gharangik and Chaudhary (1991) did the most rudimentary work on numerical
simulation of hydraulic jump. Boussinesq equations describing one-dimensional

1
unsteady, rapidly varied flows were integrated numerically to simulate both the sub- and
supercritical flows and the formation of a hydraulic jump in a rectangular channel
having a small bottom slope. For this purpose the MacCormack (second-order accurate
in space and time) and two-four (second-order accurate in time and fourth-order in
space) explicit finite-difference schemes were used to solve the governing equations
subject to specified end conditions until a steady state was reached. There was
significant agreement between the experimental and numerical results. The fourth order
accurate model was found to be more precise than the second order accurate model. But
the simulations showed that the Boussinesq terms have little effect in determining the
location of the hydraulic jump.

Chippada et al. (1994) numerically simulated hydraulic jump on a straight horizontal


channel with supercritical Froude numbers 2.0 and 4.0 by solving the Reynolds
averaged Navier-Stokes equations. Turbulence was modeled through the closure
equations. Galerkin finite element method with three-noded triangular elements is used
for spatial discretization. A detailed study of the internal and external characteristics of
hydraulic jump is done and compared with experimental values where possible.

Hamidreza et al. (2014) used the commercial CFD software FLOW 3D to simulate the
hydraulic jump in the convergence stilling basin. The software was applied to
numerically solve the Navier– Stokes equations for solution domains, namely the shout,
the stilling basin and the downstream of dam, and to estimate the turbulence flow, the
standard k-ε and RNG models was used. These models are based on the volume-of-fluid
method, and they are capable of simulating the hydraulic jump. The calculated results
such as the pressure, the velocities, the flow rate, the surface height air entranced, the
kinetics energy, the kinetics energy dissipated, and the Froude number were compared
with the scale model data where available. The physical model and CFD model results
showed good correlations.

3.0 OBJECTIVE

Hydraulic jump is a rapidly varying unsteady flow phenomenon which can be described
by the one dimensional St. Venant equation which assumes hydrostatic pressure
distribution and uniform velocity distribution.
 This research study is done to numerically simulate hydraulic jump using a finite
difference scheme to solve the governing equations.
 Experiments were conducted for different Froude numbers and three different
slopes. A comparison between the numerically computed and experimentally
measured data is made.
 Also CFD software ANSYS CFX developed by ANSYS, Inc. an American
Computer-aided engineering software is used to compare with the computed and
measured data. CFX is a solver used particularly for multiphase flows.
 As it is evident from the literature review, after Gharangik (1991) numerical
simulation of hydraulic jump using Finite Difference Method has not been done.
Gharangik conducted the experiments with horizontal bed for different Froude
numbers. But natural channels usually have some slope. In this study experiments
are conducted for two non-zero slopes for different Froude numbers ranging from
2.17 to 7. To which degree the numerical model is capable of modeling the jump is
investigated.

4.0 METHODOLOGY

Experiments were conducted in the laboratory flume for different upstream Froude
numbers and different bed slopes. For the numerical model the St. Venant equation is
used as the governing equation. The governing equations are solved numerically using
the MacCormack method. A source code is written in MATLAB (matrix laboratory),
which is a proprietary product of MathWorks, to do the numerical computations. The
details of the governing equations; numerical scheme and the experimental investigation
are given in the subsequent sections. Also the various aspects of ANSYS CFX
simulation are provided in a later section.

4.1 Governing equations


Unsteady flow phenomenon such as hydraulic jump in rectangular channel are often
modeled as one dimensional flow which can be described by a set of quasi-linear,
hyperbolic partial differential equations, called the Saint-Venant equations. The
derivation of the equations involves the following assumptions: a) The pressure
distribution is hydrostatic. b) The channel bottom slope is small enough such that the
depth measured normal to the channel bed is approximately equal to the depth measured

3
vertically. c) The velocity distribution is uniform over the entire channel cross-section.
d) The channel is prismatic –i.e., the channel bottom and cross-section remains the same
along the distance. e) Friction losses for a given flow velocity during unsteady flow is
same as that during steady flow.

The one dimensional St. Venant equation may be written as follows:

Continuity:

…………………………………………………………… (1)

Momentum:

, - ( ) ………………………………. (2)

Where x = distance along the channel bottom (considered positive in the downstream
direction); t = time; u = flow velocity in the x-direction; h = flow depth; g = acceleration
due to gravity; = channel bottom slope; and = slope of the energy grade line.

Manning equation may be used to calculate the friction slope, i.e., = Mn2u|u|Co2R4/3 in
which Mn = Manning roughness coefficient; R = hydraulic radius; C0 = correction factor
for the units; C0 = 1 in the SI units; and C0 = 1.49 in the English units.

The equations are in conservation form and can be written as:

……………………………………………………..……… (3)

In which

G=* +;

F=[ ];

S=[ ( )];
4.2 Numerical scheme

A hydraulic jump can be simulated by solving the above governing equations with
appropriate boundary conditions. The initial conditions are specified and the iterations
are continued until steady state is reached. If shock capturing numerical technique is
used, then the jump forms as a part of the steady state solution.

Several second order schemes have been used in CFD for solving hyperbolic equations
of which MacCormack scheme is widely popular because of its simplicity, robustness,
and capability for capturing shocks. The scheme is second order accurate in space and
time and is used to solve the governing equations.

4.2.1 MacCormack method

The MacCormack‟s scheme developed by MacCormack(1969) is a two level predictor


corrector scheme. In the predictor part, forward finite difference is used for the spatial
derivative terms and backward finite difference approximation is used in the corrector
part. With reference to Figure 4.1 the finite difference expressions are:

Predictor step:

……………………………..…...…………………………………….. (4)

.........……………………………………………………………….. (5)

Where i=the node in x direction; k=node in time direction; asterisks refers to the
predicted values of the variable; and and spatial grid size and time step size
respectively.

Based on the above finite difference approximation (1) and (2) may be written for
determining the predicted values hi* and ui*hi*:

………………………..………….……………….. (6)

,*( ) ( ) + *( ) ( ) +-

( )……………………………………………………………………… (7)

5
𝑥

k+1
𝑡
k

i-1 i i+1

Figure 4.1 Finite difference grid

Where

………...……………………………………………………….…….. (8)

Corrector step:

……………………….……...………………………………………… (9)

…………..……….………………………………………………… (10)

Two double asterisks denote the values of the variable after the corrector step. Based on
the above finite difference approximations the variables after the corrector step can be
written as:

………………………..…..……………...……... (11)

,* + * +-

( )……………………………………………...…………………….... (12)
Where

…………………………………………………………..………….. (13)

The channel is divided into n equal reaches. Thus if the upstream end is numbered
section 1, then the downstream end will be n+1.

4.2.2 Initial and boundary conditions

The flow velocity and flow depth is specified at time t=0 as the initial condition. The
flow is assumed to be supercritical initially in the entire channel. The initial steady state
flow depth is determined by integrating the gradually varied flow equation; starting with
the specified flow depth and velocity at section 1:

……………………………………..…………………………....... (14)

The MacCormack‟s scheme is used to compute the variables at the interior nodes. At the
boundaries the flow conditions are specified. The flow depth and velocity are specified
at the upstream boundary and they are same as the initial conditions. At the downstream
boundary, a constant flow depth is specified and the flow velocity is calculated from the
characteristic form of (1) and (2) using a forward finite-difference approximation, i.e.,
the velocity; at the unknown time level k + 1 is determined from the following
equation (Chaudhry 1987):

( ) ( ) ( )…………………………..…... (15)

Where √ is the celerity of a gravity wave.

4.2.3 Stability criteria and Artificial viscosity

The MacCormack scheme is stable if the following Courant-Friedrichs-Lewy (CFL)


criterion is satisfied:

……………………..………………………………………. (16)

7
In which = desired Courant number; must be less than or equal to 1.0 for the
MacCormack scheme.

A modified-equation analysis of finite-difference schemes presented by Warming and


Hyet (1974) shows that higher order terms are introduced which are not present in the
governing partial differential equations. These terms represent the truncation errors
which affect the behavior of the scheme and contribute to the scheme what is known as
diffusion and dispersion. The solution has dissipative errors if the leading term in the
truncation error contains an even derivative and dispersive errors if the leading term has
odd derivatives (Anderson et al., 1984). For less than that required by the Courant-
Friedrich-Lewy (CFL) limit, the dispersive errors result in introducing numerical
oscillations in the solution. Therefore it becomes necessary to add artificial viscosity to
smooth these oscillations. Several procedures have been reported in the literature for this
purpose.

The procedure, developed by Jameson et al. (1981) and used herein, has the advantage
of smoothing regions where the solution has large gradients while leaving relatively
smooth areas undisturbed; i.e., high-frequency oscillations are smoothed. A parameter
is first computed from a normalized form of the gradients of one variable. For the
studies reported herein, the depth of flow h was selected for determining the parameter:

| |
……………………………… ………….…………….. (17)
| | | | | |

( )
……………………………….…….………..…………. (18)

In which is used to regulate the amount of dissipation. The computed variables are
then modified as

( ) ( )…………………………..(19)

4.3 EXPERIMENTAL INVESTIGATIONS

The experiment was carried out in a rectangular Perspex flume 36.5cm width, 45cm
height and 5m long. The discharge was measured by a digital Flow meter. The flume
has a tail gate to control the water depth. By adjusting the tail water depth the jump
position was varied. The flow depths at the upstream end and in the section of the flume
with metal walls were measured at equally spaced intervals by a point gauge having an
accuracy of 1 mm. There were continuous undulations in the water surface downstream
of the jump. The maximum and minimum levels of these undulations at a location were
marked and an average of these levels was considered the depth at that location.
Twenty laboratory experiments were carried out with the Froude number upstream of
the jump ranging from 2.17 to 7 for three different slopes 0.0435, 0.02174 and 0. The
slopes were created by tilting the flume using two blocks of wood. The details of the
laboratory experiments are listed in Table 5.1.

5.0 NUMERICAL MODEL

A second order accurate numerical model is developed. The St. Venant equations are
solved by the MacCormack‟s scheme. The time-step size was restricted by the Courant
stability condition and the spatial grid size. The Courant number was set equal to 0.65
since best results are obtained when it is approximately equal to 2/3.

To smooth the high-frequency oscillations near the jump, the dissipation coefficient k in
the Jameson's formula, was determined by a trial-and error procedure. It is desirable to
keep its value as low as possible while still smoothing the high-frequency oscillations.
Trials with values ranging from 0.01 to 0.05 indicated that a value of 0.03 provided the
best results.

To run the numerical model, the flow depth and velocity at the upstream end and only
the flow depth at the downstream end are specified. The upstream flow velocity for each
run was computed from the continuity equation , where Q = discharge; B =
flume width;

h = measured depth; and u = flow velocity. The Froude number of the incoming flow is
determined from the equation √ .

9
Figure 5.1 Experimental flume

Figure 5.2 Pump Figure 5.3 Wooden block to tilt


the flume
Table 5.1: Experimental details

Experiment Discharge Upstream Froude No Slope Downstream


No. Depth Depth

1 19.1 0.039 2.3 0.0435 0.234


2 19.1 0.033 2.9 0.0435 0.2
3 19.1 0.04 2.17 0.0435 0.26
4 19.1 0.032 3.06 0.0435 0.268
5 19.1 0.027 3.83 0.0435 0.215
6 19.1 0.026 4.23 0.0435 0.265
7 19.1 0.023 5 0.0435 0.272
8 19.1 0.02 6.18 0.0435 0.273
9 19.1 0.018 7 0.0435 0.285
10 19.1 0.039 2.3 0.02174 0.167
11 19.1 0.033 2.9 0.02174 0.17
12 19.1 0.04 2.17 0.02174 0.182
13 19.1 0.032 3.06 0.02174 0.16
14 19.1 0.027 3.83 0.02174 0.175
15 19.1 0.026 4.23 0.02174 0.205
16 19.1 0.023 5 0.02174 0.215
17 19.1 0.02 6.18 0.02174 0.205
18 11.2 0.023 3.29 0 0.07
19 15 0.027 3.06 0 0.08
20 19.1 0.027 3.83 0 0.11

11
5.1 Simulation procedure

The upstream flow depth, velocity, Froude numbers, and downstream flow depth for
different runs are listed in Table 5.1. In these runs, the Manning coefficient (Mn) for the
flume was determined by trial and error by matching the computed water surface profile
with the measured water levels in the flume during the initial steady supercritical flow.
For the range of Froude numbers tested, these values varied from 0.014 to 0.016
depending upon the flow depth since the bottom of flume is made up of metal and the
sides are made up of glass. First, the initial steady-state depth and velocity at every
computational node were computed by assuming the flow to be supercritical throughout
the flume. Then, the unsteady computations were started by increasing the downstream
depth to the value measured during the experiment (see Table 5.1). The computations
were continued until they converged to the final steady state for the specified end
conditions.

5.2 Spatial grid size

One very important parameter in the simulation of a hydraulic jump is the size of the
spatial grid, . Its value was varied from 0.1 m to 0.4 m. Values greater than 0.4 m and
in some cases 0.3 m could not be used since they resulted in the jump forming at less
than three computational nodes. For Fr = 7, and slope 0.0435 the jump was computed to
form between 1.60 m and 2.10 m from the upstream end of the flume, with the average
distance of four different values of being 1.95 m. In other words, the jump location
may be predicted in a satisfactory manner for typical engineering applications by using a
reasonable value of .

6.0 COMPARISON BETWEEN EXPERIMENTAL AND


NUMERICAL RESULTS

Once the numerical solution converged to a steady state the depths at the corresponding
grid points is obtained, which gives the flow profile along with the jump. The numerical
data is compared with the experimentally obtained flow profile. Figure 6.1 to 6.20 give a
comparative view between numerical and experimental results for flows with different
pre jump Froude number and different slopes. The Figures show the side view and top
view of the jump location during the experiments and a graph to compare the
experimental and numerical jump profile. The depth of flow in meters is plotted in the
vertical axis and „x‟ in meters is the distance along the flume, plotted in the horizontal
axis. The profile obtained from experiment is plotted in red and the numerical profile is
plotted in black.

The St. Venant equation assumes hydrostatic pressure distribution. But the pressure
distribution in rapidly varied flow phenomenon like hydraulic jump which has steep
water surface gradients is not hydrostatic. Equations describing these flows were
derived by McCowan (1985) and Basco (1983) assuming that the vertical velocity varies
from zero at the channel bottom to its maximum value at the free surface. These
equations referred to as the Boussinesq equations. But Gharangik (1991) in his work
showed that the Boussinesq terms have little effect in predicting the jump location.
Numerical models gave the same result for the St. Venant equation and Boussinesq
equations when fourth order accurate numerical scheme is used. Thus the accuracy was
more dependent on the order of the numerical scheme. MacCormack scheme being
second order accurate in space and time could not predict the jump location precisely for
higher Froude numbers but gave good results for low Froude numbers as is evident from
the given graphs comparing the experimental and numerical results.

13
Experiment no. 1

Figure 6.1.a Side view of the jump Figure 6.1.b Top view of the jump

Figure 6.1.c Jump Profile for Fr =2.30, Slope= 0.0435


Experiment no. 2

Figure 6.2.a Side view of the jump Figure 6.2.b Top view of the jump

Figure 6.2.c Jump Profile for Fr =2.90, Slope= 0.0435

15
Experiment no. 3

Figure 6.3.a Side view of the jump Figure 6.3.b Top view of the jump

Figure 6.3.c Jump Profile for Fr =2.17, Slope= 0.0435


Experiment no. 4

Figure 6.4.a Side view of the jump Figure 6.4.b Top view of the jump

Figure 6.4.c Jump Profile for Fr =3.06, Slope= 0.0435

17
Experiment no. 5

Figure 6.5.a Side view of the jump Figure 6.5.b Top view of the jump

Figure 6.5.c Jump Profile for Fr =3.83, Slope= 0.0435


Experiment no. 6

Figure 6.6.a Side view of the jump Figure 6.6.b Top view of the jump

Figure 6.6.c Jump Profile for Fr =4.23, Slope= 0.0435

19
Experiment no. 7

Fugure 6.7.a Side view of the jump Figure 6.7.b Top view of the jump

Figure 6.7.c Jump Profile for Fr =5.0, Slope= 0.0435


Experiment no. 8

Figure 6.8.a Side view of the jump Figure 6.8.b Top view of the jump

Figure 6.8.c Jump Profile for Fr =6.18, Slope= 0.0435

21
Experiment no. 9

Figure 6.9.a Side view of the jump Figure 6.9.b Top view of the jump

Figure 6.9.c Jump Profile for Fr =7.0, Slope= 0.0435


Experiment no. 10

Figure 6.10.a Side view of the jump Figure 6.10.b Top view of the jump

Figure 6.10.c Jump Profile for Fr =2.30, Slope= 0.02174

23
Experiment no. 11

Figure 6.11.a Side view of the jump Figure 6.11.b Top view of the jump

Figure 6.11.c Jump Profile for Fr =2.90, Slope= 0.02174


Experiment no. 12

Figure 6.12.a Side view of the jump Figure 6.12.b Top view of the jump

Figure 6.12.c Jump Profile for Fr =2.17, Slope= 0.02174

25
Experiment no. 13

Figure 6.13.a Side view of the jump Figure 6.13.b Top view of the jump

Figure 6.13.c Jump Profile for Fr =3.06, Slope= 0.02174


Experiment no. 14

Figure 6.14.a Side view of the jump Figure 6.14.b Top view of the jump

Figure 6.14.c Jump Profile for Fr =3.83, Slope= 0.02174

27
Experiment no. 15

Figure 6.15.a Side view of the jump Figure 6.15.b Top view of the jump

Figure 6.15.c Jump Profile for Fr =4.23, Slope= 0.02174


Experiment no. 16

Figure 6.16.a Side view of the jump Figure 6.16.b Top view of the jump

Figure 6.16.c Jump Profile for Fr =5, Slope= 0.02174

29
Experiment no. 17

Figure 6.17.a Side view of the jump Figure 6.17.b Top view of the jump

Figure 6.17.c Jump Profile for Fr =6.18, Slope= 0.02174


Experiment no. 18

Figure 6.18.a Side view of the jump Figure 6.18.b Top view of the jump

Figure 6.18.c Jump Profile for Fr =2.90, Slope= 0.0

31
Experiment no. 19

Figure 6.19.a Side view of the jump Figure 6.19.b Top view of the jump

Figure 6.19.c Jump Profile for Fr =3.06, Slope= 0.0


Experiment no. 20

Figure 6.20.a Side view of the jump Figure 6.20.b Top view of the jump

Figure 6.20.c Jump Profile for Fr =3.83, Slope= 0.0

6.1 Discussion

The comparison between the computed and the numerical results shows that as the
Froude number increases the jump forms downstream of the location obtained
experimentally. For the slope 0.0435 for Froude numbers more than 4.23 the jump is
located downstream of the location measured experimentally by a distance of 0.25m to
0.45m. For the slope 0.02174 the jump location is about 0.26m to 0.475m downstream
from that measured. The jump location is simulated better at lower Froude numbers and
matches closely for Froude numbers 3.83 and less. The overall comparison shows, as
shown in Figure 6.21, the deviation of the jump location obtained numerically from the
measured location is within ±25%.
33
Figure 6.21 Comparison between the experimental and numerical jump location

7.0 GENERAL EQUATION USING REGRESSION ANALYSIS


Sometimes it may be useful to form an empirical relationship between the parameters
determining the jump location. The results obtained from the empirical equation are then
compared with the numerical results.

It is assumed that the dependent variable (L), which is the distance from the beginning
of the flume to the location of hydraulic jump as shown in Figure 7.1 is a function of the
following independent variables: density of flow , the upstream water depth (hu), the
tail water depth (ht) at the end of the channel, the upstream velocity (u), the acceleration
of gravity (g), bed slope (S0).

ht

hu

Figure 7.1 Schematic Diagram of Hydraulic Jump


The general function relationship between the above variables can be written as:

f(L, ,hu,ht, u, g,S) = 0.……………….………………………………………………..(20)

Using the dimensional analysis, the π terms obtained are, π1 = L / hu, π2 = ht / hu,
π3 = u2 / ghu = Fr, and π4 = S0. These π terms may be arranged in the following non
dimensional form:

⁄ ⁄ ……………..……………………………………………. (21)

The general form of equations relating a dependent π-term with a number of


independent π-terms using the regression analysis in this work is in the form of the
product of powers of relevant π terms, i.e.,
…………………………...……………………… (22)

Equation (22) can be transformed to a linear expression by taking logarithms of both


sides of the equation, as follows:

…………..…… (23)

Finally, the equation can be rewritten in matrices form as follows:

Where, N is the number of observations. Then, the values of the parameters


C, , ,…, are obtained and can be replaced back into equation (22).

If the term ⁄ is taken as a dependent term, the equation form will be as follows:

( ) ( ) ……………………………..…….……… (24)

Equation (24) was used to predict the location of hydraulic jump. The predicted values
were compared with the simulated values to check the validity of the above equation.
The comparison between the predicted and observed values shows a good agreement as
illustrated in Figure 7.2.

35
The empirical equation is not valid when the bed slope is zero. So the experiments
conducted with zero bed slopes are not used. As most of the natural channels always
have small slope it plays a role in determining the jump location. The equation may not
give satisfactory results for other experimental work as the equation has been developed
based on a small number of experiments. But as it can be observed that the numerical
model can predict the jump parameters reasonably well there is scope for further work
to develop empirical equations like (24) from the numerical data using which the jump
location may be predicted without high speed computers.

Figure 7.2 Comparison between the numerical and empirical results.

8.0 SIMULATION BY CFD SOFTWARE

The Navier-Stokes Solver ANSYS CFX is employed to simulate the hydraulic jump.
The results from the simulation are then analysed to see whether there is any agreement
with the results from the numerical model. The VOF model is used to capture the water-
air interface.
8.1 Volume of fluid method (VOF)

The VOF method is appropriate for flow where immiscible fluids have a clearly defined
interface. The VOF method was introduced by Hirt and Nichols (1981) and has been
effectively used in numerical simulation of free surface flows.

The method assumes that each control volume contains just one phase (or the interface
between phases). For volume fraction of kth fluid, three conditions are possible:

 ek = 0 if cell is empty (of the kth fluid);


 ek = 1 if cell is full (of the kth fluid);
 0 < ek < 1 if cell contains the interface between the fluids.

Tracking of interface(s) between phases is accomplished by solution of a volume


fraction continuity equation for each phase:

…..……………………………………………………………. (25)

It solves one set of momentum equations for all fluids.

( ) ( ) ( ) ……………..……… (26)

For turbulent flows, single set of turbulence transport equations solved.

8.2 Simulation Procedure

Due to unavailability of time the simulation was done for two experimental conditions.
A geometry of length = 5.0 m, height = 0.45 m, and width w = 0.02 m was imported to
CFX-Pre together with the mesh file. Tetrahedral mesh was used for this analysis. The
mesh quality was checked by parameters like skewness, orthogonality and aspect ratio.
Among the two types of options i.e. steady and transient analysis - transient type was
chosen. The simulation was done for 100s with a time step of 0.025s.

37
Figure 8.1 Cross-Section of the Computational Grid
turbulence modeling was used. Gravity in the x- and y- directions was specified.

The boundary conditions and the initial conditions were specified using CFX expression
language (CEL). At the inlet boundary the water and air volume fractions were specified
according to the upstream water depth along with the inlet velocity. Similarly for the
outlet boundary the volume fractions were specified and the outlet static pressure was
given. The bottom bed was assigned no slip wall boundary condition. For the top
boundary volume fraction of air was assigned one and that for water was assigned to
zero.

The simulation was initialized by setting the upstream volume fractions, inlet velocity
and static pressure at the upstream end.

8.3 Results

Two simulations one with Froude number 3.83 and zero slope and another with Froude
number 3.06 and slope 0.02174. The contour lines for the water volume fraction are
shown in Figure 8.2-8.3
Figure 8.2 Contour lines of water volume fraction showing the interface as the jump
profile for Froude number 3.83 and slope-0

Figure 8.3 Contour lines of water volume fraction showing the interface as the jump
profile for Froude number 3.06 and slope-0.02174

39
The jump profile obtained from CFX analysis matched fairly with the experimental
observation. The profile near the jump is not smooth, there is a lot of oscillation near the
jump location so the exact location is hard to determine. The jump could be said to be
located where the gradient was steep and for Froude number 3.06 and slope 0.02174 it
was at a distance of 2.2m to 2.4m from the upstream end, which matches closely with
experimental observation. Similarly for Froude number 3.83 and slope 0 the jump is
located at 1.4 m to 1.5m distance from the upstream end which also matches with the
experimental result. As the number of simulations is only two no further correlations
could be drawn.

9.0 SUMMARY AND CONCLUSION

A numerical model is presented to simulate the hydraulic jump numerically in a


rectangular channel. Data obtained by laboratory tests for Froude numbers ranging from
2.17 to 7.0 for three different slopes 0.0475, 0.02174, 0 are used to verify the numerical
models.

The one-dimensional Saint Venant equations are solved numerically to simulate the
hydraulic jump. By starting with the computed initial conditions, these equations are
solved in time subject to appropriate boundary conditions until a final steady state is
reached.

The MacCormack‟s scheme (second-order accurate both in time and space) was used.
As with most of the higher-order methods, high-frequency oscillations were produced
near the hydraulic jump. These were smoothed by introducing artificial viscosity.

The results obtained from the numerical model were compared with the experimental
data. For both the non-zero slopes the jump location was downstream of the location
obtained experimentally. The jump location is simulated better at lower Froude numbers
and matches closely for Froude numbers 3.83 and less.

Based on the jump locations predicted numerically a generalised empirical equation was
developed which relates the jump location with the other parameters. The equation is
valid for non-zero slopes. Though the equation may not give satisfactory result for other
experimental work as the equation has been developed based on a small number of
experiments but this is an area where further work is can be done to establish an
empirical relationship between the hydraulic jump parameters.

The CFD software ANSYS CFX was used to simulate the jump for two Froue numbers.
The results match closely with experimental observations. As the number of simulations
was only two therefore no further correlations could be drawn.

10.0 FURTHER SCOPE

There remain a lot of aspects which can be further studied:

 The numerical model is second order accurate in space and time which cannot
predict the jump location for higher Froude numbers precisely. So fourth order
model can be developed to check the accuracy.
 The empirical equation is not generalized. Further work can be done to form a
generalized empirical relationship between the jump parameters.
 The ANSYS CFX simulation should be studied in further details, preferably by
simulating all the experiments.

11.0 REFERENCES

 Abbott, M. B., Marshall, G., and Rodenhius, G. S. (1969). "Amplitude-dissipative


and phase-dissipative scheme for hydraulic jump simulation." 13th Congress
IAHR, (Int. Assoc. of Hydr. Res.). Tokyo, Japan, 1, 313-329.
 ASCE Task Force. (1964). “Energy dissipators for spillways and outlet works.” J.
Hydr. Div., 90(1), 121-147.
 Basco, D. R. (1983). "Introduction to rapidly-varied unsteady, free-surface flow
computation." USGS, Water Resour. Invest. Report No. 83-4284. U.S. Geological
Service, Reston, Va.
 Bakhmeteff, B. A., and Matzke, A. E. (1936). "The hydraulic jump in terms of
dynamic similarity.” Trans. 101, 630-680.
 Belanger, J. B. (1828). “Essai sur la solution numeric de quelques problems
relatifs an mouvement permenent des causcourantes” (in French), Paris, France.
 Bidone, G. (1819). "Observations sur le hauteur du ressaut hydraulique en 1818."
Report (in French), Royal Academy of Sciences, Turin, Italy.

41
 Bradley, J. N., and Peterka, A. J. (1957). "Hydraulic design of stilling basins." J.
Hydr. Div., 83(5), 1401-1406.
 Chaudhry, M. H. (1993), “Open-Channel Flow”, Prentice Hall, Inc., New Jersey,
USA.
 Chippada, S., B. Ramaswamy and M. F. Wheeler (1994): “Numerical simulation of
hydraulic jump.” International Journal of Numerical Methods in Engg. 37, 1381-
1397.
 Chow, V. T. (1959), "Open Channel Hydraulics", McGraw-Hill Book Co., Inc.,
New York, USA.
 Garg, S. P., and Sharma, H. R. (1971). "Efficiency of hydraulic jump, "J. Hydr.
Div., 97(3), 409-420.
 Gharangik, A. M. and Chaudhry, M. H. (1991), “Numerical Simulation of
Hydraulic Jump", Journal of Hydraulic Engineering, 117, No.9, 1195-1210.
 Hager, W. H., and Wanoschek, R. (1987). "Hydraulic jump in triangular channel."
J. Hydr. Res., 25(5), 549-564
 Hamidreza,B et al.(2015): “Computational Modeling of the Hydraulic Jump in the
Stilling Basin with ConvergenceWalls Using CFD Codes.” Arab J Sci Eng 40:381–
395.
 Jameson, A., Schmidt, W., and Turkel, E. (1981). "Numerical solutions of the Euler
equations by finite volume methods using Runge-Kutta time-stepping schemes."
AIAA 14th Fluid and Plasma Dynamics Conf.: AlAA 81-1259, American Institute
of Aeronautics and Astronautics, Palo Alto, Calif.
 Katopodes, N. D., "A Dissipative Galerkin Scheme for Open Channel Flow", J.
Hydraul. Eng., 110, No. 4, 450-466, 1984.
 Leutheusser, H. J., and Kartha, V. (1972). "Effects of inflow conditions on
hydraulic jump.” J. Hydr. Engrg. Div., 98(8), 1367-1385
 MacCormack, R. W. (1969). "The effect of viscosity in hypervelocity impact
cratering." Paper 69-354, American Institute of Aeronautics and Astronautics,
Cincinnati, Ohio.
 McCorquodale, J. A., and Khalifa, A. (1983). "Internal flow in hydraulic jumps." J.
Hydraul. Eng., 109(5), 684-701.
 MacCowan, A. D. (1985). "Equation systems for modeling dispersive flow in
shallow water," XXIIAHR Congress, Int. Assoc, of Hydro. Res., Melbourne,
Australia, 2, 50-57.
 Rajaratnam, N. (1968). "Profile of the hydraulic jump." J. Hydr. Div., 94(3), 663-
673.
 Rajaratnam, N. (1967). "Hydraulic jump."Advances in Hydrosci, Academic Press,
4, 198-280.
 Rouse, H., Siao, T. T., and Nagaratnam, S. (1959). "Turbulence characteristics of
the hydraulic jump." Trans., 926-966.
 Safranez, K. (1929). "Researches relating to the hydraulic jump" (English
translation by D. P. Barnes), Bureau of Reclamation Files, Nos. 37 and 38, Bureau
of Reclamation, Denver, Colo.
 Sarma, K. V. N., and Newnham, D. A. (1973). "Surface profiles of hydraulic jump
for Froude number less than four." Water Resour. London, England, 25 (April),
139-142.
 Silvester, R. (1964). "Hydraulic jump in all shapes of horizontal channels." J. Hydr.
Div., 90(1), 23-55.
 Terrence, J.T. (1994), “Applied Numerical Methods for Engineering”, John Wiley
& Sons, Inc., New York, USA.

43

Potrebbero piacerti anche