Вы находитесь на странице: 1из 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/320811766

An alternative approach to the Von Karman vortex problem in modern


hydraulic turbines

Conference Paper · November 2017

CITATIONS READS

0 124

4 authors, including:

John Gummer
Hydro-Consult
68 PUBLICATIONS   129 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Drin river rehabiliation View project

Hydrogen economy View project

All content following this page was uploaded by John Gummer on 02 November 2017.

The user has requested enhancement of the downloaded file.


An alternative approach to the Von Karman vortex
problem in modern hydraulic turbines

Thomas Neidhardt Alexander Jung


Product Owner Turbine Head of Methods & Science
Voith Hydro Holding GmbH & Co. KG Voith Hydro Holding GmbH & Co. KG
Alexanderstr. 11 Alexanderstr. 11
89522 Heidenheim, Germany 89522 Heidenheim, Germany

Sebastian Hyneck John Gummer


Design Engineer Director
Voith Hydro Holding GmbH & Co. KG Hydro-Consult Pty Ltd
Alexanderstr. 11 15 McLeod St
89522 Heidenheim, Germany Rye, VIC 3941, Australia

Introduction
The traditional method of dealing with resonance resulting from Von Karman vortex shedding from the
trailing edges of hydraulic turbine components is to calculate the shedding frequency of the vortex street
using a suitable Strouhal number and compare this frequency with the natural frequencies of the component.
If the natural frequency was higher than the highest shedding frequency (say by 30%) then the component
was considered safe from vortex induced resonance.
In the past, Strouhal numbers were based upon experimental data possibly modified by boundary layer
analysis (the Universal Strouhal number approach) and the natural frequency of the component was
computed using approximate closed solutions. With the advent and universal use of Computational Fluid
Dynamics (CFD), shedding frequencies can be more accurately calculated than in the past and dynamic
Finite Element Analysis (FEA) will give a far more accurate values of natural frequencies than the closed
solutions, which with all their assumptions and approximations, ever could.
Whilst the time honoured and more modern computer approaches are perfectly adequate for bluff bodies
such as trash rack bars, they are increasingly less applicable for the analysis of the stay vanes and runner
blades of modern day turbines. In the past little attention was given to stay vane and runner blade thicknesses
and profiles but, with the increasing demand for higher and higher efficiencies, every component in a
modern day turbine, pump turbine and storage pump is finely tuned to give maximum efficiency. Whereas
30 years ago a peak efficiency of 95% was considered good for a large Francis unit, these days 97% and
even higher is within reach. In other words, during that period the hydraulic losses of the units were reduced
by as much as 40%.
Accordingly, in a modern hydraulic machine hydraulic profiles are aerodynamically shaped to minimise
turbulent boundary layer growth and, in particular, to restrict boundary layer separation, thus minimising
vortex generation at the trailing edge. This is very different to the profiles of yesteryear when stay vanes
were little more than cut plates or heavy castings and runner blades were of necessity thicker to
accommodate the unknowns involved in estimating hydraulic forces and the resulting blade stresses. With
modern CFD, hydraulic forces are known with far greater accuracy and with FEA the resulting stresses can
be more confidently calculated.
With these improvements in the design of hydraulic components the traditional methods of dealing with Von
Karman vortex induced resonance have been found to be inappropriate. In this paper the authors review the
various traditional approaches to the problem and detail the advantages of adopting a new approach in the
context of modern turbine design features.

1 Flow conditions at the trailing edges stayvanes or runner blades


Flow conditions in the stayvanes or runner blades of a typical hydraulic turbine are shown in Fig l. For
convenience a stay vane or runner blade will be referred hereinafter as “blade”. As an acceptable
approximation flow around a blade can be considered to be composed of two distinct regions, one a viscous
boundary layer region close to the vane surface, and the other, the general inviscid flow region along the
blade. The boundary layer flow has a negligible influence on the inviscid flow; however the growth of the
boundary layer is determined by the pressure gradients in the inviscid region.
The boundary layers commence at the stagnation point and, after a short distance of laminar growth along
the blade, the layers reach a point where their flow changes from laminar, to turbulent. This transition can
occur as a result of critical boundary layer thickness, or, if the boundary layer is close to the critical
thickness, can be triggered by abrupt changes in the blade profile or by surface defects.

Figure 1 – Flow conditions around a stay vane or runner blade

The turbulent layers continue to grow in thickness along the two surfaces on the blade, under the influence
of the inviscid flow, until the layers separate from the blade surfaces. Separation occurs either because the
layers reach a critical thickness, or because there is an abrupt change in blade geometry, to force separation.
The layers on each side of the blade need not necessarily separate at the same distance from the stagnation
point especially where there is no marked change in geometry to force separation to occur.
The location and flow conditions at the separation point are susceptible to perturbations in the flow external
to the boundary layer, which results in a well-defined vortex street issuing from the trailing edge of the
blade. The vortexes forming in the wake close to the blade (near wake) create variations in the hydraulic
pressure distribution acting upon the blades, the frequency of the resulting hydraulic forces being the same
as the frequency of the near wake vortexes. If the hydraulic forces are large enough and are oscillating at or
near the natural frequency of the vane then resonance will occur, resulting in large amplitude vibrations of
the vane which can lead to severe vibrations and possible cracking of stay vanes (References 1 and 2) and
runner blades.

Figure 2 a) - Experimental data Figure 2 b) - Schematic representation

Figure 2 – Fingerprint of von Karman vortices in a prototype Francis runner


Should the amplitudes be large enough, these too influence the location of the separation point, possibly
leading to a “lock in” phenomenon in which the shedding frequency of the vortexes are forced to be the
same as the natural frequency of the blades, over a wide range of free stream velocities (Reference 3). Unlike
stay vanes where the first torsion and bending modes are most likely to be excited, the complex geometry of
a Francis runner comes along with many natural frequencies in the relevant range for vortex shedding.
Consequently vortex induced vibrations occur at various distinct load-dependent frequencies. Figure 2 a)
above depicts the results of a short-time Fast Fourier Transform (FFT) of measurement data of a prototype
Francis turbine with a blunt trailing edge during load increase from part load to full load. At around t=150s,
a first “lock-in” effect can be observed with a frequency of approximately 1100 Hz. As the discharge (and
thus the flow velocity) is increased with time subsequently the shed vortices experience more “lock-ins” to
other natural modes of the runner structure. This is shown schematically in Figure 2 b) above.
Hence the propensity for the blade to be excited by the vortexes depends upon the frequency of the vortex
compared with the blade natural frequency, the intensity of the oscillating hydraulic forces acting on the
blade, and the effect of additional feedback from the movement of the vane and resonance of the hydraulic
channel. For any design situation these phenomena are considered separately in order to determine how
much is known and how well they can be predicted.

2 Von Karman Vortex shedding from bluff bodies and chamfered trailing edged
plates

2.1 Historical methods of calculating the frequency of vortex shedding


That the shedding frequency is one of the most important factors in determining whether or not stress
amplitudes in the vane will be high, can be seen from Fig 3 in which gain factor is plotted against the ratio of
shedding frequency coefficients. Typically the damping coefficient will be small, the only damping being
internal friction in the blade and churning in the water, hence the gain curve will be sharply peaked rising to
high values at a frequency ration of l.0.

Figure 3
Influence of frequency ratio
on amplitude gain

The natural frequency of the blade in air can be calculated to acceptable accuracy from standard formulae or
by finite element methods. The results thus obtained have to be modified to account for the effect of the
water. Most investigators have used a simple correction factor of .70, which has been confirmed by the
results obtained from tests on strain gauged blades. However upon closer investigation it becomes obvious
that this factor has a significant dependency on the profile shape and the mode type, i.e. bending or torsion.
The frequency of vortex shedding is calculated from the Strouhal Number, which in its simplest form is
defined as:
𝑓𝑑
𝑆 =  
𝑣
Where:-
𝑓 = frequency of vortex
𝑣 = some reference velocity (e.g. free stream velocity at the point of boundary layer separation)
𝑑 = some reference thickness (e.g. thickness of the wake at the point of boundary layer separation).
The Strouhal Number in this form is called the Universal Strouhal Number and as such should be
independent of geometry and (at Reynolds Numbers of interest) is independent of the Reynolds Number
calculated from wake conditions.
Flow conditions around bluff bodies are easier to define than those for flat plates with tapering trailing
edges. The bluff body has a unique, easily determined, separation point, and flow conditions at the point are
accordingly simple to calculate. Accordingly Strouhal numbers used in work on hydraulic turbines lean
heavily on experimental studies conducted upon bluff bodies such as spheres, cylinders and cones. These
have given reasonably consistent values of the Universal Strouhal Number in the range 0.18 to 0.25.
The relatively few investigations on vortexes from the trailing edges of chamfered plates do not have this
advantage of consistency of approach. In some cases the dimension of the trailing edge is used, others use
the thickness of the plate and seldom are the boundary layer thicknesses taken into consideration.
That consistency of approach in selecting the parameters for establishing Strouhal Number from
experimental data is important is evident from the results of Reference 4. As pointed out by the authors in
this reference, all the plates tested have the same characteristic dimension, but because of the differences in
trailing edge geometry, boundary layer separation point and possibly Strouhal Number the vortex shedding
frequency can differ by a factor of 2. Where the geometry of the trailing edge is such as to force separation
to occur at a well-defined point, such as in the plates of Reference 5 then the situation is far more akin to that
of a bluff body, and thus the flow conditions at the separation point are easier to define.
However despite the reported problems with accurately in determining Strouhal Numbers for chamfered flat
plates the usual procedure for checking the propensity for resonance and potential cracking of the blade is to
assume a Strouhal Number in the range 0.18 to 0.25. Then to calculate shedding frequency check its
correlation with the calculated natural frequency and, if the natural frequency is say 30% higher than the
shedding frequency at maximum turbine flow to allow for inaccuracies and lock-in, then to declare the blade
free from vortex induced problems.
In the past, despite all its potential pitfalls, this approach has been successful, probably because the design
methods for hydraulic turbines in the past have not been as exact as those demanded by today’s market and
the chamfering used was more akin to that of a bluff body than an aerofoil. However today we live in a very
competitive world where 0.10% of turbine efficiency typically involves a lot of money. Modern design tools
such as Computer Fluid Dynamics (CFD) and Finite Element Analysis (FEA) has resulted in finely honed
designs in which stay vanes and runner blades are more akin to aerofoils than those of yesteryear.
Chamfering of the trailing edges of these components is now part of the drive for higher efficiencies and
improved cavitation performance (Reference 6) and the inherent inaccuracies in the historical Strouhal
Number approach based upon bluff body work and sharp plate trailing edges as the means of ensuring the
absence of blade vibration is now questionable and a new design philosophy is needed.
The basic problem with the Strouhal Number approach is the assumption that it is constant regardless of
trailing edge geometry. As shown in Reference 4, although this is true for bluff bodies and sharp chamfers it
becomes less so for long slender tapering chamfers typical of today’s hydraulic designs.
The effect of thinning and fine tapering of the trailing edge of the stay vane of a large hydraulic turbine was
investigated by Voith Hydro using a dynamic CFD analysis. The geometrical configuration for the study is
shown in Figure 4.
Figure 4a) – Stayvane for CFD analysis Figure 4b) – Pressure monitoring points

Figure 4 – CFD Geometry and pressure monitoring points

The results of the study are shown in Figure 5.

Figure 5 – Frequencies, pressure amplitudes and Strouhal numbers for thin chamfered aerofoils
Evident from Figure 5 is the inescapable conclusion that Strouhal
number is not constant over the range of interest and below a trailing
edge thickness of 5 mm there is no established vortex street which
could promote resonance. These CFD results confirm those of
Reference 4 which are reproduced in Figure 6 i.e. as the trailing
edge is more finely tapered the frequency of the shedding vortex
increases, even though the Strouhal number and diameter are
typically considered to be constants.
Accordingly with very fine trailing edge thickness the historical
Strouhal Number approach is misleading at the best and incorrect at
the worst. Another approach is required which correctly takes the
true conditions at the trailing edge of a fine chamfer into
consideration. As will be shown in the next section the basis for this
approach can be seen in the results of CFD analysis of Figure 4 and
the accompanying boundary layer visualisations reproduced here as
Figure 7.

2.2 Intensity of the hydraulic forces


The blade will only be subjected to severe oscillations if the
fluctuating hydraulic forces are powerful enough. If the wake
vortexes are unorganized and very weak then oscillations of the
blade will remain within acceptable limits, even at apparently
resonant conditions. From the CFD results of Figure 5 it is evident
that below a chamfer thickness of about 6 mm the pressure Figure 6 – Shedding frequencies
amplitudes are negligible and obviously incapable of promoting
resonance. The reason for this can be seen in the CFD boundary visualisations of Figure 7. Here the typical
start of the vortex street is not evident below a trailing edge chamfer thickness of 12mm; the wake being
disorganised and having no structure typical of a vortex.

Figure 7 – Boundary layer flow visualisation at the chamfer trailing edge

That the shape of the trailing edge has a governing effect on the intensity of the wake induced forces evident
in the current CFD study has been experimentally proven. Figure 8 taken from Reference 7 shows a
summary of experimental results presented by various investigations and gives a qualitative indication of
wake intensity. Reference 8 has confirmatory curve from the work by Liess.
From Figure 8 it is evident that, as the s/d ratio is increased beyond 1.5 the vortex intensity drops to zero and
does not increase again. For trailing edges in this range it would appear that the wedge shaped trailing edge
for minimum wake intensity is one which puts the end of the vane within the near wake i.e. in the region of
formation of the first vortex. The vane then interferes with the formation of the vortex and leads to an
unorganized and weak vortex street.

Figure 8 – Vortex street intensity

On the basis of these results it is fair to question whether the chamfering solutions found for some resonant
problems of stay vanes and runner blades in the past were less due to a change of shedding frequency and
more because of a substantial reduction in shedding intensity.

3 Conclusions
A new CFD analysis and confirmatory experimental results from literature has shown that the common
assumption of a constant Strouhal Number to calculate the vortex shedding frequency from chamfered
trailing edges as typically used for the design of hydraulic turbine runners and stay vanes trailing edges is
incorrect for other than very coarse chamfers.
With the finely tuned profiles typical of today’s hydraulic turbine stay vanes and runner blades shedding
frequencies calculated using historical Strouhal Numbers from bluff body and sharp chamfer literature will
probably be at the best in error or at the worst dangerous.
Accordingly it is recommended that for fine chamfers above an s/d ratio of 1.5 the traditional resonant
frequency analysis be dispensed with and the fact that these profiles result in virtually zero dynamic forces at
the trailing edge (at least up to an impractical s/d ratio of 3.5) be relied upon to prevent undue vibrations of
stay vanes and runner blades.
For convenience this should be called the “low energy profile” approach as opposed to the historical
“frequency approach”.
References
1. J.H. Gummer and P.C. Hensman, “A review of stayvane cracking in hydraulic turbines”, Water Power and Dam
Construction, August 1992
2. R.K. Fisher Jr, J.H. Gummer and C. Liess “Stayvane vibrations in the Nkula Falls turbines”, Hydropower and
Dams”, January 1994
3. T.H. Toebes and P.S. Eagleson “Hydroelastic vibrations of flat plates related to trailing edge geometry” Journal of
Basic Engineering”, December 1961.
4. G. Hesleastad and D.R. Olberts “Influence of trailing edge geometry on hydraulic-turbine blade vibration
resulting from vortex excitation”, Journal of Engineering for Power, April 1960.
5. M.E. Greenway and C.J. Nood “The effect of a bevelled trailing edge on vortex shedding and vibration”, Journal
of Fluid Mechanics, Volume 61, part 2, 1973.
6. S. Etter, A. Otto and J.H. Gummer “The benefits of chamfering the trailing edges of Francis turbine blades”,
Hydropower and Dams, Issue 2, 2007.
7. Chen and Florjancic “Vortex induced vibrations on guide vanes with slender trailing edges”, Conference on
Vibrations and Noise in pump, fan and compressor installations, Inst. Mech. Engineers, Sept 16-18, Southampton
UK, 1975.
8. Liess C “The cause and avoidance of cortex excited vibrations of turbine stay vanes”, Voith research and
Construction Issue 32 Paper 6 Reprint t2664e, 1986.

The Authors

Thomas Neidhardt is Product Owner Turbine at the Voith Hydro’s Headquarter engineering center in Heidenheim,
Germany. He graduated from the University of Karlsruhe, Germany, in mechanical engineering and joined Voith Hydro
in the year 1996. Since then he has worked in the fields of structural and dynamic analysis of turbine components,
software development and mechanic design. From 2010 to 2016 he was heading the turbine system engineering team in
Voith Hydro with additional support to the product standardisation and turbine development.

Dr. Alexander Jung studied mechanical engineering at University of Stuttgart and Northwestern University and
graduated in 1993. He obtained his doctoral degree on the development and application of numerical methods for
unsteady flow prediction in turbo machines. He joined Voith in 2004 and worked several years as a hydraulic designer
for large Francis and pump turbine projects. After that he was heading the turbine development methods department
before he took over the responsibility for “Methods & Science” within Voith Hydro in 2016.

Sebastian Hyneck graduated in mechanical engineering at University of Applied Science in Ulm, Germany 2006 and
joined Voith Hydro in 2007. Since then he designed numerous turbine components of all types, collecting a large range
of experience, working in different Voith locations and intercultural teams. As mechanic lead design engineer he takes
over responsibility for system engineering of worldwide large hydro power units. In the recent years he is focused on
high head turbines and pump storage plants.

John Gummer was educated at Imperial College and Bristol and Southampton Universities in the UK, and has spent
more than 50 years working on major hydroelectric and pumped storage projects throughout the world. Since leaving the
Itaipu Hydroelectric Project where he was the Chief Mechanical Engineer of the IECO-ELC Co-ordination Group, he
has acted as consultant to the World Bank, ODA, and many leading engineering organisations and lending authorities.
He is on the Editorial Board of The International Journal on Hydropower and Dams and has authored over 60 papers on
hydropower and fluid dynamics.

View publication stats

Вам также может понравиться