Sei sulla pagina 1di 20

Acta Mech 201, 211–229 (2008)

DOI 10.1007/s00707-008-0060-4

Helmut Keck · Mirjam Sick

Thirty years of numerical flow simulation in hydraulic


turbomachines
Dedicated to Professor Wilhelm Schneider on the occasion of his 70th birthday

Received: 6 February 2008 / Revised: 23 May 2008 / Published online: 2 September 2008
© Springer-Verlag 2008

Abstract The application of computational fluid dynamics (CFD) in the design of water turbines and pumps
started about 30 years ago. This paper reviews the main steps and breakthroughs in the methods that were made
during this period, through the eyes of one particular water turbine company which spear-headed some of the
first developments for practical applications. Practical examples are used to illustrate the developments of the
tools from 1978 to 2008 and to give an overview of the complete revolution in hydraulic turbine design that has
occurred over this time. Several periods with distinct levels of complexity, and hence accuracy of the physical
models and of the simulation methods can be distinguished. The first steps coincided with the introduction
of the Finite Element Method into CFD, and were characterized by simplified Quasi-3D Euler solutions and
Fully 3D potential flow solutions. Over the years the complexity continuously increased in stages: via 3D
Euler solutions, to steady RANS simulations of single blade passages using finite volume methods, extending
to steady simulations of whole machines, until today unsteady RANS equations are solved with advanced
turbulence models. The most active areas of research and development are now concerned with including the
effects of 2-phase flows (free surface flow in Pelton turbines and cavitation) and fluid–structure interaction.

1 Introduction

In 1978, when a young engineer started to develop and apply numerical methods for the hydraulic design of
water turbines and pumps, there were some tremendous challenges: up to then the software and hardware had
been rather rudimentary and only 2D geometries could be treated by conformal mapping or by the singularity
method. But in the 1970s numerical methods, such as finite difference and finite element methods, had made
big progress, first in aerodynamics and then in structural mechanics. The challenge at the time was to intro-
duce these new numerical methods into the traditional world of hydraulic turbomachines with its emphasis on
experimental development. An overview over the different types of hydraulic turbines and their design aspects
is given by [6].
The name CFD was already known in 1978, having been coined in the title of Roache’s book [37], and the
key developments in the k-ε turbulence model had already been made [28]. At that time however CFD was
not known at all by real turbine designers, and was far less accepted as an engineering tool. It is interesting to
look back to the first steps of CFD applications to hydraulic turbines and to imagine the complete revolution
in the design technologies since then.

H. Keck (B) · M. Sick


Andritz VA TECH HYDRO Ltd, Hardstrasse 319, Postfach 2621, 8021 Zurich, Switzerland
E-mail: helmut.keck@vatech-hydro.ch

M. Sick
E-mail: mirjam.sick@vatech-hydro.ch
212 H. Keck, M. Sick

Fig. 1 Main sections of a bulb turbine where one half of the symmetric intake is modelled by FEM

2 The potential flow and Q-3D period (1978–1987)

The first success of CFD in modelling turbine flows came from the fact that suddenly it was possible to treat
very complex geometries with good accuracy. It was the step from “simple equations in simple geometries”
to “simple equations in complex geometries”.
The finite element method (FEM) seemed to be ideally suited as it had the natural capability to model
complex geometries by using triangular elements or rectangular 8-node parabolic elements. The FEM-codes
for structural analysis were solving the 3D-Laplace equation:
 = 0 in (x, y, z). (1)
And by interpreting  as the velocity potential
grad = c (2)
these codes were able to solve potential flow problems both in 2D and 3D.
The first step of using FEM for a CFD application was just to program a post-processing tool that allowed
velocity components (and static pressure) to be calculated from the -distribution over the nodal points of the
grid and to formulate proper boundary conditions.
The specification of boundary conditions turned out to be the most critical task in order to get useful results
(which, by the way, is still the case today even with the most advanced simulation methods!).
First results were quite promising, see [18]. The flow in a bulb turbine intake which is shown in Figs. 1, 2,
3 and 4, was actually well suited to such methods as the assumption of a potential flow was justified and the
3D geometry was rather complex. The transition from a rectangular to a circular cross-section and finally to a
ring section could be treated with good accuracy in spite of a very coarse grid.
Much more important than intakes are the stationary and rotating blade rows. For axial turbines the quasi-3D
approach with calculating the meridional flow and the blade-to-blade flow separately is a reasonable approx-
imation and was again a nice application of the FEM method as shown by [19]. By combining 3D-intake
modelling and quasi-3D blade row modelling, the flow in a complete bulb turbine (except the outlet draft tube)
was calculated [16].
It should be mentioned that the flow in the blade rows deviates (sometimes quite significantly) from a
potential flow, even if the flow is considered as inviscid. The reason is that the circulation of the blades is
usually not constant between hub and shroud and that between the inlet and outlet of a (3D) blade row shed
vorticity is generated. By introducing stream functions both in the meridional and the blade-to-blade analysis
it was possible to calculate such rotational (non-potential) flows by a quasi-3D approximation of the 3D-Euler
equations.
Thirty years of numerical flow simulation in hydraulic turbomachines 213

Fig. 2 Three-dimensional potential flow analysis of the intake (main section through FEM grid)

Fig. 3 Selected elements of the 3D FEM grid in the transition between rectangular and circular cross section

The results obtained showed very good agreement with measurements and clearly indicated the strong
(non-linear) influence of the swirl on the flow between blade rows of Kaplan turbines, see [20] and [21].
In this period it was clear that inviscid CFD simulations could only be reasonably accurate close to the
best efficiency point. Off-design calculations were beyond the state-of-the-art. In spite of these limitations the
quasi-3D Euler calculations already provided useful guidance for turbomachinery runner designs [17].

3 The decision towards 3D-Euler

Based on the first success with above mentioned CFD-methods the strategy for further refinements had to be
defined, especially for radial runners of Francis turbines and for off-design conditions. The following options
were at hand:
214 H. Keck, M. Sick

Fig. 4 Cross section perpendicular to the axis of the machine

(A) Add 3D-boundary layer calculations to the inviscid solutions (which were either 3D-potential flows or
quasi-3D rotational flows).
(B) Develop a 3D-Euler code for incompressible flows in complex geometries.
(C) Step into the emerging technology of solving the Reynolds-averaged (turbulent) Navier–Stokes equations.
Option B was clearly superior for the following reasons:
• the flow inside radial runners (Francis turbines) is not well approximated by the quasi-3D method and is
by no means irrotational. Adding boundary layer calculations to an incorrect inviscid solution would not
provide any progress.
• in turbine runners the flow is accelerating and the Reynolds number is very high (thin boundary layers).
The major 3D-effects are not driven by viscosity but by vorticity → and inviscid, so that a fully 3D Euler
simulation is the proper approach.
• The step into Navier–Stokes simulations was too early as the methods were still in their infancy, and was
postponed until real viscosity driven flows (diffusers, separated boundary layers, vortices, etc.) could be
calculated.
In 1983 a research project together with the technical university of Lausanne (EPFL) was launched to develop
a 3D-Euler code for Francis runners. This development was based on a time marching code for transonic
flows that had previously been developed by Rizzi. Its adaptation and modification for incompressible flows
in hydraulic turbines was a long process. In 1987 the break-through was achieved: Göde et al. published the
first successful 3D-Euler simulation in a Francis runner, see [12].

4 The 3D-Euler period (1987–1994)

Thanks to the fully 3D-Euler analysis it was possible to demonstrate that the flow inside a Francis runner by
no means followed the axial symmetric stream-sheets as assumed by the classical quasi-3D methods but that
a complex “twist of stream sheets” was obtained, see Fig. 5. With this method it was also possible to perform
Thirty years of numerical flow simulation in hydraulic turbomachines 215

Fig. 5 Visualisation of streamlines of the 3D Euler

Fig. 6 Leading edge vortex modelled by the 3D Euler code

off-design calculations with high accuracy as all the vorticity-driven secondary flows were modeled correctly.
And as long as the real flow was not dominated by viscous effects, the non-viscous but fully rotational 3D-Euler
simulation showed very good results. An example for a vorticity driven secondary flow nicely modelled by
3D-Euler is shown in Fig. 6: the leading edge vortex, often observed in test rigs at off-design conditions, was
numerically modeled for the first time by [13].
Of course, qualitative comparisons with experiments were important to aid understanding, but a big effort
went into the detailed quantitative validation of the numerical results with experimental data. Comprehensive
L2F (Laser two focus) measurements at the runner inlet and outlet with scanning in circumferential direction
were performed, Fig. 7 and published by Göde et al. in 1989 [14].
Figure 8 shows the detailed variation of velocity components in circumferential direction. The upstream
effect of the leading edge is stronger in the 3D-Euler simulations than the measurements, whereas the wake-
effect downstream of the blade trailing edge is more pronounced in the L2F results (as to be expected).
Except for some local effects the agreement for the average velocity components was excellent.
In the late 1980s the progress in CFD was so fast that a fair international comparison of the accuracy
against experiment was launched. Test data of a Francis turbine from EPFL were made available as boundary
conditions for a CFD analysis. In the GAMM workshop 1989 in Lausanne the numerical results of more
than ten international CFD teams were presented and compared. The agreement at runner inlet and outlet was
generally good, but only the 3D-Euler analysis based on a stacking technique between the stationary upstream
part and the rotating runner part as presented by [13] predicted the low pressure peak at runner inlet correctly,
Fig. 9. The pronounced inlet edge cavitation observed on the EPFL-test rig for this “GAMM-runner” was not
known to any of the CFD-teams and only one of them was able to predict it. Interesting: the winning team
calculated its own inlet boundary conditions by means of its stacking technique, whereas all the other teams
216 H. Keck, M. Sick

Fig. 7 Laser 2 focus (L2F) measurements at inlet and outlet of a Francis runner

Fig. 8 Comparison of L2F measurement results with 3D Euler calculations

took the measured inlet velocity profiles which were circumferentially averaged. So again: handling boundary
conditions is crucial and even measured boundary conditions can be a pitfall (especially if the averaging of
data suppresses important information). The possibility to model the flow in all kinds of turbine runners with
high accuracy in a wide operating range opened up completely new opportunities in the hydro industry. The
transfer of the 3D-Euler analysis from a research tool to a practical design tool was a challenging and exciting
task [23].
Thirty years of numerical flow simulation in hydraulic turbomachines 217

Fig. 9 Pressure distribution at runner inlet as predicted with Euler code and stacking technique

In this period the hydraulic industry was changing: fewer new plants were being built, but a tremendous
demand for rehabilitation and upgrading of old plants was emerging. The new design tools were ideally suited
for the new challenges: it was no longer sufficient to provide new designs by interpolation and minor evolution
from previous well proven and tested designs, now new runners with typically more than 20% power increase
over the existing runner had to be designed that also had to match the often poor and hardly understood existing
stationary parts. In addition, the price of one medium size new runner was of the same order of magnitude
as a full model test. The pressure to supply new replacement runners based on CFD only, without any model
tests, increased. Again, a drastic change in the industry took place: the upgrading business started to follow
a step-by-step approach whereby the first step was a contractual CFD-analysis of the existing unit, with a
comparison between the new and the old runner by CFD (partly supported by model tests) as a second step.
The new approach, firstly promoted by [22], pushed CFD-analysis into the role of THE key pacing technol-
ogy of the hydro industry. This role was then continuously consolidated by the introduction of Navier–Stokes
codes for turbulent flow computation in the practical design work.

5 Towards the virtual test rig (1990–2000)

5.1 Steady state Reynolds averaged Navier–Stokes equations

From 1990 on, finite volume methods, which are inherently conservative with respect to mass and momen-
tum, became the dominant tools for solving the Reynolds averaged Navier–Stokes (RANS) equations in water
turbine flow analysis and design. All examples discussed in the following sections have been analyzed by
applying commercial software packages, which are based on this approach. Most of the development was
made through close cooperation with a commercial software developer (ASC, CFX and now ANSYS) so that
many of the developments ultimately became available for other turbine manufacturers to use.

5.2 Viscosity and turbulence

In contrast to the Euler equations, the Navier–Stokes equations take into account both viscous and turbulent
effects and can therefore be used for
• Loss analysis.
• Analysis of boundary layer effects such as boundary layer separation in all turbine components.
• Component analysis for decelerated flow.
While the flow in a water turbine accelerates from runner inlet to runner outlet, it decelerates in the draft
tube in order to recover static pressure from the available kinetic energy. In a pump turbine deceleration of
the flow, and the associated pressure rise, are characteristic for the flow in pump direction through almost all
218 H. Keck, M. Sick

Stator Runner Stator Runner Stator Runner

Stage interface Frozen rotor interface Unsteady rotor-stator interface


Fig. 10 Three major types of rotor–stator interfaces

components. Consequently, the analysis of draft tube flow (see [8]) and of the flow in pumps (see [39]) and
pump turbines in the pump mode became far more realistic with the possibility of solving the RANS equations.
The long calculation times with fine grids that resolved the boundary layers of the CFD RANS codes meant
that during the nineties pump turbine runner design was done by a combination of both Euler and RANS
solvers as described by Sebestyen et al. in [42]. But for more sophisticated studies where viscous effects are
totally dominant such as in diffusers, see [30], and the onset of flow instabilities in pump turbines, it was clear
that RANS solver had to be used, see [10]. By 2000 it had become state-of-the-art in major refurbishment
projects to analyse and optimise components of a pump turbine entirely by means of RANS simulations. Two
such projects, Solina and Zarnowiec in Poland, are described in [38].

5.3 Coupled flow simulation of rotating and stationary components

During the mid-eighties interface models between rotating and stationary domain were developed and imple-
mented into academic CFD codes. Such methods allow the steady interaction between components to be
modelled and thus reduce the uncertainty with respect to the inlet conditions at the turbine runner inlet. The
models opened the era of CFD simulation of entire turbines, including the interaction of all turbine components.
Interface models have since then been developed for use both in Euler codes for the preliminary design and in
Navier–Stokes applications for detailed analysis and optimization.
Various mathematical models were developed in order to take into account the interaction between station-
ary and rotating components, starting from relatively simple circumferential mixing planes, over frozen rotor
interfaces, to more sophisticated stage interface models which allow local non-uniformities of the flow field to
be extended to the stage interface, see Fig. 10. All these models have in common the fact that they reduce the
time dependent interaction between rotating and stationary components to a steady-state problem. But they
differ in the way how disturbances like the wake of stator blades are transferred. Most realistic is the unsteady
rotor–stator interface in which disturbances are properly transported across the interface (Fig. 10, right hand
side) whereas in a stage interface these disturbances are mixed out circumferentially (Fig. 10, left hand side). A
frozen rotor algorithm enables disturbances to be transported from one frame of reference into the other—but
in the wrong way which can lead to an unrealistic overestimation of the effect of this disturbance as visualized
in Fig. 10, middle, showing the wake of a stator blade hitting a runner blade along the full length.
Validation of CFD has always been an important issue and particularly so with the interface algorithms.
A comparison between velocity measurements and CFD predictions before and after the runner when a stage
interface between guide vanes and runner is applied and, as contrast, when a velocity profile is prescribed at
runner inlet demonstrates the high quality of CFD prediction when a stage interface is used. At the same time,
it identifies the huge error a poor runner inlet condition can cause at the runner outlet [43].
With the ability of modelling the flow through rotating and stationary components in one single CFD sim-
ulation the concept of the virtual test rig was born. The first “purely CFD” hill chart of a Francis turbine was
published in 1996 [24,44] showing both the enormous potential of CFD prediction and also some limitations.
The CFD prediction reflected very well the shape of the hill chart but also showed between 4 and 8% deviation
from the measured absolute efficiency values depending on the operating point, see Fig. 11.
Thirty years of numerical flow simulation in hydraulic turbomachines 219

Fig. 11 Entire Francis turbine, numerical hill chart versus experimental hill chart

These developments provided the basis for the shift of CFD applications from a mere design and optimi-
sation tool to a highly complex analysis tool which today is used through all stages of a turbine refurbishment,
from the feasibility study to design and optimization [48].

5.4 Colours for directors?

The increasing significance of RANS methods, both in turbine design and in component analysis, caused the
accuracy and reliability of CFD simulations to become a major issue. It became clear that it was a notoriously
difficult technology to apply in practice. This gave rise to dicta for CFD such as “Cheats, Frauds and Deceivers”
or “Colours For Directors”. Among CFD users in the industrial environment considerable efforts were made
to improve reliability and accuracy of CFD, and to understand the causes of errors and uncertainties, see [7].
In water turbines the draft tube is the one component for which numerical flow simulation is most difficult
and least reliable. Validation of the computational method and parameters such as the computational domain
and boundary conditions, the computational grid, discretization scheme and turbulence model has become
a significant task. The European Research Community on Flow, Turbulence and Combustion (ERCOFTAC)
organized a continuing series of workshops on validation of CFD for basic test cases such as backward facing
step, curved pipes, horseshoe vortex or flow over an obstacle. Workshops were also organized to examine the
accuracy of draft tube flow prediction (such as Turbine 99 organized by IAHR and ERCOFTAC, [1]). The
outcome of that workshop is shown in Fig. 12 demonstrating the relative randomness of the CFD prediction
of pressure recovery in a turbine draft tube of the different participants at that time. A second workshop was
held of which results are reported by Andersson et al. in [2], and this identified many of the issues related to
boundary conditions which allowed more consistent predictions than those of the first workshop.
Following this, draft tube flow analysis was the focus of the international research project FLINDT which
was carried out at the Swiss Federal Technology Institute Lausanne (EPFL), see [4]. Continuous effort since
then is illustrated in Fig. 13 which shows the model of the FLINDT Francis turbine on the left hand side and the
measurement data of the pressure recovery (red line) together with CFD predictions of all different FLINDT
partners (other colours). Between the upper graph and the graph below lie 5 years of validation work in order
to improve CFD methods for draft tube flow modelling. With the CFD approach producing the light blue line
the prediction of the pressure recovery finally seems reliable (compare dark blue line, data, with light blue line,
CFD prediction without any data input from measurement). It should be mentioned that this reliability is not
220 H. Keck, M. Sick

1.6
Experimental value: 1.12
1.2

0.8

0.4

Iowa K-w R

LMH-EPFL N3S k-e R


Iowa k-w T

Iowa k-e R

ckd RSM R
LMH-EPFL N3S k-e T

TEV RSM T

HOF k-e R

UF-GE k-e R
NUT k-e T

NUT k-e T R

KV1 k-e R
KV2 k-e R
HOT k-e R
Iowa k-e T

ckd RSM T

LMH-EPFL Fluent a k-e R


LMH-EPFL Fluent b k-e R
SWECO k-e T

fluent RSM T
HQT k-e T
HQF k-e T

UF-GE k-e T

LTU RSM T

HTALuzern k-e R
KV2 k-e T

KIT Japan ak-e T

AEAT RSMssg T
KV1 k-e- T

AEAT RSMlrr T

LTH LES T
CD cubic k-e T
LTH S-A T

LMH-EPFL Fluent a k-e T


LMH-EPFL Fluent b k-e T
fluent k-e T

HTALuzern k-e T
AEAT k-e T

KIT-Japan eke-T
KIT-Japan k-e T

KIT-Japan rk-e T
Fig. 12 Turbine 99 workshop, courtesy Vattenfall Utveckling

0,84

0,74

0,64

0,54

0,44

0,34
0, 33 0,35 0, 37 0, 39 ϕ 0,41

0.9

0.8

0.7
RC

0.6

0.5

experimental
Messung data
0.4 prof ile f rom meas urement
calculation c oar se
calcualtion f ine
0.3
0.33 0.35 0.37 0.39 0.41

Phi

Fig. 13 CAD model of FLINDT turbine and diagrams showing predicted versus measured pressure recovery over nondimensional
throughflow

reached by manipulation of obscure fudge factors but by clean and precise modelling only. Factors such as the
nature of the turbine leakage flow and fine details of the impeller outlet flow distribution need to be correctly
modelled. Nevertheless, it should be mentioned that CFD prediction of draft tube flow is still an open issue
for bulb turbines which react extremely sensitively to draft tube flow quality, see for example [11].

5.5 Pelton CFD

Component analysis by RANS in Francis and Kaplan turbines also opened the door for CFD based Pelton
turbine analysis. Pelton distributors which until the middle of the nineties had been developed and optimized
based on empirical methods and experiments only could now be thoroughly analyzed and optimized with
Thirty years of numerical flow simulation in hydraulic turbomachines 221

Fig. 14 CFD analysis of a Pelton distributor flow, from [47] (1)

the help of CFD, see [32] and [45]. Figure 14 shows the computational grid of a five nozzle Pelton turbine
distributor (left) and the predicted velocity distribution in the plane of symmetry (right) revealing the typical
zone of deceleration and backflow at all bifurcations. This recirculation zone near the bifurcation is not only a
cause of additional head loss in the distributor. Far more significant are the resulting disturbances to the flow
in the nozzles and subsequent deterioration of the jet at nozzle exit. Measurements show that poor jet quality
can lead to more than 0.5% loss in efficiency because the jet bucket interaction is no longer optimal.
While CFD prediction of Pelton distributor flow is well validated and delivers reliable results the inter-
pretation of the flow features within the distributor with respect to their impact on jet quality and resulting
jet–bucket interaction is not so clear. Turbulence intensity and strength of secondary flow or vortex structures
respectively are analyzed in order to conclude on jet quality. Thus, the CFD simulation of the jet flow itself and
of the jet–bucket interaction was the next obvious step, leading CFD engineers onto a new field of modelling
and application: time dependent multiphase free-surface flow simulation.

6 More than just water (2000–today)

6.1 Multiphase, unsteadiness and multiphysics

With increasing computational power the questions to be answered by CFD are still growing at least as fast
as processor speed and disk space. More and more complex flow phenomena in water turbines are being
investigated, requiring the development and validation of more sophisticated models for turbulence, cavitating
and multiphase flow as well as fluid structure coupling. In the following sections the most important new
developments are presented and discussed.

6.2 Pelton jet simulation

Flow observations at the test rig as shown in Fig. 15, middle and right hand side, demonstrate the complexity
of free jet flow in a Pelton turbine. The surface of the water jet is characterized by a mixing zone between
water and air. Furthermore, the jet shape can deteriorate or become off-axis. Both phenomena are now known
to be caused by the characteristics of the Pelton distributor flow, [54]. From 1998 first results of Pelton jet flow
simulation were published by [3,31], and [46, p. 2].
Excellent validation work has been done and published by [33]. In the left hand picture of Fig. 16 CFD
prediction of the axial velocity in a Pelton jet is shown to be in good agreement with experimental data (red
crosses). In order to carry out LDV measurement of the velocity field within the Pelton jet a glass plate had
to be positioned directly on the jet surface thus enabling the Laser beam to penetrate the water surface. Both
measurement and CFD prediction reveal the wake of the needle in the centre of the nozzle. Two-phase model-
ling has been found to be extremely difficult in the region of detachment at nozzle outlet when it is not forced
by a specific geometry feature such as a sharp edge.
222 H. Keck, M. Sick

Fig. 15 Pelton jet

Fig. 16 Validation of CFD predicted velocity field in the Pelton jet

A new measurement technique with a needle touching the water surface had to be developed in order to
measure the jet shape which is displayed in the middle of Fig. 16 revealing a deformation which can also be
seen in the picture presented in Fig. 15, right. This deformation is known to not only cause additional loss but,
what is more severe, to cause cavitation damage in Pelton buckets. The origin of the deformation could be
understood with the help of a CFD simulation, see Fig. 16, right: the flow field in Pelton nozzles downstream
of a curved duct is characterized by a secondary flow which leads to a deformation of the jet at the stagnation
point of the secondary flow motion. Again, CFD prediction compared to measurement shows very good results.

6.3 Jet-bucket interaction

In a Pelton runner energy conversion takes place in the Pelton bucket where the kinetic energy of the free
jet is converted into momentum driving the runner. Clearly, non-viscous forces dominate this flow, a fact
leading to a fairly precise CFD prediction of the pressure distribution in a Pelton runner as demonstrated
in [27] and [34], see Fig. 17, left hand side. At the outer side of the Pelton bucket flow mechanisms are
dominated by viscous effects, surface tension and cavitation. Therefore the CFD prediction is not as pre-
cise as for highly loaded positions on the inner side of the bucket, see [35] and Fig. 17, right hand side.
If CFD is to be used as optimization tool for Pelton bucket design these details have to be modeled with
high accuracy. Therefore, more investigation, model refinement and validation will be done during the next
years.
Even with these weaknesses a CFD prediction of torque and efficiency has already become possible with
remarkable accuracy as shown in Fig. 18 on the left hand side. Transferring the CFD predicted pressure distri-
Thirty years of numerical flow simulation in hydraulic turbomachines 223

Fig. 17 Measurement and CFD prediction of static pressure over time in a model Pelton turbine runner bucket

Fig. 18 CFD prediction and measurement of efficiency curves (left), CFD prediction serves as input to FEM analysis (right)

bution as input to a FEM analysis of displacement and stress enables a complete analysis and optimization of
both hydraulic design and life time of a Pelton runner, see Fig. 18 on the right hand side [35], [36].

6.4 Cavitation

When the static pressure drops below vapour pressure, vapour bubbles develop which collapse as soon as the
pressure rises again. The high pressure caused by the bubble implosion causes surface erosion often damaging
the turbine surfaces. Cavitation occurs mostly in zones of low overall pressure at the runner outlet but, in case
of high incidence, at the blade inlet. It can also occur in zones of strong streamline curvature such as the bend
of the casing between guide vanes and runner, and at the tip clearance vortex. Design engineers generally
identify cavitation risk by evaluating zones of pressure below vapour pressure in computed flow fields without
a two phase model. With this approach the effect of a cavitation bubble on the flow field is neglected which in
most cases is an adequate approach.
But if more information is needed, such as the effect of cavitation on the efficiency or a more precise
prediction of the extent of a cavitation bubble, a two-phase flow simulation has to be carried out. Today, the
224 H. Keck, M. Sick

Fig. 19 CFD simulation of cavitating flow in a bulb turbine runner; two phase flow—Rayleigh Plesset model; [41]

Rayleigh–Plesset approach is being used for modelling the formation and decay of vapour bubbles, thereby
enabling a more accurate prediction of the cavitation zone and the associated drop in efficiency. Validation of
the CFD approach is shown for a bulb turbine runner in Fig. 19. The predicted size and position of the vapour
zone is found to be in very good agreement with the observation at the test rig, whereas the related rise and
sharp drop in efficiency with decreasing Thoma number is predicted with good, but not excellent, accuracy.
Here too, more work is needed.

6.5 Unsteady flow phenomena

Modelling unsteady flow phenomena which cause dynamic loads on the structural components is one major
step forward in recent CFD applications to water turbine flow. Figure 20 shows the three most important causes
for dynamic loads in Francis turbines and pump turbines: draft tube vortex at part load, von Kármán vortex
shedding at trailing edges of stay vanes, guide vanes or runner blades and the interaction between guide vanes
and runner blades. These phenomena and their effect on the life time of water turbines has become a major
issue during the past years mainly because water turbines are being optimized with respect to weight, which
requires more knowledge about safety margins. Safety margins can be defined more precisely on the basis
of more accurate calculation methods of both the load on and the response of the structure. An overview of
today’s status of dynamic load prediction is given by Sick et al. in [50].

6.6 Draft tube vortex at part load

In a certain range of part load operation the flow at the draft tube inlet is characterized by a strong swirl which
causes a rotating vortex as shown in Fig. 21 with low pressure in the vortex core and thereby a rotating pressure
field. In some hydro power stations additional measures such as aeration through the hub or fins in the draft
tube cone have to be taken in order to prevent damage due to these strong pressure pulsations. In order to learn
Thirty years of numerical flow simulation in hydraulic turbomachines 225

Fig. 20 Unsteady flow phenomena causing dynamic load: draft tube vortex (left), von Kármán vortex shedding (middle), inter-
action between stationary and rotating components (right)

Fig. 21 Cavitating draft tube vortex as observed at the test rig (left) and predicted by CFD (right), [51]

more about the nature of this flow phenomenon CFD modelling has been carried out and validated throughout
recent years. The simulations identify the rolling up of the vortices at the shear layer between the outer flow
and the inner region with back flow leading to a periodically rotating helical vortex as can be seen in Fig. 21.
Most important was the finding that the CFD simulation of the draft tube vortex requires great care with respect
to turbulence modelling, see [47]. Generally, two-equation models fail because of the assumption of isotropy
which is not applicable for such strongly curved flow paths. Instead, turbulence modelling has to be done with
the Reynolds stress models, Large Eddy Simulation or similar approaches.
Due to the strong curvature of the flow path, the static pressure in the vortex core drops below vapour
pressure causing a cavitating vortex rope which can then be observed at the test rig. Draft tube vortex simula-
tion with a two-phase approach using the Rayleigh Plesset model is a serious challenge with respect to spatial
resolution and accuracy because of the steep pressure gradient to be resolved as can be seen in Fig. 22 which
shows the pressure distribution across the vortex core as predicted with several levels of spatial discretization
and with different turbulence models. In order to predict cavitation as observed at the test rig, the pressure has
to drop below vapour pressure (dashed line), which is achieved with high quality spatial resolution only. In
his PhD thesis [52] Stein showed that an accurate simulation of the draft tube vortex would require about 500
million grid nodes in order to achieve the accuracy needed for a realistic prediction of the volume of vapour in
the vortex core. The latter is important for an accurate prediction of both the rotation frequency of the vortex
and the dynamic response of the entire system to the excitation by the draft tube vortex. But such a size of
computation is not yet possible in the industry.

6.7 Von Kármán vortex shedding

Finer grid resolution together with improved discretization methods (schemes of second order accuracy or
TVD schemes) may reveal small scale unsteadiness which could not be found beforehand. This sometimes
happens without being wished by the CFD engineer who often is interested in steady state average flow infor-
mation. But in case of fatigue failure of stay vanes the detailed flow analysis around the stay vanes revealed
von Kármán vortex shedding, thereby contributing to a better understanding of the cause of the failure and
problem solution, see [29]. Again, in the case of von Kármán vortex shedding the turbulence model has been
226 H. Keck, M. Sick

Fx Fy
800 600

600
400
400
200
200
Measurement f Measurement
0 0
Calculation Calculation
0.0 2.0 4.0 6.0 8.0 10.0 0.0 2.0 4.0 6.0 8.0 10.0
-200
-200
-400
-400
-600
Number of revolutions Number of revolutions
-800 -600

Mx My
80 100

60 80

40 60

20 40

0 Measurement 20 Measurement
0.0 2.0 4.0 6.0 8.0 10.0
-20 Calculation 0 Calculation
0.0 2.0 4.0 6.0 8.0 10.0
-40 -20

-60 -40

-80 -60
Number of revolutions Number of revolutions
-100 -80

50000

45000

40000
Pressure [Pa]

35000
plane of
30000
observation
SST_310k nodes_75% 2nd Order
25000

Reynolds_Stress_1mill nodes_75% 2nd Order


20000

Reynolds_Stress_1mill nodes_2nd Order


15000

Reynolds_Stress_1mill nodes_2nd
10000
Order+cavitation
vapour pressure
5000

-0.3 -0.2 -0.1 0 0.1 0.2


Radius [m]

Fig. 22 Validation CFD prediction: lateral forces and bending moments on shaft due to draft tube vortex (top) pressure drop in
vortex core (bottom): influence of discretization method and turbulence model, [49]

Fig. 23 CFD prediction of vorticity: effect of turbulence model; standard k-ε model (left), SAS model (right)

identified as being an important numerical parameter with respect to the accuracy of the solution, see Fig. 23
which demonstrates the unrealistic high diffusion in case of the k-ε model leading to a smoothed shape of the
vortex street, [25]. Furthermore, a detailed study of the effect of the transition laminar—turbulent boundary
layer on the Strouhal number which is described by [40], could be identified in a CFD study by [53].
Thirty years of numerical flow simulation in hydraulic turbomachines 227

Fig. 24 CFD simulation of the pressure field in a pump turbine (left), comparison CFD prediction—measurement (right) [26]

6.8 Rotor–stator interaction: fluid mechanics meets structural mechanics

Between rotating and stationary components unsteady effects are inevitably caused by the upstream potential
effect of the leading edges (stagnation pressure) and the downstream effect of the wake disturbances. While
the excitation frequencies due to the interaction of guide vanes and runner blades as shown in Fig. 24 can be
analytically determined [9], the amplitudes are determined by the flow field and need to be either measured
or predicted by means of CFD. Validation of the pressure pulsation due to rotor stator interaction has been
done for the vaneless space between guide vanes and runner of model pump turbine, see [25]. The agreement
between CFD prediction and measurement data has been found to be very good, see Fig. 24.
For turbine design the major step is now to analyze the response of the structure to the dynamic load due
to the flow field. CFD analysis allows the harmonic response method to be applied. The pressure amplitudes
are extracted from the CFD predicted flow field for all frequencies of interest, i.e. natural frequencies of the
structure which have been identified of being critical for life time. Structural analysis by means of finite element
analysis (FEA) is then carried out in order to determine the response of the structure, which means the stress
due to dynamic load for each individual frequency. This approach, the harmonic response method, is now more
and more being validated and applied with considerable success throughout the water turbine industry, see
[50]. The coupling of fluid mechanics with structural mechanics becomes more and more close and efforts are
already under way to simulate fluid–structure interaction. For water turbine applications questions of dynamic
load and fluid–structure interaction are investigated in the international research project HydroDyna, see [5].

7 Conclusion and outlook

The application of computational fluid dynamics (CFD) in the field of water turbines and pumps started about
30 years ago. This paper has reviewed the main steps and breakthroughs in the methods that were made during
this period, with practical examples to illustrate the main issues and level of complexity during each period.
Over this period the complexity of the calculations has increased and the level of difficulty of the problems
that can be tackled has become higher. CFD has become a powerful tool which requires validation versus
smartly designed and executed experiments as well as profound knowledge in fluid mechanics in order to
identify numerical problems on one side and to most effectively interpret proper CFD predictions on the other
hand.
228 H. Keck, M. Sick

Extrapolation from the experience over the last 30 years would suggest that large strides are still possi-
ble, and that careful systematic analysis and validation of the numerical methods will be the key to the next
breakthroughs.
Similar to detailed experimental investigations highly sophisticated and expensive CFD studies will be
increasingly used to understand complex flow mechanisms and multi physical problems. The knowledge
gained from complex numerical studies will lead to improved design rules and to better understanding and use
of standard design tools. A two-fold development is foreseen: fast, well calibrated CFD methods for design
which are automatized to a high degree on one hand and, on the other hand, highly complex multi physics
studies for special applications and investigations in order to improve design procedures.

Acknowledgments The authors would like to thank the large number of present and former colleagues at Andritz VA TECH
HYDRO (formerly Escher Wyss/Sulzer Hydro) for their contributions during the last 30 years to CFD in hydraulic turbomachines.
Special thanks are expressed to the authors’ teachers in Fluid Mechanics, Prof. Dr. W. Schneider, Vienna, Prof. Dr. K.O. Felsch
and Prof. Dr. J. Zierep, Karlsruhe.

References

1. Andersson, U., Dahlbäck, N.: Experimental evaluation of draft tube flow—a test case for CFD simulations. In: XIX IAHR
Symposium on Hydraulic Machinery and Systems, Singapore (1998)
2. Andersson, U., Engström, F., Gustavsson, H., Karlsson, R.: The Turbine-99 workshops—lessons learned. The QNET-CFD
Network Newsletter, vol. 2(3) (2003)
3. Avellan, F., Dupont, Ph., Kvicinsky, S.: Flow calculations in Pelton turbines—Part 2: free surface flows. In: XIXth IAHR
Symposium on Hydraulic Machinery and Systems, Singapore (1998)
4. Avellan, F.: Flow investigation in a Francis draft tube: the flindt project. In: Proceedings of the Hydraulic Machinery and
Systems 20th IAHR Symposium, Charlotte (2000)
5. Avellan, F.: http://hpc.epfl.ch/bgl/scientific-projects/hydrodyna/hydrodyna-project
6. Casey, M.V., Keck, H.: Hydraulic turbines. Handbook of Fluid Dynamics and Fluid Machinery, Chap. 27.9, vol. 3. Wiley,
London (1996)
7. Casey, M.V., Wintergerste, T.: Best practice guidelines for industrial CFD. Published by ERCOFTAC (2000)
8. Drtina, P., Goede, E., Schachenmann, A.: Three dimensional turbulent flow simulation for two different hydraulic turbine
draft tubes. In: Proceedings of the First European Computational Fluid Dynamics Conference. Brussels, Belgium, 7–11 Sep.
1992
9. Dubas, M.: Über die Erregung infolge der Periodizität von Turbomaschinen. Ingenieur Archiv. Springer, Verlag (1984)
10. Eisele, K., Muggli, F., Zhang, Z., Casey, M., Sallaberger, M., Sebestyen, A.: Experimental and numerical studies of flow
instabilities in pump-turbine stages. In: XIX IAHR Symposium, Singapore (1998)
11. Gehrer, A., Egger, A., Riener, J.: Numerical and experimental investigation of the daft tube flow downstream of a bulb
turbine. In: Proceedings of the XXI IAHR Symposium on Hydraulic Machinery and Systems, Lausanne (2002)
12. Göde, E., Ryhming, I.L.: 3D-Computation of the flow in a Francis runner. Sulzer Technical Review No. 4 (1987)
13. Göde, E., Cuénod, R., Pestalozzi, J.: Visualization of flow phenomena in a hydraulic turbine based on 3D flow computations.
Waterpower (1989)
14. Göde, E., Cuénod, R., Bachmann, P.: Theoretical and experimental investigation of the flow field around a Francis runner.
IAHR Symposium, Trondheim (1988)
15. Göde, E.: A Stacking technique for multistage 3D flow computation in hydraulic turbomachinery. GAMM workshop, Lau-
sanne (1989)
16. Höller, H.K., Keck, H.: The complete finite element flow computation in a bulb turbine from inlet trash rack to runner outlet.
IAHR Symposium Tokyo (1980)
17. Jermann, P., Keck, H., Pestalozzi, J.: FEM—flow analysis in hydraulic turbomachinery runners. Water Power Conference,
Las Vegas (1985)
18. Keck, H.: Die Anwendung der Finite-Elemente-Methode auf die Berechnung von 3D-Potentialströmungen. Internal Escher
Wyss Report WT-78–566 (1978)
19. Keck, H., Haas, W.: Vorticity and blade circulation modeling in the calculation of quasi-three-dimensional cascade flows
with finite elements. 3. In: Int. Conf. on Finite Elements in Flow Problems, Banff/Canada (1980)
20. Keck, H.: Finite Element Analysis of Flows in Hydraulic Turbomachines. Escher Wyss News, vol. 53 (1980)
21. Keck, H., Haas, W.: Finite Element Analysis of quasi-3D flow in turbomachines. Finite Elements in Fluids, Chap. 24, vol.
4. Wiley, London (1982)
22. Keck, H., Göde, E., Pestalozzi, J.: Experience with 3D Euler flow analysis as a practical design tool. IAHR Symposium,
Belgrade (1992)
23. Keck, H., Göde, E., Grunder, R., Pestalozzi, J.: Upgrading by new runners based on 3D-flow simulation and model testing.
IAHR Symposium, Sao Paulo (1992)
24. Keck, H., Drtina, P., Sick, M.: Numerical hill chart prediction by means of CFD stage simulation for a complete francis
turbine. In: Proceedings of the XVIII IAHR Symposium, Valencia (1996)
25. Keller, M.: CFD of Unsteady Phenomena in water turbines. In: 24th CADFEM User’s Meeting (2006)
26. Keller, M., Sallaberger, M.: (2) Modern hydraulic design of pump turbines. In: Proceedings of the 14th International Seminar
on Hydropower Plants, Vienna (2006)
Thirty years of numerical flow simulation in hydraulic turbomachines 229

27. Kvicinski, S., Kueny, J.L., Avellan, F., Parkinson, E.: Experimental and numerical analysis of free suface flows in a rotating
bucket. In: XXI IAHR Symposium, Lausanne, Switzerland (2002)
28. Launder, B.E., Spalding, D.B.: Mathematical Models of Turbulence, Academic Press, London/New York (1972)
29. Lockey, K., Keller, M., Sick, M., Gehrer, A.: Flow induced vibrations at stay vanes: Experience at site and CFD simulation
of von Kármán vortex shedding. Hydro2006, Porto Carras, Greece (2006)
30. Muggli, F.A., Eisele, K., Casey, M.V. et al.: Flow analysis in a pump diffuser—Part 2: Validation and limitations of CFD for
diffuser flows. ASME J. Fluids Eng. 119, 978–984 (1997)
31. Muggli, F., Zhang, Z. Schärer, C., Geppert, L.: Numerical and experimental analysis of Pelton turbine flow, Part 2: the free
surface jet flow. XX IAHR Symposium, Charlotte (2000)
32. Parkinson, E., Lestriez, R., Chapuis, L.: Flow calculations in Pelton turbines. Part 1: Repartitor and injector numerical
analysis. In: Proceedings of the XIX IAHR Symposium, Singapore (1998)
33. Parkinson, E., Garcin, H., Vullioud, G., Muggli, F., Zhang, Z., Casartelli, E.: (1) Experimental and numerical investigations
of the free jet flow at a model nozzle of a Pelton turbine. In: Proceedings of the XXI IAHR Symposium on Hydraulic
Machinery and Systems, Lausanne, Switzerland (2002)
34. Parkinson, E., Vullioud, G., Geppert, L., Keck, H.: (2) Analysis of Pelton turbine flow patterns for improved runner component
interaction. Hydropower Dams (5) (2002)
35. Parkinson, E., Neury, C., Garcin, H., Vullioud, G., Weiss, Th.: Unsteady analysis of a Pelton runner with flow and mechanical
simulatons. Hydro (2005)
36. Parkinson, E., Angehrn, R., Weiss, Th.: Modern design engineering applied to Pelton runners. Hydropower Dams (4) (2007)
37. Roache, P.J.: Computational Fluid Dynamics, Hermosa Publishers, Albuquerque, New Mexico (1972)
38. Sallaberger, M., Staehle, M., Thoma, W., Kiedrowski, T., Krasicki, R., Lewandowski, S.: (2) Major progress in upgrading
of reversible pump turbines. In: Proceedings of Hydro, Bern, Switzerland (2000)
39. Schachenmann, A., Muggli, F., Gühlich, J.: Comparison of three Navier–Stokes codes with lda measurements on an industrial
radial pump impeller. In: ASME Fluids Engineering Conference, Washington DC (1993)
40. Schlichting, H.: Grenzschicht Theorie. Braun Verlag, Karlsruhe, 8. Auflage, Kapitel II, p. 32 (1984)
41. Schmidl, R.: Contribution to the cavitational optimization of Kaplan blades by numerical flow simulation and evolutionary
strategies. Doctoral Thesis, Faculty of Mechanical Engineering, Graz University of Technology, Graz, Austria (2007)
42. Sebestyen, A., Jaquet, M., Keck, H.: CFD-Design procedure for runner replacement of reversible pump-turbines.
In: Proceedings of the XIX IAHR Symposium, Singapore (1998)
43. Sick, M., Casey, M., Galpin, P.: Validation of a stage calculation in a Francis turbine. In: Proceedings of the XVIII IAHR
Symposium, Valencia, Spain (1996)
44. Sick, M., Drtina, P., Casey, M.V.: The use of stage capability in CFD for turbomachinery, with application in a Francis
turbine. Int. J. Comput. Appl. 11(3/4/5), 219–220 (1998)
45. Sick, M., Schindler, M., Drtina, P., Schärer, C., Keck, H.: (1) Numerical and experimental analysis of Pelton turbine flow.
Part 1: Distributor and Injector. XX IAHR Symposium, Charlotte (2000)
46. Sick, M., Keck, H., Parkinson, E., Vullioud, G.: (2) New challenges in Pelton research. In: Proceedings of Hydro, Bern,
Switzerland (2000)
47. Sick, M., Dörfler, P., Lohmberg, A., Casey, M.: Numerical Simulations of Vortical Flows in Draft Tubes. WCCMV, Vienna,
Austria (2002)
48. Sick, M.: State of the art of CFD based feasibility studies for water turbines. Technology Review for HPLIG (CEATI), CEATI
Report No. T042700-0324, Montreal (2004)
49. Sick, M., Stein, P., Dörfler, P., Sallaberger, M.: Part-load instabilities in francis turbines and pump turbines. Int. J. Hydropower
Dams (2005)
50. Sick, M., Lais, S., Weiss, Th.: Numerical prediction of flow induced dynamic load in water turbines: recent developments
and results. In: Proceedings of Hydro2007, Granada, Spain (2007)
51. Stein, P., Sick, M., Doerfler, P., White, P., Braune, A.: Numerical simulation of the cavitating draft tube vortex in a Francis
turbine. In: Proceedings of the XXIII IAHR Symposium, Yokohama (2006)
52. Stein, P.: Numerical simulation and investigation of draft tube vortex flow. Ph.D. Thesis, Coventry University, Coventry
(2007)
53. Vu, T.C., Nennemann, B., Ausoni, Ph., Farhat, M., Avellan, F.: Unsteady CFD prediction of von Kármán vortex shedding in
hydraulic turbine stay vanes. In: Proceedings of Hydro2007, Granada (2007)
54. Zhang, Zh., Casey, M. V.: Experimental studies of the jet of a Pelton turbine, vol. 221. In: Proc ImechE. Part A J. Power
Energy, pp. 1181–1192 (2007)

Potrebbero piacerti anche