Numerical Simulation of A Cavitating Draft Tube Vo

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Journal of Physics: Conference Series

PAPER • OPEN ACCESS Related content


- Numerical investigation of the upper part
Numerical simulation of a cavitating draft tube load vortex rope behaviour in a Francis
turbine
vortex rope in a Francis turbine at part load Z W Guan, Z N Wang, Y Z Zhao et al.

- Unsteady hydraulic simulation of the


conditions for different σ-levels cavitating part load vortex rope in Francis
turbines
J Brammer, C Segoufin, F Duparchy et al.
To cite this article: J Wack and S Riedelbauch 2017 J. Phys.: Conf. Ser. 813 012019
- Investigation on pressure fluctuation in a
Francis turbine with improvement
measures
J J Feng, W F Li, H Wu et al.

View the article online for updates and enhancements.

This content was downloaded from IP address 139.81.250.246 on 15/02/2019 at 18:44


Hyperbole IOP Publishing
IOP Conf. Series: Journal of Physics: Conf. Series 813 (2017) 012019 doi:10.1088/1742-6596/813/1/012019

International Conference on Recent Trends in Physics 2016 (ICRTP2016) IOP Publishing


Journal of Physics: Conference Series 755 (2016) 011001 doi:10.1088/1742-6596/755/1/011001

Numerical simulation of a cavitating draft tube vortex rope in


a Francis turbine at part load conditions for different σ-levels

J Wack and S Riedelbauch


Institute of Fluid Mechanics and Hydraulic Machinery, University of Stuttgart,
Pfaffenwaldring 10, 70555 Stuttgart, Germany

E-mail: jonas.wack@ihs.uni-stuttgart.de

Abstract. In the present study, numerical investigations of a Francis turbine at model scale are
performed. A part load operating point is selected for the analysis with a focus on the
cavitating draft tube vortex rope. For three σ-levels, different aspects are investigated. First, the
extension of the vapor volume is analyzed and compared to high-speed visualization
recordings from the measurements. Furthermore, the pressure fluctuations for different σ-levels
are analyzed and compared to experimental data at a measuring point in the draft tube cone.
Finally, the impact of the different cavitation volumes on the axial and circumferential velocity
distribution is investigated. In addition, the SST and SAS turbulence models are compared for
two pressure levels.

1. Introduction
The safe integration of renewable energies like photovoltaics is a key issue of nowadays energy
market. Hydro power plants play an important role for this topic as they can be easily regulated.
Nevertheless, this comes across with the fact that turbines are more and more operated at off-design
conditions which are characterized by strongly swirling flows and the occurrence of cavitation.
In a Francis turbine, different cavitation phenomena can occur depending on the operating point
[1]. At part load conditions, a helically shaped vortex rope forms in the draft tube cone. For low σ-
levels, the pressure in the vortex core falls below the vapor pressure which results in cavitation. In
experiments the appearance of cavitation allows to visualize the shape of the vortex rope (e.g. [2, 3]),
which can be used for comparison with simulation results. Typically the frequency of the vortex rope
rotation is between 0.2 and 0.4 times the runner frequency [4]. Favrel et al. [5] observed that the
precession frequency is a function of the discharge factor while the σ-level has no significant impact.
The occurring pressure oscillations depend on the discharge, the σ-level and the Froude number [5, 6].
Furthermore, the presence of a cavitation volume significantly reduces the natural frequency of the
system [7]. When the natural frequency and the precession frequency coincide it comes to resonance
conditions with high pressure fluctuations.

2. Modelling and numerical setup


The simulation of cavitating flows is more complex compared to single phase flows. In the scope of
this study, the homogenous model is applied for the two-phase treatment which assumes that the liquid
and vapor phase share a common pressure and velocity field [8]. As cavitation model, the Zwart model

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution
of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
Published under licence by IOP Publishing Ltd 1
Hyperbole IOP Publishing
IOP Conf. Series: Journal of Physics: Conf. Series 813 (2017) 012019 doi:10.1088/1742-6596/813/1/012019

is used and the four model parameters are set according to Zwart et al. [9]. Furthermore, for URANS
simulations the widely used SST turbulence model is applied to close the system of governing
equations. Additionally, two simulations are performed with the SAS-SST model which can operate in
scale resolving simulation mode [10].
The simulations are performed with Ansys CFX on a Francis turbine at model scale. The
simulation domain is ranging from the spiral case inlet to the draft tube outlet. For all investigations a
structured mesh with approximately 25 million cells is used. Most of the cells are located in the draft
tube as the main flow phenomenon is located there.
For the treatment of the interfaces between stationary and rotating domains, a transient rotor stator
approach is applied. As inlet boundary condition, a constant total pressure is set and at the outlet an
opening boundary condition is prescribed. The total pressure at the inlet and static pressure at the
outlet are set that both the required discharge and pressure level in the turbine are met. For the spatial
discretization, a high resolution scheme is used for all simulations performed with the SST turbulence
model and a bounded central differencing scheme is applied for the simulation with the SAS model.
As temporal discretization, a second order backward Euler scheme is used for all setups.

3. Numerical results
In the scope of this study, an operating point at part load conditions is investigated for three different
σ-numbers (0.06, 0.09 and 0.11). The discharge of the examined operating point is 81 % compared to
best efficiency point conditions.

3.1. Cavitating vortex rope


The regions where cavitation occurs can be visualized with an isosurface of the void fraction. In
figure 1 the simulation results for different cavitation numbers and turbulence models are displayed
and can be compared to high-speed visualization recordings from the measurements. Additionally, in
the draft tube cone two lines are shown that are used for the evaluation of the velocity profiles.

σ=0.06, SST σ=0.09, SST σ=0.09, SAS σ=0.11, SST σ=0.11, SAS

Figure 1. Cavitating vortex rope visualized with an isosurface of the void fraction (value 0.25)
top: measurement – bottom: simulation results

2
Hyperbole IOP Publishing
IOP Conf. Series: Journal of Physics: Conf. Series 813 (2017) 012019 doi:10.1088/1742-6596/813/1/012019

As expected, the cavitation volume increases with reduction of the σ-number. For the lowest
pressure level, a large vapor volume forms in the draft tube cone which is almost axisymmetric in the
upper part for both simulation and measurement. The typical helical shape for part load is only
observable in the lower region of the cone where the diameter of the vortex rope is already
significantly smaller. In addition, cavitation occurs on the suction side of the runner blades close to the
trailing edge due to the very low pressure level.
For σ=0.09, a comparison of the results obtained with the different turbulence models shows that
the SAS model resolves smaller turbulence structures that are noticeable in the shape of the cavitation
volume. Especially in the lower region of the draft tube cone where the vapor phase decays,
differences are clearly observable to the SST results. Nevertheless, in terms of vapor volume size only
minor differences are present between both turbulence models. This is in contrast to observations
made for a deep part load operating point where the SAS model predicted a significantly bigger
cavitation volume of the inter-blade vortices in the runner compared to the SST model [11]. This can
be explained by the fact that the occurring vortex rope at part load conditions has a much bigger
diameter compared to the inter-blade vortices. Consequently, a different resolution of the pressure
minimum in the vortex core caused by the turbulence model plays a less significant role at part load
condition.
For the highest pressure level similar observations like for σ=0.09 can be made. In the
measurement it can be noticed that the cavitating vortex rope decays into small structures and vapor
bubbles. This can be better reproduced by the SAS model. Nevertheless, the chosen two-phase
modeling approach is not suitable to resolve single vapor bubbles.

3.2. Pressure fluctuations


The pressure fluctuations for σ=0.09 and σ=0.11 are evaluated at a measuring point in the draft
tube cone. The results are displayed in figure 2. Both turbulence models can reproduce changes in
amplitude that are present especially for σ=0.09. While for the lower σ-level the amplitude of the
measurement is well met except for the last cycle, for σ=0.11 the simulation results slightly
underestimate the occurring pressure variations. In terms of frequency, both models are in good
agreement with the measurements over a wide range. However, the SAS model tends to overestimate
the frequency for cycles with lower amplitude.
A comparison of the cases with different σ-numbers shows that the pressure level has a significant
impact on the amplitude of the pressure fluctuations that has also been observed by Favrel et al. [6].
Due to the fact that the investigated σ-levels are below the resonance conditions, the pressure
fluctuations decrease with the σ-number. Between σ=0.09 and σ=0.11 the amplitude changes by a
factor of approximately two. This is on the one hand caused by getting away from resonance
conditions. On the other hand the increased cavitation volume leads to a damping of the pressure
fluctuations. The results suggest that in the presence of a cavitating vortex rope, two-phase simulations
are necessary to correctly predict the occurring pressure amplitudes.

meas SAS SST meas SAS SST


0.12 0.16
0.15
0.11
0.14
p/ρgH [-]

p/ρgH [-]

0.1 0.13
0.12
0.09 0.11
0.1
0.08
0.09
0.07 0.08
0 2 4 6 8 10 12 0 2 4 6 8 10 12
t x n [-] t x n [-]

Figure 2. Pressure at measuring point in the draft tube cone for σ=0.09 (left) and σ=0.11 (right)

3
Hyperbole IOP Publishing
IOP Conf. Series: Journal of Physics: Conf. Series 813 (2017) 012019 doi:10.1088/1742-6596/813/1/012019

3.3. Axial and circumferential velocity distribution in the draft tube cone
The influence of the σ-level on the axial and circumferential velocity distribution in the draft tube
cone is investigated along two lines that are displayed in figure 1. Line 1 and 2 are located 0.39 and
1.02 times the outer runner outlet diameter below the runner outlet. Measurement data is available for
two σ-numbers close to the investigated pressure levels. The results are presented in figure 3.
At line 1, the results of the averaged axial velocity cax for σ=0.11 show a good agreement with the
measurements over a wide range. Deviations can be found close to the draft tube wall which might be
a result of neglecting the flow through the runner side gap on the crown side and in the stagnant region
in the center. There, the simulation is forming a plateau while for the measurement at σ=0.12 the
backflow increases to the center of the draft tube. This can be explained by the occurrence of a region
with permanent cavitation in the middle of the measuring line for the slightly lower pressure level in
the simulation. At σ=0.10 the measurement shows a similar course compared to the simulation. Close
to the center no data is available due to measuring problems caused by the presence of vapor.

line 1 line 2
meas sig010 SST sig011 SST sig006 meas sig010 SST sig011 SST sig006
meas sig012 SAS sig009 meas sig012 SAS sig009
SAS sig011 SST sig009 SAS sig011 SST sig009
0.6 0.4
0.4
0.2
0.2
Velocity cax/cref [-]

Velocity cax/cref [-]

0 0
-0.2 -0.2
-0.4
-0.6 -0.4
-0.8 -0.6
-1
-0.8
-1.2
-1.4 -1
-1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1 -1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1
Radius R/Rref [-] Radius R/Rref [-]

1.5 1
0.8
1 0.6
Velocity cu/cref [-]

Velocity cu/cref [-]

0.5 0.4
0.2
0 0
-0.2
-0.5 -0.4
-1 -0.6
-0.8
-1.5 -1
-1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1 -1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1
Radius R/Rref [-] Radius R/Rref [-]

Figure 3. Axial (top) and circumferential (bottom) velocity distribution along two lines in the draft
tube cone averaged over ten vortex rope revolutions

With a reduction of the pressure level, the extension of the stagnant region increases as the
cavitation volume grows. For σ=0.06, the region of backflow is huge and occupies almost the inner
half of the draft tube cone. This results in a significantly higher axial velocity in the outer part of the
draft tube cone which can be explained by the conservation of mass. A comparison of the different
turbulence models shows that the biggest differences between the SAS and SST model are observed in
the outer region close to the draft tube wall.
Further downstream (at line 2), for σ=0.09 and σ=0.11 a good agreement of the axial velocity with
the measurements can be stated. Due to the helical shape and smaller diameter of the vortex rope in
that area, there is no region with permanent cavitation which results in the different course of the
velocity distribution in the stagnant region. The biggest deviation to the measurement again can be

4
Hyperbole IOP Publishing
IOP Conf. Series: Journal of Physics: Conf. Series 813 (2017) 012019 doi:10.1088/1742-6596/813/1/012019

found close to the draft tube wall. Contrary to the cases at higher pressure levels, the axial velocity
distribution for σ=0.06 looks different. The reason is that a region with permanent cavitation persists
which has an impact on the velocity field.
In the circumferential velocity distribution, the impact of cavitation can also be determined. All
simulation results at line 1 show a maximum circumferential velocity at a dimensionless radius of
around 0.5. Going to the center of the draft tube, there is first a steep velocity gradient which
significantly reduces around a dimensionless radius of approximately 0.25. Contrary to this, the
measurement with σ=0.12 has a higher velocity gradient in the center and no change in slope. At line 2
the results for σ=0.09 and σ=0.11 show the same characteristics like the measurements while the low
pressure level still shows the change in slope in the center due to the presence of permanent cavitation.

4. Conclusion and outlook


Cavitation simulations were carried out on a Francis turbine at model scale for a part load operating
point at different σ-levels. As expected, the vapor volume increases with decreasing pressure level.
The axial and circumferential velocity distribution is affected by the σ-level. With increasing
cavitation volume the stagnant region is getting bigger. Furthermore, the occurrence of regions with
permanent cavitation significantly influences the shape of the velocity profiles. Moreover, the σ-level
affects the amplitude of the pressure fluctuations which necessitates the usage of two-phase
simulations. Below resonance conditions the amplitude decreases with reduction of the σ-number. The
results of the SST and SAS model show a similar size of the cavitation volume. Nevertheless, the SAS
model can resolve smaller vortex structures which results in differences in the shape of the vortex rope
especially in the region of decaying vapor volume.
For further improvement of the simulation results the use of more complex two-phase models
should be investigated in the future.

Acknowledgments
The research leading to the results published in this paper is part of the HYPERBOLE research
project, granted by the European Commission (ERC/FP7-ENERGY-2013-1-Grant 608532). Special
thanks to EPFL, in particular Keita Yamamoto, Andres Müller, Arthur Favrel, Christian Landry and
François Avellan, for providing all experimental data enabling comparisons presented here.
Furthermore, the authors would like to thank the HLRS Stuttgart within the Bundesprojekt framework.

References
[1] Avellan F 2004 6th Int. Conf. on Hydraulic Machinery and Hydrodynamics Timisoara Romania
[2] Favrel A, Müller A, Landry C, Yamamoto K and Avellan F 2015 Exp. in Fluids 56(215)
[3] Houde S, Iliescu M S, Fraser R, Lemay S, Ciocan G D and Deschênes C 2011 Proc. ASME-
JSME-KSME 2011 Joint Fluids Engineering Conf. pp 171-182
[4] Nicolet C, Zobeiri A, Maruzewski P and Avellan F 2010 IOP Conf. Series: Earth and Env. Sci.
12 012053
[5] Favrel A, Müller A, Landry C, Yamamoto K and Avellan F 2016 Proc. 28th IAHR Symp. on
Hydraulic Machinery and Systems Grenoble France
[6] Favrel A, Landry C, Müller A, Nicolet C, Yamamoto K and Avellan F 2014 IOP Conf. Series:
Earth and Env. Sci. 22 032035
[7] Landry C, Favrel A, Müller A, Yamamoto K and Avellan F 2014 IOP Conf. Series: Earth and
Env. Sci. 22 032037
[8] Bouziad Y A 2006 Ph.D. Thesis EPFL Lausanne
[9] Zwart P J, Gerber A G, Belamri T 2004 Int. Conf. on Multiphase Flow Yokohama Japan
[10] Menter F R, Schütze J and Gritskevich M 2012 Progress in Hybrid RANS-LES Modelling ed S
Fu, W Haase, S H Peng and D Schwamborn (Berlin Heidelberg: Springer) pp 15–28
[11] Wack J, Riedelbauch S, Yamamoto K and Avellan F 2016 Proc. 9th Int. Conf. on Multiphase
Flow Firenze Italy

You might also like