Download as pdf or txt
Download as pdf or txt
You are on page 1of 97

2008:016 CIV

MA S T E R’S T H E SI S

Dynamic stress assessment


in high head Francis runners

Yasir Dawood Siwani

MASTER OF SCIENCE PROGRAMME


Engineering Physics

Luleå University of Technology


Department of Applied Physics and Mechanical Engineering
Division of Fluid mechanics

2008:016 CIV • ISSN: 1402 - 1617 • ISRN: LTU - EX - - 08/016 - - SE


Dynamic stress assessment in
high head Francis runners

Yasir Dawood Siwani

Master thesis
Department of Applied Physics and Mechanical Engineering
Luleå University of Technology
SE-971 87 Luleå
Sweden

January 25, 2008


Preface

This work was conducted under the supervision of Associate Professor


Michel Cervantes, at the department of fluid mechanics at luleå univer-
sity of technology, Sweden and Trond Moltubakk technical leader for
HHT mechanical Norway at Rainpower Norway AS.

The work was preformed during 2007/08-12 at HHT mechanical Noway,


Rainpower Norway AS in Kjeller, Norway.

There are some people I would like to acknowledge for their help and
support during the thesis period. First I would like to thank Trond
moltubakk and Michel Cervantes for the time given me both within
and outside the thesis, their guidance, patient, criticism and suggestions
throughout the work.

I would also like to thank Ståle Risberg, hydraulic design engineer at


Rainpower Norway AS, for his help and guidance with the CFD calcula-
tions preformed and for the many discussions not only on the hydraulic
part of the thesis but also on the mechanical calculations. Ståle Risberg
was responsible for obtaining the CFD results to be used for further

i
ii

analysis in the thesis.

Special thank to Øystein Gjerde, Mechanical design engineer at Rain-


power Norway AS, for his patient and discussions within the mechanical
work in the thesis.

Yasir Dawood Siwani


January 25, 2008
Abstract

The deregulation of the energy market in the recent years has lead to an
increased willingness to operate hydro turbines over a larger operational
range and with increased number of start and stops. This means that
the turbine will be subject to a more frequent and larger load variation.
Especially for high head turbines, the impact of dynamic forces has been
of considerable interest. High head turbine runners are subjected to dy-
namic forces originating from various sources. One significant source is
the interaction between the runner and guide vanes, also called rotor -
stator interaction (RSI). RSI creates periodic pressure fluctuations in the
runner. Combined with stress concentration and material defects, the in-
duced pressure fluctuations can give an increased risk of fatigue cracks.
For Francis runners, the areas near the trailing edge of the runner blade
towards the band and/or crown has been identified as critical areas for
fatigue, and for high head runners there has been several examples where
fatigue failure have been initiated near the trailing edge toward the band.

The thesis presents a validation of dynamic stress obtained via computa-


tional fluid dynamics and the finite element method. A new runner was
installed at Tonstad hydropower plant in Norway. Site measurements

iii
iv

were carried out on the new replacement runner and one of the original
design runners. The tests were conducted by Halvard Bjørndal in 2003
[4]. The measurements were partially done to compare the mechanical
properties of the new design to the original one, during operation [4],
and also for validation of FEA stress prediction.

Measurement results on the new runner were used for validation of the
finite element method. A computational fluid dynamic geometry model
of the new runner was modeled. Geometrical symmetry allowed for a
third of the real runner including guide vanes and an extension of the
outlet to be modeled.

Unsteady CFD simulations providing the load for FEA calculations were
preformed by Ståle Risberg a hydraulic design engineer at Rainpower
Norway AS. The transient CFD solution showed as predicted a fluctuat-
ing torque on the blades due to the runner blade - guide vane interaction
(or Rotor Stator Interaction (RSI)). The calculations were carried out
on 81% load, which gives a mean torque of about 3.3 M N m which is an
output of 129.6 M P a.

The CFD analysis resulted in a mean torque of 3.2 M N m and at the


guide vane frequency an amplitude of 76.3 kN m. The result yielded a
mean torque deviation of 1.8% between the CFD torque and the mean
torque calculated based on a nominal output value of 160 M P a.

The dynamic pressure for one cycle, one blade passing a hole guide vane
passage, is the dynamic pressure load used in the FEA. The FEA model
consisted of a portion of the band and crown, a runner blade and a
splitter blade. The mean torque obtained by the FEA was found to be
3.21 M P a. That amounted to a deviation of only 2.71% between the
FEA torque and the mean torque calculated for 81% load.

The maximum principle quasi-static mean stress calculated was found


to be located at the trailing edge of the blade near the band. The high-
est mean stress, at the predefined validation points, was located 15 mm
from the trailing edge of the blade on the pressure side. Compared to the
measured value, the calculated stress was 32.7% higher. The deviation
was 33.8%, 9% and 1% for 25 mm, 200 mm and 300 mm respectively on
the pressure side. Looking at the suction side, the mean stress calculated
v

deviated by 121% at 15 mm and 7.5% 25 mm from the trailing edge.

The stress amplitudes at the guide vane passing frequency (150 Hz)
were much lower than the measured values, for four out of the validation
points. On the pressure side, the deviation was 67.9%, 63.2%, 32.1%
and 13.3% at 15 mm, 25 mm, 200 mm and 300 mm respectively. The
amplitudes on the suction side deviated by 94.3% 15 mm and 71.1%
25 mm from the trailing edge of the blade.

Although the deviation was higher for the amplitudes than for the mean
stress, the tendencies were similar. The deviation seamed to decrease for
distance further from the trailing edge of the blade. It was concluded
that one cause could be, thickness difference between the design model
and the real one.
Contents

Preface i

Abstract iii

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Hydraulic machines . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Kaplan turbines . . . . . . . . . . . . . . . . . . . 4
1.2.2 Pelton turbines . . . . . . . . . . . . . . . . . . . . 6
1.2.3 Francis turbines . . . . . . . . . . . . . . . . . . . . 8
1.3 Site measurements . . . . . . . . . . . . . . . . . . . . . . 10

2 Theory 13
2.1 Fluid mechanics . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Fluid flows . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2 Boundary layer equations . . . . . . . . . . . . . . 15
2.1.3 Navier-Stoke and continuity equation . . . . . . . . 17
2.2 Solid Mechanics . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.1 Mechanical stress and strain . . . . . . . . . . . . . 18

vii
2.2.2 Fatigue . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Numerics . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.1 Computational Fluid Dynamics (CFD) . . . . . . . 25
2.3.2 Finite Element Method (FEM) . . . . . . . . . . . 27

3 Modeling 31
3.1 CFD model . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.1.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . 33
3.1.2 Preprocessing . . . . . . . . . . . . . . . . . . . . . 33
3.2 FEM model . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . 36
3.2.2 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.3 preprocessing . . . . . . . . . . . . . . . . . . . . . 39

4 Results 41
4.1 CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 FEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.2.1 Mean stress . . . . . . . . . . . . . . . . . . . . . . 48
4.2.2 Stress amplitudes . . . . . . . . . . . . . . . . . . . 50

5 Discussion and
Conclusions 53
5.1 CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.2 FEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . 56

Bibliography 59

A Figures 61
A.1 Figure from site . . . . . . . . . . . . . . . . . . . . . . . . 61
A.2 Figures from the FEA . . . . . . . . . . . . . . . . . . . . 64

B Tables 69
B.1 Site measurements . . . . . . . . . . . . . . . . . . . . . . 69
B.2 FEA results . . . . . . . . . . . . . . . . . . . . . . . . . . 72

C Matlab source code 75


C.1 Result presentation . . . . . . . . . . . . . . . . . . . . . . 75
C.2 Deviation angle calculations . . . . . . . . . . . . . . . . . 79
List of Figures

1.1 Simplified schematic of a Hydropower plant. . . . . . . . . 3

1.2 Turbine types and there operational head span. The im-
age is gathered from Krivchenko 1994 [16]. . . . . . . . . . 4

1.3 Figure showing illustration of a Kaplan turbine and a


schematic of an adjustable blade Kaplan turbine. (The
images are printed with permission from Rainpower Nor-
way AS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.4 Illustration of a Pelton turbine.(The images are printed


with permission from Rainpower Norway AS) . . . . . . . 6

1.5 Simplified schematic of a Pelton turbine.(The image is an


illustration printed with permission from Rainpower Nor-
way AS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.6 Illustration of a Francis turbine. (The image is an illus-


tration printed with permission from Rainpower Norway
AS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

ix
x LIST OF FIGURES

1.7 Simplified schematic of a low head Francis turbine with


terminology. 1. Stayring vanes, 2. Guide vanes, 3. Wicket
gate lower ring, 4. Turbine cover, 5. Support bearings of
guide vane upper pivot, 6. Attachment of cover of stayring
upper band, 7. Shaft flange, 8. Shaft, 9. Runner crown,
10. Runner band, 11. Runner blades, 12, 13, 14. Oper-
ating gear, 15- Servomotor, 16. Bearing, 17. Generator
thrust bearing support, 18. Runner cone, 19. Runner
band seal, 20. Blanching hole. The image is gathered
from Krivchenko 1994 [16]. . . . . . . . . . . . . . . . . . . 9
1.8 Different designs for Francis runners [2]. a.High head, b.
Medium head and c. Low head. (The image is an illus-
tration printed with permission from Rainpower Norway
AS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.9 The figure shows the location of the strain gauges on the
new runner, blade 7. Picture (a) shows the locations of
the gauges on the pressure side. Picture (b) shows the
locations on the suction side. The images are prints from
the site measurement report [4]. . . . . . . . . . . . . . . . 11
1.10 The formation of the rosette on blade 14 of the original
runner. The figure also illustrates the error estimation
shown to the right. The image is a print from the site
measurement report [4]. . . . . . . . . . . . . . . . . . . . 12

2.1 Figure showing a bar with a diameter D, length L, cross


section area after deformation is A, the material has the
Youngs modulus E. The force F applied on the bar has
resulted in an elongation ². b) Shows the nominal stress
σ at the cross section. . . . . . . . . . . . . . . . . . . . . 18
2.2 The figure shows a standard stress strain curve obtained
by tensile testing. The figure to the right shows the elastic
contra the plastic region in the curve. On the figure to the
left, σt is the tensile strength, σy is the yield strength, ²t
is the tensile strain and ²f is the fracture strain. . . . . . . 20
2.3 The figure shows the same cross section as in figure 2.1b
but here a defect in the form of a vacant sphere. . . . . . 20
2.4 The figure shows the stress concentration condition for a
small hole with diameter d in an infinitely large and thin
plate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
LIST OF FIGURES xi

2.5 Figure showing simplistic view of the methodology of the


discretization process. The image is a modified one that
originally was found in Felippa 2002 [10]. . . . . . . . . . . 24
2.6 Figure illustrating the shapes, degrees and node positions
for some elements. . . . . . . . . . . . . . . . . . . . . . . 28

3.1 Figure of the designed runner model. (a) Showing a trig


view, (b) a side view,(c) a top view and (d) a bottom view. 32
3.2 Figure showing the CFD model used in the analysis. It
represents a third of the total geometry shown in figure 3.1. 33
3.3 Figure of the CFD Mesh. (a) Showing a top view, (b) a
zoom in, for a closer look at the mesh. . . . . . . . . . . 34
3.4 Figure showing one of the simplifications on the design
model in order to save CPU time. (b) The zoom in design
model contour on the band, (c) a zoom in, for a look at
simplification done at the area. . . . . . . . . . . . . . . . 35
3.5 Figure showing the finite element geometry of model one
used in the analysis. Containing a splitter blade, blade
and part of the crown and band. . . . . . . . . . . . . . . 36
3.6 Figure showing the finite element geometry of model two
used in the analysis. a) showing a back view and b) a front. 37
3.7 Figure showing the type of solid mesh element used in
the analysis. It is a ten node tetrahedra element with
quadratic displacement behavior. The image is one gath-
ered from the Ansys 11 manual pages [1]. . . . . . . . . . 38
3.8 Figure showing the type of surface mesh element used in
the analysis. It is a Second order elements with four (on
the right) to eight (on the left) nodes. The image is one
gathered from the Ansys 11 manual pages [1]. . . . . . . . 38
3.9 Figure showing the finite element mesh of model two used
in the analysis. The figure is showing a pressure side view. 38
3.10 Figure showing the boundary conditions for the finite el-
ement analysis. . . . . . . . . . . . . . . . . . . . . . . . . 40
xii LIST OF FIGURES

4.1 Figure showing the results obtained by the CFD calcula-


tion. The figure on the top, figure a, shows the pressure
results normalized with respect to the head of the power
plant and figure b shows a contour plot of the velocity
at the same positions. The velocity is normalized with
respect to the the maximum velocity vmax . . . . . . . . . . 42

4.2 The figure is showing the pressure results on the surfaces


of the blades. It should be noted the the contour plot
presented here is taken at on certain time step out of the
transient CFD calculation. . . . . . . . . . . . . . . . . . . 43

4.3 Figure showing the results obtained by the FEA calcula-


tion. The figure on the top, figure a, shows the maximum
principle stress for a given time step. The plotted contour
is of σ/σmax . For a closer look of the legend values se
figures given in the appendix, figure A.3-A.5 . . . . . . . . 45

4.4 The figure shows the quasi static stress variation on the
pressure side at different point, at various locations from
the trailing edge of the blade. . . . . . . . . . . . . . . . . 46

4.5 The figure shows the quasi static stress variation on the
suction side at two different locations from the trailing
edge of the blade. . . . . . . . . . . . . . . . . . . . . . . . 47

4.6 A plot of the mean stress at different locations from the


trailing edge on the blade. The mean stress calculated
from the site measurements is here compared to the FEA
calculated results. . . . . . . . . . . . . . . . . . . . . . . . 48

4.7 A figure showing plots of the mean stress deviation as a


function of the distance from the trailing edge. . . . . . . 49

4.8 A plot of the mean stress at different locations from the


trailing edge on the blade. The mean stress calculated
from the site measurements is here compared to the FEA
calculated results. . . . . . . . . . . . . . . . . . . . . . . . 50

4.9 A figure showing plots of the stress amplitude deviation


as a function of the distance from the blade trailing edge. 51
5.1 The figure illustrates the geometry difference at the trail-
ing edge of the blade between the FEA model and the
real runner. The figure to the left represents the shape
profile of the FEA model and the one on the right is that
of the real runner. It should be noted that the figure is
only an illustration of the difference and is in no way an
exact representation of the geometry. . . . . . . . . . . . . 55

A.1 The figure shows the location of the strain gauges on blade
7. The Picture shows the locations of the gauges on the
pressure side. The image is a print from the site measure-
ment report[4]. . . . . . . . . . . . . . . . . . . . . . . . . 62
A.2 The figure shows the location of the strain gauges on blade
7. The Picture shows the locations on the suction side.
The image is a print from the site measurement report[4]. 63
A.3 Figure showing the results obtained by the FEA calcula-
tion. The figure shows the maximum principle stress for
a given time step. The plotted contour is of σ/σmax . . . . 65
A.4 Figure showing the results obtained by the FEA calcula-
tion. The figure shows a zoom in for a better picture of
the trailing edge at the pressure side. The contour plot
gives the distribution of the maximum principle stresses
for a given time step. The plotted contour is of σ/σmax . . 66
A.5 Figure showing the results obtained by the FEA calcula-
tion. The figure shows a zoom in for a better picture of
the trailing edge at the suction side. The contour plot
gives the distribution of the maximum principle stresses
for a given time step. The plotted contour is of σ/σmax . . 67
A.6 Figure showing the results obtained by the FEA calcula-
tion. The contour plot gives the distribution of the max-
imum principle stresses for a given time step. And the
streamlines color code responds to the local velocity of the
fluid. The figure was only meant to serv as an illustrative
picture of the interaction between fluid and structure and
nothing else. . . . . . . . . . . . . . . . . . . . . . . . . . . 68
xiv LIST OF FIGURES
List of Tables

B.1 The table presents the normalized mean stress measured.


The stresses are normalized with respect to the maximum
measured mean stress. . . . . . . . . . . . . . . . . . . . . 70
B.2 The table presents the normalized stress amplitudes at the
guide vane passing frequency 150Hz. The amplitudes are
normalized with respect to the maximum measured stress
amplitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
B.3 The table presents the normalized FEA principle stresses.
The stresses are normalized with respect to the maximum
measured mean stress on site. σ denotes the mean stress
and ∆σ2 stands for the stress amplitude. It shoould be
noted that the ∆σ 2 was normalized with respect to the
maximum measured stress amplitude. . . . . . . . . . . . 73
CHAPTER 1. INTRODUCTION

Chapter

Introduction

The mechanical energy in flowing water has been exploited for centuries.
Waterwheels driven by the flow of water in rivers were used in ancient
India mainly for irrigation of crops. They were used for grinding grain by
the ancient Greeks and the Romans were known for using waterwheels
for mining. Although useful for many purposes, waterwheels are not
very effective in harnessing the full potential of the kinematic energy of
flowing water.

The first development of a more modern turbine was done in the 8th
century by professor Ján Andrej Segner and is today called the Segner
turbine. The unit was an axial machine, much smaller than a water-
wheel, that raised the limit for operating heads and flow rate [21]. The
development of water turbines took off during the industrial revolution
and in 2005 the existing capacity of hydroelectric power was 816 GW
which was about 21% of the worlds electricity supply at that time [18].

1
1.1. BACKGROUND CHAPTER 1. INTRODUCTION

1.1 Background

The mechanical integrity of turbine runners is put to its very limits with
increasing demand for higher efficiency to a lower cost. Designers today
have access to design tool that enable them to push the mechanical limits
further than ever before. Computational methods have given designers a
deeper understanding of the complex turbulent flow and a greater knowl-
edge of the loads they impose on mechanical structures.

Turbines are pushed to be more available and made to operate at a


wider rang of operation loads, which means frequent start and stop cy-
cles and a drift at of best efficiency operational conditions. Since turbines
are manufactured to work at best efficiency for a longer period of time,
they experience loads they are not dimensioned for and thus the risk
of fatigue failure is increased. Designers are facet with fatigue, fatigue
loading, corrosion, cavitation, corrosion fatigue, stress fatigue and many
other challenges when designing a new runner.

Turbine runners are subjected to high dynamic loading that could even-
tually lead to fatigue failure. Dynamic fatigue loading cycles are of major
concern. They are manly categorized in two sections, High Fatigue Cy-
cles (HCF) and Low Fatigue Cycles (LCF). LCF originates from Start
stop cycles and HCF from fatigue loading during operation. The main
source for HCF is thought to be the rotor stator interaction [12].

Fatigue failure generally occurs in areas of high stress concentration


where dynamic stress amplitudes are the largest. Structural loading
fatigue failure in runners has been observed both in early life and after
several years of operation. Fatigue cracking propagate from areas of high
stress concentration [12]. For high head Francis runners, the transitional
welded area between the blade and band or crown have been identified
as critical areas [8]. The risk for fatigue failure is specially high near
the trailing edge of the blade. Structural failure originate from vari-
ous different factors but the major causes are thought to be dynamic
hydraulic load fluctuations combined with material defects, especially
defects within the weld [8].

With computational methods such as Computational Fluid Dynamics


(CFD) and Finite Element Analysis (FEA) we are able to better un-

2
CHAPTER 1. INTRODUCTION 1.2. HYDRAULIC MACHINES

derstand the cause and that way improve the mechanical properties of
future runners.

1.2 Hydraulic machines

There are many different types of hydraulic machines in use at different


power and pump stations worldwide. The choice of a machine depends
on the application and the installation site. They are chosen mainly
based on the head, headwater height over the tailwater, and flow rate at
the specific reservoir. Figure 1.1 shows a simplified schematic of a Hy-
dropower plant. The most commonly used turbines are Kaplan, Francis
and Pelton turbines. The reason why these types are the most predom-
inate today [12], is that they supplement each other. They cover most
operation heads and discharge water flows, figure 1.2 shows an approxi-
mate head span for different turbine types [16].

Figure 1.1: Simplified schematic of a Hydropower plant.

3
1.2. HYDRAULIC MACHINES CHAPTER 1. INTRODUCTION

Figure 1.2: Turbine types and there operational head span. The image
is gathered from Krivchenko 1994 [16].

1.2.1 Kaplan turbines


Another classification for Kaplan turbines is axial flow turbine, this is
because the fluid flows axially through the runner. It is a low-head
machine used for heads ranging approximately from 10 to 70 meters and
high flow rates [16]. Figure 1.3(b) shows a schematic of a Kaplan turbine.
As the fluid flows through the runner the static pressure decreases and
mechanical energy in transferred to the runner driving the shaft. Kaplan
runners with adjustable blades have a larger range of discharges and are
used at sites where there is a high head variation [16].

4
CHAPTER 1. INTRODUCTION 1.2. HYDRAULIC MACHINES

(a) Illustration of a Kaplan turbine (b) 1. Runner, 2. Guide vane, 3. Outer


cover, 4. Lower cover, 5. Runner chamber, 6.
Stay ring, 7. Spiral casing, 8. Vane arm, 9.
Link, 10. Regulating ring, 11. Inner cover,
12. Turbine shaft, 13. Turnťbine bearing,
14. Shaft seal box, 15. Thrust bearing, 16.
Draft tube cone, 17. Drain pump, 18. Inter-
mediate shaft, 19. Servo motor.

Figure 1.3: Figure showing illustration of a Kaplan turbine and a


schematic of an adjustable blade Kaplan turbine. (The images are
printed with permission from Rainpower Norway AS)

5
1.2. HYDRAULIC MACHINES CHAPTER 1. INTRODUCTION

1.2.2 Pelton turbines

Figure 1.4: Illustration of a Pelton turbine.(The images are printed with


permission from Rainpower Norway AS)

Pelton turbines are used manly for extremely high heads with low
discharge [16]. A Pelton turbine operates by one or more nozzles im-
pinging jets onto buckets placed circumferentially on the runner disk, se
figure 1.5. The buckets change the direction of the flow and the resulting
change in momentum exerts a force on the buckets, thus driving the run-
ner [12]. The resulting change in momentum is why Pelton turbines are
also classified as impulse machines. The runner is mounted vertically or
horizontally and both runner and nozzle are located over the tailwater
[16].

6
CHAPTER 1. INTRODUCTION 1.2. HYDRAULIC MACHINES

Figure 1.5: Simplified schematic of a Pelton turbine.(The image is an


illustration printed with permission from Rainpower Norway AS)

For impulse turbines the static pressure is constant over the runner.
In the case of a reaction turbines the pressure decreases as the fluid
flows through the runner. Francis and Kaplan are examples of reaction
turbines [12].

7
1.2. HYDRAULIC MACHINES CHAPTER 1. INTRODUCTION

1.2.3 Francis turbines

Figure 1.6: Illustration of a Francis turbine. (The image is an illustration


printed with permission from Rainpower Norway AS)

Francis turbines are also classified as radial-axial flow turbines. The


fluid flows radially, with respect to the shaft, in at the inlet of the run-
ner and exits axially at the outlet. Francis turbines are use for heads
typically ranging from 20 to 900m [16]. A simplified scheme of a Francis
turbine can be seen in figure 1.7. Francis runners are as the case for Ka-
plan runners submerged in water. Newtons third law of motion describes
the energy transfer from fluid to runner for these types of turbines, se
below [5].

Newtonťs third law:

When two particles interact, the force on one particle is equal and oppo-
site to the force on the other.

The water is lead from the reservoir via a pressure conduit, the pen-
stock, to the spiral casing. The spiral casing is designed to distribute
a uniform flow in to the runner. The fluid passes the stay vanes and
goes through the guide vane channels onto the runner blades. Kinematic

8
CHAPTER 1. INTRODUCTION 1.2. HYDRAULIC MACHINES

Figure 1.7: Simplified schematic of a low head Francis turbine with ter-
minology. 1. Stayring vanes, 2. Guide vanes, 3. Wicket gate lower ring,
4. Turbine cover, 5. Support bearings of guide vane upper pivot, 6.
Attachment of cover of stayring upper band, 7. Shaft flange, 8. Shaft,
9. Runner crown, 10. Runner band, 11. Runner blades, 12, 13, 14. Op-
erating gear, 15- Servomotor, 16. Bearing, 17. Generator thrust bearing
support, 18. Runner cone, 19. Runner band seal, 20. Blanching hole.
The image is gathered from Krivchenko 1994 [16].

energy is then absorbed from the fluid by the blades driving the shaft
connected to the generator. The water is then lead through the draft
tube to the tailwater.

Francis runner

Francis runner designs vary depending on the head. The designs are
distinguished between low, medium and high head turbines [2], se firure
1.8. A high head Francis runner may also consist of splitter blades as well
as blades to obtain the maximum efficiency possible for higher heads.

9
1.3. SITE MEASUREMENTS CHAPTER 1. INTRODUCTION

Figure 1.8: Different designs for Francis runners [2]. a.High head, b.
Medium head and c. Low head. (The image is an illustration printed
with permission from Rainpower Norway AS)

1.3 Site measurements

A new unit was installed at Tonstad hydropower plat in Norway. The


unit has a nominal head of 430 m, a nominal power of 160 M P a and
a rotational speed of 375 rpm. The runner replaced one of the original
runners and had a new design.

Measurements where carried out to investigate the difference in mechan-


ical properties between the new and the original design. Strain gauges
were mounted near the trailing edge of the blade close to the band, on
unit I (the original runner) and unit II (the new runner). The strain
gauges were mounted on blade 7 and 14 both unites. The strain gauges
mounted at blade seven were installed as a precaution, in case the gauges
on blade 14 would destroy during the test [4].

Six strain gauges were installed on each blade, four on the pressure side
and two on suction side. Gauges were installed at 15 mm (P S15), 25 mm
(P S25), 200 mm (P S200) and 300 mm (P S300) on the pressure side,
for the new runner, and 15 mm and 25 mm on the suction side, se figure
1.9. All the strain gauges were mounted at a distance of 25 mm from
the band. The strain gauges installed on the original runner had the
same positions except for P S200 which was mounted 175 mm from the

10
CHAPTER 1. INTRODUCTION 1.3. SITE MEASUREMENTS

trailing edge to make room for a rosette strain gauge [4].

(a) (b)

Figure 1.9: The figure shows the location of the strain gauges on the
new runner, blade 7. Picture (a) shows the locations of the gauges on
the pressure side. Picture (b) shows the locations on the suction side.
The images are prints from the site measurement report [4].

The major principle stress, σ1 , was assumed to have a direction perpen-


dicular to the band. The transverse stress was assumed to be insignifi-
cant. That introduced an error which needed to be investigated [4].

The error of measuring the strain in a single direction was estimated


by installing a rosette on the original runner. The rosette was installed
according to figure 1.10 and was placed on the pressure side 200 mm
from the trailing edge on blade 14 on the original runner. As seen to the
right in figure 1.10, an error of 0.14% was found after the test [4].

Site measurements on the new runner were used for validating the dy-
namic stress investigation in the thesis, therefore site measurement re-
sults presented in the thesis are from measurements on the new unit.
The measurement for that test was not fulfilled according to the planed
test. During the first try, the data logger used, only manege to register
results for the first 30 minutes of a 60 minute test session. Other test
tries were unsuccessful in improving the first try. The first test mea-
surements covered start up and uploading to 290 mm servomotor stroke
(81% load), full load is at 343 mm servo stroke [4].

11
1.3. SITE MEASUREMENTS CHAPTER 1. INTRODUCTION

Figure 1.10: The formation of the rosette on blade 14 of the original


runner. The figure also illustrates the error estimation shown to the
right. The image is a print from the site measurement report [4].

The maximum mean stress at 290 mm servomotor stroke was registered


to be located at P S15. The mean stresses gradually decreased as the
distance from the trailing edge increased. The largest fluctuations oc-
curred during startup, at speeds near the natural frequency of one of the
blades. That seemed to have triggered interference between mechanical
and hydraulic resonance [4].

12
CHAPTER 2. THEORY

Chapter

Theory

2.1 Fluid mechanics


Fluid mechanics describes the behavior of fluids. It can be classified in
three major subbranches. The science describing fluids at rest called fluid
statics, in motion (fluid dynamics) and the interaction between solid and
fluid [25].

2.1.1 Fluid flows


Laminar flow is typically apparent for high viscosity fluids with low ve-
locity. A laminar flow is smooth and ordered. If the flow is disordered it
is classified as turbulent. Turbulent flow usually appear at high veloci-
ties and one of the characteristics is velocity fluctuations throughout the
domain [25].

A flow is defined as steady if the flow is time independent and unsteady


otherwise. As steady implies no change at a point in time, uniform
implies no change in location within the domain. Unsteady flow which
fluctuate in the same manner about a steady state are defined as periodic
[25].

13
2.1. FLUID MECHANICS CHAPTER 2. THEORY

Viscosity
When a fluid in contact with another fluid is forced to move, it exerts
a shear force through the contact surface in the flow direction on the
stationary fluid. The same is apparent if the moving fluid is replace for-
instance with a solid plate in motion. This is due to the exitance of a
property quantified by what is called viscosity. It is the frictional force
developed as the two fluids are forced to move relative to each other.
Viscosity can be seen as a measure of the resistance of deformation in
the fluid [25].

If a fluid rate of deformation is proportional to the shear stress on the


contact surface, the fluid is classified as Newtonian fluid other wise the
fluid is non-Newtonian. The relationship between rate of deformation
and the shear stress for a Newtonian fluid in one-dimension is described
by equation 2.2 [25]. The relationship is derived and expressed by an
experiment with laminar flow of a fluid between two large plates held
at a certain distance from each other. One plate is forced to move at
a certain velocity while the other is kept stationary. The velocity field
between is linear and can be expressed by equation 2.1 [25].
y
u(y) = V (2.1)
l
Where u is the the velocity at a certain point, y is the vertical distance
from the stand-still plate, l is the distance between the plates and V
stands for the velocity of the moving plate.

du
σs = µ (2.2)
dy

Where σs stands for the shear stress acting on the fluid layer and µ stands
for the dynamic viscosity. The relation 2.2 can be used to calculate the
viscosity of a fluid. Therefore the above described experiment can give
a measure of the viscosity [25].

The ratio between the dynamic viscosity and the density of a fluid is
a recurring term in fluid mechanics. The term is therefore given a name,
kinematic viscosity ν, se equation 2.3.
µ
ν= (2.3)
ρ

14
CHAPTER 2. THEORY 2.1. FLUID MECHANICS

Viscous effects are involved in all fluid flows. The effects are though
more significant in some regions of the flow domain than others. In
these regions the flow is classified as viscous. Regions where the internal
frictional forces are negligibly small compared to for-instance pressure
forces, the flow region is called inviscid flow region [25]. This enables
approximations simplifying the complexity of the flow, while still keeping
the accuracy level of the analysis high.

Compressibility
Compressibility defines the degree in which a fluid changes its density
throughout a flow domain. If a volume portion of a fluid varies depending
on the pressure at different points in the flow field, the fluid is defined
as compressible and incompressible if otherwise. All fluids change in
density to some degrees when pressurized, so incompressibility is actually
an approximation [25]. A fluid is said to be incompressible if the density
remain reasonably constant. Liquids keep there densities nearly constant
and therefore are usually called incompressible substances [25]. Gases
compressibility is partially model after the mach number, the speed of
the gas relative to the speed of sound, se equation 2.4. If the change in
density of a gas is under 5%, the flow is approximated as incompressible,
usually when M a < 0.3 [25].
u
Ma = (2.4)
c
u is the velocity of the fluid and c stands for the speed of sound.

2.1.2 Boundary layer equations


Professor Claude-Louis Navier and George Gabriel Stokes formulated the
equations of fluid motion, further discussed in section 2.1.3, during the
beginning of the 1800s. Even though the Navier-Stokes equation was
well known in the mid seventeens century, it is a complex differential
equation that could only be solved for simple flow geometries. Scientists
were able to solve approximations of the equation (the Euler equation,
equation 2.5, and the Bernoulli equation, equation 2.6) [25], but these
simplification by them self were not enough to resolve the complete flow,
therefore these solutions could not be validated by experiments.

The Euler equation approximation is derived from the Navier-Stokes

15
2.1. FLUID MECHANICS CHAPTER 2. THEORY

equation by the following assumption [25].

Euler equation approximation:

Viscous forces << pressure and/or inertial forces

The bernoulli equation eq. 2.6 is derived for stationary and incompress-
ible flows[25].

∂~u 1
+ ~u · ∇~u = − ∇P + F~ (2.5)
∂t ρ
P U 2
~u · ∇( + + F~ ) = 0 (2.6)
ρ 2

u is the velocity of the fluid, U is the velocity average, P is the pressure


and F stands for the external force acting on the fluid.

The assumption means that the viscous term in the Navier-Stokes equa-
tion can be neglected. This assumption is not valid close to solid walls
where it would mean free slip condition, which does not yield a physical
solution. The viscous forces near a solid wall cause the no slip condition,
they are therefore significant and can not be neglected.

The breakthrough came in 1904 by the German physicist Ludwig Prandtl.


He suggested that the flow was to be divided into two regions, an inner
and an outer flow region. The outer flow regions velocity field would
be solved by the Euler and continuity equation. The Bernoulli equation
would be used for solving the pressure field for the region, where viscous
forces can be neglected. The inner region, called boundary layer, is a
very thin layer where viscous forces are not negligible and the flow is
rotational. Professor Prandtls approximation is what is known as the
boundary layer approximation [25].

The boundary layer is assumed to be very thin and the pressure is shown
to be constant across a boundary layer [25]. The boundary layer for in-
compressible flows is described by equation 2.7.

1
~u · ∇~u = − ∇P + ν∇2 ~u (2.7)
ρ

16
CHAPTER 2. THEORY 2.2. SOLID MECHANICS

2.1.3 Navier-Stoke and continuity equation

The motion of fluid is mainly described by two differential equations. The


conservation of mass also known as the continuity equation, equation 2.8,
and the Navier-stokes equation, equation 2.9. These equations combined
are able to fully solve the details of the flow for every point in the domain
[25]. These differential equations are extremely complex and difficult
to solve. The equations are usually combined with other equations and
necessary approximation in order to be solved analytically. However with
the help of computational fluid dynamics a solution of the equations of
motion can be obtain, the method is described further in section 2.3.1.


+ ρ(∇ · ~u) = 0 (2.8)
Dt
D~u 1
= − ∇P + ν∇2 ~u + F~ (2.9)
Dt ρ

2.2 Solid Mechanics

Material strength testing and theories thrived during the 20th century.
The engineer Alan Arnold Griffith made a profound contribution the the
study of materials when he published his first work on the subject in
1920, "The phenomenon of rupture and flow in solids". He made a series
of tensile tests with glass bars. The test showed that the tensile strength
σt , the stress needed to pull a material to the breaking point, seemed to
decrees with increasing specimens thickness. The discovery lead him to
the following conclusion, there are defects within the material leading to
it breaking at much lower stresses than theoretically anticipated. The
number of defects is increased with increased bar thickness and that is
the cause for lower tensile strength. The conclusion of the study was
only valid for extremely brittle materials, since it only took the elastic
work into consideration [14].

The theory was further modified in the 1940s by the scientist George
Rankine Irwin and the team under him during his time at the US Naval
Research Laboratory. He included the plastic work done during tensile
testing and developed the concept of stress intensity factors [14].

17
2.2. SOLID MECHANICS CHAPTER 2. THEORY

2.2.1 Mechanical stress and strain


The cohesive forces between atoms within a material defines the ideal
strength of the material. The word ideal is used here because such
strength refers to a material with no discontinuities, defects, and as is
known such a material does not exist in reality. The theory is through
still applicable in certain loading cases discussed further in the following
sections.

The strength is quantified by the material property Youngs modulus


E. E is given by Hooks law and it is obtained by calculating the ratio
between the normal stress applied and the strain, elongation or elas-
tic deformation, it forces on the material, se equation 2.12 [9]. Figure
2.1 illustrates the standard visualization of the stress-strain relation, the
stress is calculated by equation 2.10 and the strain by equation 2.11 [5].

(a)

(b)

Figure 2.1: Figure showing a bar with a diameter D, length L, cross


section area after deformation is A, the material has the Youngs modulus
E. The force F applied on the bar has resulted in an elongation ². b)
Shows the nominal stress σ at the cross section.

F
σ= (2.10)
A

18
CHAPTER 2. THEORY 2.2. SOLID MECHANICS

∆L
²= (2.11)
L
σ
E= (2.12)
²
Where σ is the stress and ² is the elongation. As seen in equation 2.12,
the relation between stress and strain is assumed to be linear. This is
approximately true for a material, except for polymers, in the elastic
domain before plastic deformation occurs.

In the case of a nonuniform cross section stress, due to discontinuities or


geometrical changes in the structure, the following derivation results in
an expression for the stress distribution.

Equation 2.10 gives the following for a one dimensional stress distri-
bution:
∆F
σc = (2.13)
∆A
σc stands for the cross section stress. To look at the stress in one point
in the cross section, δA → 0, equation 2.13 becomes:
∆F dF
σc = lim = (2.14)
∆A→0 ∆A dA
So the stress can vary at different points on the cross section area. The
net force can then be obtained as follows [9].
dF = σc (A)dA ⇒ (2.15)
Z
⇒ Fnet = σc (A)dA (2.16)

Deformation
If a test specimen is loaded with a load that does not exceed the materials
yield strength σy , the deformation forced on the specimen will disappear
when the load is removed, the test subject returns to its original shape.
This type of deformation is classified as elastic. Plastic deformation
occurs if σy is exceeded. In this case, when the load is released the
material will recover the elastic deformation but keep the plastically
deformed length [23, 14]. Figure 2.2 shows a standard stress-strain curve
plotted for a material after tensile testing. The figure also gives an idea
of the information such a curve reveals about the material properties of
the specimen.

19
2.2. SOLID MECHANICS CHAPTER 2. THEORY

Figure 2.2: The figure shows a standard stress strain curve obtained by
tensile testing. The figure to the right shows the elastic contra the plastic
region in the curve. On the figure to the left, σt is the tensile strength,
σy is the yield strength, ²t is the tensile strain and ²f is the fracture
strain.

Stress concentration factor


The definition of stress concentration is, a local elevation of stress levels.
Consider the setup in figure 2.1. If a deficiency is introduced within the
bar, in the form of a vacant sphere of radius R, take out a small sphere
of material as sown in figure 2.3. The same loading condition is applied.

Figure 2.3: The figure shows the same cross section as in figure 2.1b but
here a defect in the form of a vacant sphere.

The stress sheared by the present material in the previous case is now
sheared by the material around the defect. This leads to elevated stress
levels around the sphere, an area of stress concentration appears [9]. The
cross section stress distribution is no longer homogeneous. The stress dis-
tribution seen in figure 2.1b is called the nominal stress and is noted as
σn . The stress distribution around the sphere is much more complex and

20
CHAPTER 2. THEORY 2.2. SOLID MECHANICS

three dimensional than illustrated in figure 2.3 [9].

Lets look at the case for an infinite and thin plate, loaded with the
nominal tensile stress σn . The plate has a hole with the diameter d, see
figure 2.4.

Figure 2.4: The figure shows the stress concentration condition for a
small hole with diameter d in an infinitely large and thin plate.

The stress distribution near the hole is the same as in the previous case,
it is complex, but the stress at the circle curve is shown to be described
by equation 2.17 [9].

σθ = (1 − 2cos(2θ))σn (2.17)

Equation 2.17 gives σθ = 3σn for θ = 90◦ and θ = 270◦ , which is noted
as σmax . The equation also gives σθ = −1σn for θ = 0◦ and θ = 180◦ ,
which is the minimum stress at the circle. The negative value means
that the stresses in these areas are contractive. The stress concentration
factor is derived from the maximum stress and is formulated by equation
2.18 [12], it should be noted that the equation is only valid for a circular
hole. In this case the concentration factor is Kt = 3.
σmax
Kt = (2.18)
σn
The expressions of Kt differ for different types of discontinuities. Dia-
grams, tables and exact expressions of the solution of Kt are today given
in books available for construction engineers and scientists.

2.2.2 Fatigue
Mechanical structures have been known to reach fracture failure at load
levels below the tensile strength of the material, fractures have also oc-
curred at stress levels even below the yield strength. This phenomenon is

21
2.2. SOLID MECHANICS CHAPTER 2. THEORY

a progressive structural damage known as fatigue. A structure subjected


to load cycles can fail if the load amplitude is high enough to achieve
fracture, it is called fatigue failure or fatigue fracture [12].

Fatigue loading
Fatigue life assessment are often characterized by S − N curves, also
known as Wöhler curves. The curve plots the stress (S) against the
number of cycles (N ) needed to achieve fatigue failure [12]. Fatigue fail-
ure often starts with crack nucleation (initiation) and then as the load
cycles continue, the crack grows continuously eventually leading to frac-
ture [12].

There are two main types of fatigue loading, Low cycle fatigue (LCF)
and high cycle fatigue (HCF). Low cycle fatigue, is defined at approxi-
mately N = 103 , and occur where the stress levels are high enough to
cause plastic deformation. LCF is often characterized by Coffin-Manson
relation given by equation 2.19.
∆²p
= ²0f (2N )c (2.19)
2
∆²
Where 2 p is the plastic strain amplitude, ²0f is the empirical con-
stant fatigue ductility coefficient and c is the empirical constant fatigue
ductility exponent. HCF is usually defined for cycles ranging between
N = 104 − 108 . HCF occur for load cycles with stress levels below the
yield strength where the deformation is primarily elastic.

Fatigue crack and crack growth


G. R. Irwin extended theory of A. A. Griffiths work lead Irwin and
his team to the formulation of the stress intensity factor, K. Griffith
formulated the concept, that an initiated crack in a brittle environment
will grow during loading if the total energy of the system is lower by the
crack propagation. Írwens work based on Griffiths findings developed a
theory for crack propagation in ductile material and formulated equation
2.20 [12]. √
K = Cσn πa (2.20)
Here a is the crack length and C is a dimensionless constant derived for
the crack shape and the geometry around the crack. Exact solutions of

22
CHAPTER 2. THEORY 2.3. NUMERICS

the crack tip stress field for an infinite plate and relatively simple load
cases are available in engineering books.

The solutions for stress intensity factors available are often calculated for
simpler load cases. Mechanical structures inhabit more complex stress
fields originating from various sources, like residual stress fields and stress
field at notches (geometrical discontinuities). Due to these discontinu-
ities stress concentration is evident, which lead to gradient stress fields.
The stress gradient, χ, given by equation 2.21, for propagating cracks
are in most cases not constant [12]. So solving the stress intensity factor
is a complex and non trivial task.
¯ ¯
1 ¯¯ δσy ¯¯
χ= (2.21)
σmax ¯ δx ¯x=x0

The stress gradient in equation 2.21 is given for a spherical cavity in an


infinite body in tension in the y axes direction. Here x0 is where the
maximum stress is located. The crack is shown to grow in the direction
of the largest stress gradient and perpendicular to the maximum mean
stress [12].

To accurately predict the crack propagation of a crack in a mechani-


cal structure, the geometry factor, F , and the stress gradient are key
components. The stress field at the propagation front is influenced by
the stress gradient and visa versa. This means that in order to obtain a
solution as accurate as possible, χ is preferably recalculated after each
crack growing load cycle [12].

2.3 Numerics
Numerical analysis is a representation of continues mathematics by al-
gorithms. Numerical mathematics has been around for a long time.
The earliest numeric mathematical find dates back to somewhere be-
tween 1800 − 1600BC in ancient Babylonia. The find was a clay tablet

with an engraved numerical approximation of the square root of 2, 2.
Great mathematicians and physicists like Isaac Newton, Joseph Louis
Lagrange, Carl Friedrich Gauss and Leonhard Euler have formulated
numerical algorithms used for a wide range of computational applica-
tions [19].

23
2.3. NUMERICS CHAPTER 2. THEORY

Computational numerics can be seen as the discretization of the physical


setup of the problem at hand. The methodology and process involved in
applying such calculations on a physical problem is summarized by the
figure 2.5 [10].

The process called idealization describes going form the actual physical

Figure 2.5: Figure showing simplistic view of the methodology of the


discretization process. The image is a modified one that originally was
found in Felippa 2002 [10].

system to a simplified mathematical model. The model is designed ex-


tracting certain aspects of behavior of the physical system, depending on
the application. Therefore the idealization process is the most sensitive
and important step, since there are many factors that need to be taken
into account. Mathematical modeling is basically a filtering step, where
physical details which are not of interest for the analysis are filtered
out [10]. These models are usually a system of coupled partial differ-
ential equation in space and time. They are often complex not easy to
solve. The variables of such a system are called degrees freedom (DOF).
A mathematical idealized model has an infinite number of DOFs, it is
what is called a continues model [10].

It is necessary to reduce the number of DOFs to a finite number, in


order to gain a numerically handleable and solvable system, this step
is called discretization (transferring continues equations into discreet
counterparts) [10]. A discreet model is obtained by spatial and time
discretization. The discreet model is only an approximate model of the
mathematical one, so validation of the process is of upmost importance.

24
CHAPTER 2. THEORY 2.3. NUMERICS

Every step is an error source introducing errors that need to be un-


derstood and accounted for in the final solution. The only way to really
do that is by studying the real physical systems behavior, what is meant
by that statement is validation against experimental results.

2.3.1 Computational Fluid Dynamics (CFD)


Computational fluid dynamics has made a profound impact on the de-
sign process for engineer. This section will briefly go through the solution
methodology and theory of CFD calculations. The CFD method is used
for numerical analysis of the flow of fluids. The equations of motion are
as mentioned earlier a set of partial differential equation (PDEs) that
describe the flow, equation 2.8 and 2.9.

Obtaining a solution for a turbulent flow is not an easy task. The finer
features the flow are unsteady and three-dimensionally random. There
are randomly swirling vortical structures within a turbulent flow called
turbulent Eddies. They are of various sizes and time scales and add to
the difficulties of the calculation [25]. Even though the complexity of a
turbulent flow is high, a meaningful, physical and increasingly accurate
solution can still be obtained. Such a solution requirers the total devo-
tion to understanding the setup. Knowledge of the mathematical and
physical ground of the specific problem is essential. The computational
time can be largely reduced by appropriate approximations and assump-
tions.

There are many different methods, of different difficulties and accuracy,


used to solve the PDEs. The direct numerical simulation (DNS) method
is one of those. The method solves the unsteady motions of the flow
directly for all scale sizes. The method gives a highly accurate solution.
However the size and time scale difference between large and small Ed-
dies make the DNS method computationally heavy. It requires fine grids
and a great deal of CPU time [11].

A simpler technique is to only resolve the large Eddies and model the
small Eddies, the method is called Large Eddy simulation (LES). The
model of the small Eddies basically assumes that they are isotropic (inde-
pendent of direction) and behave in a statistically similar manner. This

25
2.3. NUMERICS CHAPTER 2. THEORY

method is thus not as heavy as DNS and require less computational time.
The method is still time consuming [25].

Turbulence models
Modeling all the turbulence features of the flow with what is called turbu-
lence models is the most commercially used. Here mathematical models
of all the turbulent unsteady features, such as the mixing and diffusion
caused by the turbulent Eddies, are made. The calculation difficulty is
considerably reduces by this solving method, while the accuracy is held
high depending on the chosen model.

For CFD calculation the Navier-stokes equation is time, velocity and


pressure averaged via equation 2.22 and 2.23. The resulting equation is
the Reynolds-averaged Navier-stokes equation (RANS), given in tensor
form by equation 2.24 [11].

P = P̄ + p0 (2.22)
ui = Ūi + u0i (2.23)
δUi δUi 1 δP δ 2 Ui δτij
+ Uj =− +ν 2 + (2.24)
δt δxj ρ δxi δxi δxi
Where Ui and P are statistical averages, u0i and p0 are the fluctuations
for the velocity and the pressure and τij is the Reynolds stress tensor [11].

Equation 2.24 is then modeled to be solved. There are many turbu-


lence models in use today including one-equation, two-equation and
Reynolds stress models (RSM). The first mentioned is mostly validated
for aerodynamical flows and thus primarily used within that area. The
Reynolds stress models are some of the more complex turbulence models
in common use. They are based on modeling dynamic equations for the
Reynolds stress tensor. The improvement in solution accuracy, said to
be gained with RSM models, is still to be proven.

Two-equation models are the predominate turbulence models in use to-


day. This is because they offer a balance between numerical effort and
computational accuracy. The mainly used two-equation models are the
k −² and shear stress transport (SST) models. whereas the -˛² is the more
widely used. The SST model combines the use of two models, the k − ²
and the k − ω model. Both models are modeled relating the Reynolds

26
CHAPTER 2. THEORY 2.3. NUMERICS

stresses to the mean velocity gradients and the turbulent viscosity [11].
The k − ² models is shown to handel the free stream flow outside the
boundary layer with greater accuracy, but is less efficient in the bound-
ary layer region, where the Reynolds number is low. The model needs
finer grids near the wall to account for the flow formulation of the bound-
ary layer. The k − ω on the other hand gives a better resolution of the
boundary layer but is said to be less stable within the free-stream flow.
The model is shown to be sensitive to the inlet free-stream turbulence
properties [11]. The SST model combines, as mentioned, the two previ-
ously discussed. It uses the k − ω formulation in areas of low Reynolds
number flows (near walls) and the k − ² model further away from the
walls, in the free-stream. Although the method offers a higher accuracy
solution, it requires an enormous amount of CPU time and therefore
usually chosen after an initial solution by k − ², if required. The k − ²
is by far the most used model, it is shown to be very accurate for many
diffident applications.

2.3.2 Finite Element Method (FEM)


The finite element method is a discretization method used in many ap-
plications. The method accuracy is though most frequently validated for
structural mechanic calculations. FEM analysis is carried out by decom-
posing the whole physical domain into a finite number of subdomains,
finite elements, and calculating an approximate solution for each subdo-
main.

The finite elements come in different shapes and they possess various
properties. They are of different special dimensions one, two and three
dimensions. The elements are also of different complexity. Elements of
the same classification can be of different mathematical orders, linear or
of higher order. The choice of element directly impacts the accuracy and
the convergence of the solution.

For the finite element method, the elements do not overlap in space as
for some other discretization methods like the finite difference method.
This property is called disjoint support. Each element consists of one (in
the case of a one dimensional elements) or several distinguished points
called nodes. Each node serve a purpose. The classifications are geo-
metric nodes for nodes defining the element geometry and connection

27
2.3. NUMERICS CHAPTER 2. THEORY

nodes for nodes home for the defined degrees of freedom [10]. For many
studies the two definitions can define one and the same node. Figure 2.6
illustrates the shapes, degrees and node positions for some elements.

Figure 2.6: Figure illustrating the shapes, degrees and node positions for
some elements.

An approximation of the mathematical models describing the behav-


ior of the physical problem is represented by suitable functions in each
element and solved. The mathematical models of the physical setup are
usually a set of partial differential equations or integral equations. The
element equations are then assembled and a discrete solution for the sys-
tem of equations is then obtained, meaning the values for every node
is solved. That way approximate solution for the whole domain is ob-
tained [3]. The accuracy of the FEM solution refers to the mathematical
models ability to predict the behavior of interest of the real system. The
term convergence in FEM analysis sense refers to the accuracy of the
discretization of the mathematical model [22].

The general equation of the finite element method is given by equation

28
CHAPTER 2. THEORY 2.3. NUMERICS

2.25 [22], where an approximate solution for u is of interest.


n
X m
X
u≈ uj ψj + cj φj (2.25)
j=1 j=1

n is the number of nodes, uj is the values of u at each element node,


ψj stand for the interpolation functions, cj are the nodeless coefficients
and φj are the approximation derived by using concepts of interpolation
theory functions [22].

The solution given in equation 2.25 leads to the formation of the known
matrix expression 2.26 [22].

[K]{u} = {f } + {Q} (2.26)

Where [K] is the coefficient matrix and {f } is the source vector, they are
called the stiffness matrix and the force vector respectively in solid and
structural mechanics. {u} and {Q} are the primary and the secondary
element nodal degrees of freedom [22]. The expression describes the FEM
problem to be solved.

29
CHAPTER 3. MODELING

Chapter

Modeling

When creating a computer model, approximations are made. As men-


tioned in the theory chapter each simplification introduces an error that
need to be marginalized as much as possible.

The geometry used for the calculations was, as mentioned in previous


chapters, the new design runner. A design geometry (Pre production
model, the model prior to manufacture) was used for creating a CFD
and FEA model, figure 3.1 shows the design model. It should be noted
that the design model does not include the fillet at the blade-band and
crown T-joint section.

31
CHAPTER 3. MODELING

(a) (b)

(c) (d)

Figure 3.1: Figure of the designed runner model. (a) Showing a trig
view, (b) a side view,(c) a top view and (d) a bottom view.

32
CHAPTER 3. MODELING 3.1. CFD MODEL

Figure 3.2: Figure showing the CFD model used in the analysis. It
represents a third of the total geometry shown in figure 3.1.

3.1 CFD model


3.1.1 Geometry
The fluid domain is extracted from the design model. Geometrical sym-
metry allowed for a reduction of the model by two thirds, and thereby
gaining a reduction in computational time. The CFD model is shown
in figure 3.2. The model contains five blades, five splitter blades, eight
guide vanes and an extension of the outlet. The geometry was extracted
using the contour of the band and crown. The extension to the outlet
was added to avoid any effect it would have had on the flow near the
areas of interest. It should be noted that the extension is not a part of
the draft tube.

3.1.2 Preprocessing
The mesh used in the CFD calculation is a tetrahedra mesh with a size
of approximately 710, 000 nodes, se figure 3.3. The simulation setup
was a transient rotor stator simulation. The turbulence model chosen
for the simulation was the two equation k − ² model. The advection
scheme was high resolution with a second order backward Euler transient
scheme. The inlet boundary conditions were chosen to be of the type
total pressure, set to the pressure equal to the head hight. The flow
direction at the inlet was specified to flow in at an appropriate angle

33
3.2. FEM MODEL CHAPTER 3. MODELING

(a) (b)

Figure 3.3: Figure of the CFD Mesh. (a) Showing a top view, (b) a zoom
in, for a closer look at the mesh.

with respect to the angle of the stay vanes. The outlet conditions were
set to average static pressure 0. A time step of 0.5◦ was decidedly used
and a maximum of 10 iterations per loop was set. The torque of the
blade and splitter was monitored during the simulation. The monitoring
of the blades was done to enable a manual stop of the simulation once
the torque reached a steady periodicity.

3.2 FEM model


As mentioned earlier some approximation was made in order to keep the
computational time limited. The original sketch had complex band and
crown contours which needed to be smoothed or in some cases remade.
When changing the geometry, there is a constant struggle between sim-
plifying to gain a manageable simulation time and keeping the changes to
areas where they do not affect the results. Figure 3.4 shows an example
of a simplification made on the band geometry.

34
CHAPTER 3. MODELING 3.2. FEM MODEL

(a)

(b) (c)

Figure 3.4: Figure showing one of the simplifications on the design model
in order to save CPU time. (b) The zoom in design model contour on
the band, (c) a zoom in, for a look at simplification done at the area.

35
3.2. FEM MODEL CHAPTER 3. MODELING

Figure 3.5: Figure showing the finite element geometry of model one
used in the analysis. Containing a splitter blade, blade and part of the
crown and band.

3.2.1 Geometry
Two different FEA models were made. Model one is the simulation
model used in the analysis. Model two was used to investigate wether
the boundary conditions being close to the areas of interest had any
effect on the results, especially near the trailing edge of the blade where
the boundary conditions are closest to the blade.

Model one
The FEA model created had a portion of the band and crown, a blade
and a splitter blade, see figure 3.5. Since the runner consists of 15 blades
and splitters, model one was cut out of the total runner with an angle
of 360◦ /15 = 24◦ . The cut to the crown and band, as seen in figure 3.5,
follows the blades shape.

Model two
The model was made so that the cyclic symmetry boundary condition,
explained in section 3.2.3, was put further away from the areas of interest.

36
CHAPTER 3. MODELING 3.2. FEM MODEL

(a) (b)

Figure 3.6: Figure showing the finite element geometry of model two
used in the analysis. a) showing a back view and b) a front.

The geometry is shown in Figure 3.6. As can be seen the model has the
same shape as FEA model one, but here a splitter blade is added on one
side and a blade on the other.

3.2.2 Mesh

Model one and two were meshed with second order tetrahedra mesh ele-
ments with quadratic displacement behavior [1], se figure 3.7. They are
ten node elements with three DOF at each node, translation in x, y and z.

Second order surface elements with four to eight nodes were created,
figure 3.8. The elements were used to apply the pressure load. They
were created on the surfaces of the blade and splitter blade.

The mesh can be seen in figure 3.9. The mesh was made uniform with
one average element side length obtained from a mesh study. The mesh
study was preformed to minimize the numerical error in the simulation
model.

37
3.2. FEM MODEL CHAPTER 3. MODELING

Figure 3.7: Figure showing the type of solid mesh element used in the
analysis. It is a ten node tetrahedra element with quadratic displacement
behavior. The image is one gathered from the Ansys 11 manual pages
[1].

Figure 3.8: Figure showing the type of surface mesh element used in the
analysis. It is a Second order elements with four (on the right) to eight
(on the left) nodes. The image is one gathered from the Ansys 11 manual
pages [1].

Figure 3.9: Figure showing the finite element mesh of model two used in
the analysis. The figure is showing a pressure side view.

38
CHAPTER 3. MODELING 3.2. FEM MODEL

3.2.3 preprocessing
Model one

Since only a portion of the geometry was used certain conditions needed
to be filled. The finite element model can be seen as one part of a cou-
pled circular array. Therefore the degrees of freedom on both surfaces at
the band and crown need to be coupled, it is called cyclic symmetry.

Cyclic symmetry is essentially a coupled set of degrees of freedom, which


constrains the nodes in question to the same space of motion. The nodes
are only allowed to move in the same exact way for a certain coupled
DOF. In a coupled system of two nodes, one node is chosen to be the
master and the other in the coupled pair is the slave node. If the master
node is forced to move within the coupled DOF, the exact same motion
is applied to the slave node.

The nodes on each side of the crown and band portion were coupled
in all directions. The surface meshes on opposite side of the cut were
matched so that the sides had nodes at the same geometrical position
on the surfaces. That way cylindrical cyclic symmetry could be applied.

The bolts fixing the runner to the shaft were approximated as a fixed
line at the bolt circle. The nodes on the line were fixed in all DOF. At
each time step a new pressure load is applied to the surface of the blade
and splitter, see figure 3.10.

Model two

The constraints and boundary conditions were applied in the same way
here as in model one. The cutting angle here was 48◦ instead of 24◦ as
it was in model one.

The CFD simulation was modeled without the consideration of FEA


model two, since model two was an after construction. The boundaries
of the added splitter and blade were not defined in the CFD calculation.
Therefore the pressure load results for them could not be extracted with
the available methods. The blades were instead approximated to have
the same phase. It basically means that both blades and splitters were
at the same position with respect to the guide vanes.

39
3.2. FEM MODEL CHAPTER 3. MODELING

Figure 3.10: Figure showing the boundary conditions for the finite ele-
ment analysis.

The discretization of the pressure follows the time step size of the CFD
calculation. One hole cycle is as mentioned earlier 15◦ .

40
CHAPTER 4. RESULTS

Chapter

Results

4.1 CFD

The result presented in the present section are in no way thorough and
should not be expected to be that. The result are a mere presentation
of the quasi static pressure used in the FEM analysis of the runners me-
chanical response to those.

With that said, figure 4.1 shows the pressure drop as a point travels
along a line situated in the middle of the runner, between the band and
the crown. The pressure presented in the figure is the result of the CFD
analysis debrided in the modeling chapter.

The result of the analysis was a fluctuating dynamic pressure on the


blades that was later used for the finite element analysis, se figure 4.2.

41
4.1. CFD CHAPTER 4. RESULTS


(a)

(b)

Figure 4.1: Figure showing the results obtained by the CFD calculation.
The figure on the top, figure a, shows the pressure results normalized
with respect to the head of the power plant and figure b shows a contour
plot of the velocity at the same positions. The velocity is normalized
with respect to the the maximum velocity vmax .

42
CHAPTER 4. RESULTS 4.1. CFD

(a) The figure shows the pressure on the (b) The figure shows the pressure on the
pressure side of the blade. suction side of the blade.

(c) The figure shows the pressure on the (d) The figure shows the pressure on the
pressure side of the splitter blade. suction side of the splitter blade.

Figure 4.2: The figure is showing the pressure results on the surfaces of
the blades. It should be noted the the contour plot presented here is
taken at on certain time step out of the transient CFD calculation.

43
4.2. FEM CHAPTER 4. RESULTS

4.2 FEM

As predicted by earlier reports and publications[6, 4], the highest stress


values predicted by the finite element analysis were found near the trail-
ing edge of the blade. Examining figure 4.3, the prediction was matched
by the general look of the resulted contour plot of the maximum princi-
ple stress.

The high stresses shown to be located at the fixed line where the bolt
radius is located for the new runner are non real and caused by the lines
fixed degrees of freedom. But as seen in the figure the elevated stresses
at the area were only local and therefore assumed to not have an affect
on the areas of interest for the validation.

The FEA resulted in the quasi static stress fluctuations shown in fig-
ures 4.4 and 4.5, due to the guide vane runner interaction. The result
showed that the stress variations for P S200 and P S300 on the pressure
side were in sync, meaning they were in phase. figure 4.4 shows that
there was a small phase difference between P S15 and P S25. Figure 4.5
reveals a phase difference on the suction side between the results at SS15
and SS25.

The investigation of the cyclic boundary conditions closeness to the areas


of interest for the analysis described in the modeling chapter came to the
following conclusion. The analysis refereed to here was the FEA model
two built as described the mentioned chapter. The investigation came to
the conclusion that the cyclic conditions being close to the trailing edge
of the blade, due to the nature of the geometry section build up, did not
have a significant affect on the end results. The conclusion was based on
the mean stress and stress amplitude at all the points of interest.

44
CHAPTER 4. RESULTS 4.2. FEM

(a)

(b) The figure shows a zoom in for a (c) The figure shows a zoom in for a
closer look at the stress at the tailing closer look at the stress at the trailing
edge of the runner blade, near the band edge of the runner blade, near the band
on the pressure side. on the suction side.

Figure 4.3: Figure showing the results obtained by the FEA calculation.
The figure on the top, figure a, shows the maximum principle stress for
a given time step. The plotted contour is of σ/σmax . For a closer look
of the legend values se figures given in the appendix, figure A.3-A.5

45
4.2. FEM CHAPTER 4. RESULTS

Figure 4.4: The figure shows the quasi static stress variation on the
pressure side at different point, at various locations from the trailing
edge of the blade.

46
CHAPTER 4. RESULTS 4.2. FEM

Figure 4.5: The figure shows the quasi static stress variation on the
suction side at two different locations from the trailing edge of the blade.

47
4.2. FEM CHAPTER 4. RESULTS

Figure 4.6: A plot of the mean stress at different locations from the
trailing edge on the blade. The mean stress calculated from the site
measurements is here compared to the FEA calculated results.

4.2.1 Mean stress

Calculating the mean stress value for the validation points for the stress
variation shown in figure 4.4 and 4.5 gave the following results.

As seen in figure 4.6, the calculated mean stresses at the points P S200
and P S300 gave a better comparison with the measured. The exact devi-
ations there were only 9 and 1% respectively. The deviation was higher
for points closer to the trailing edge, where on the pressure side they
were of magnitudes over 30%, 32.7% at P S15 and 33.8% at P S25. The
deviation on the suction side was high for SS15 where it was calculated
to be 121%.The case was different for SS25 where it was only 9%. The
deviations for both sides of the blade are given as a plot in figure 4.7.
The figure shows a general decline in deviation as the distance from the
trailing edge increases.

48
CHAPTER 4. RESULTS 4.2. FEM

Figure 4.7: A figure showing plots of the mean stress deviation as a


function of the distance from the trailing edge.

49
4.2. FEM CHAPTER 4. RESULTS

Figure 4.8: A plot of the mean stress at different locations from the
trailing edge on the blade. The mean stress calculated from the site
measurements is here compared to the FEA calculated results.

4.2.2 Stress amplitudes

The results for the stress amplitudes at the guide vane passing frequency
150Hz are given in figure 4.8. The figure includes the FEA calculated
values and the measured. A comparison of the two showed more or less
the same tendencies as for the mean stress. The deviation seemed to
decrease as the distance from the trailing edge was increased if only the
deviation is observed. Looking closer at the figure 4.8 reveals that the
FEA results are consistently lower for all validation points as appose to
the mean stress where the calculated values were higher than the mea-
sured.

The amplitude deviation presented by figure 4.9 show a maximum am-


plitude deviation of 94.3% at the location SS15. The stress amplitude
for SS25 was 71.1% lower than the measured value at the same position.

50
CHAPTER 4. RESULTS 4.2. FEM

Figure 4.9: A figure showing plots of the stress amplitude deviation as a


function of the distance from the blade trailing edge.

The pressure side amplitudes were 67.9, 63.2, 32.1 and 13.3% lower than
the measured for P S15, P S25, P S200 and P S300 respectively.

51
CHAPTER 5. DISCUSSION AND
CONCLUSIONS

Chapter

5
Discussion and
Conclusions

5.1 CFD

Looking at the mean torque calculated by the CFD analysis the con-
clusion might be that the calculations were good, since the deviation
between the torque calculated by use of the nominal output of the unit
and the CFD mean torque was minimal (1.8%). A closer examination of
the flow near the guide vane walls revealed the following.

The boundary layer resolution was low and that had to do with the
fact that the mesh near the wall did not have the required fineness to
completely resolve the boundary layer. This is recommended by docu-
mentations and reports dealing with CFD flow analysis [1]. The investi-
gation of the boundary layer flow near the guide vane walls was examined
for one reason. That reason being the notion that the dynamic pressure
amplitude may be amplified by that size of the wake created at the guide
vane outlet. The size of the wake is dependent of the geometrical prop-
erties of the outlet [7], but it is also shown to be partially dependent on
the boundary layer thickness.

53
CHAPTER 5. DISCUSSION AND
5.2. FEM CONCLUSIONS

A deeper analysis of the CFD tool is needed. A mesh sensitivity study


would give some understanding of the effects and also a comparison be-
tween a structured and an unstructured mesh may contribute to a gain
in knowledge [25].

5.2 FEM
By only looking at the prediction of the mean and dynamic stress for
P S200 and P S300, a good correlation is observed for both. This fact
can lead to the conclusion that the FEA calculations gave a reasonably
accurate comparison. The conclusion of the validation is though not as
convincing for several reasons.

The mean stress calculated at the thinner areas of the blade, near the
trailing edge of the blade, were of higher deviation values then those in
the areas mentioned earlier in this section. The higher deviation, not
only for the mean stress but also for the dynamic stress amplitudes, may
have many different causes.

The FEA model was built out of the design model and geometrical differ-
ences, between the design model and the real life manufactured runner
can partially explain the deviation. The geometry used in the FEA did
not include the fillet present in the real runner due to the weld section
at the blade-band and crown T-joint transition, which ultimately could
have had an effect on the results.

Another factor for geometrical difference was that the manufacturing


manual allowed for a deficiency as high as 10% for the thickness of the
blade at the outlet. A ten percent increase in thickness at the trailing
edge of the blade could account for the higher mean stress results in the
FEA calculations. The mean stress was measured on the surface of the
blade. A 10% increase in thickness was calculated to translate to a 30%
decrease in stress, which would account for the higher mean stress found
by the finite element analysis.

The outlet shape of the blade also differed in profile. The outlet profile
of the real runner is more of a rounded shape, the difference is illustrated
in figure 5.1. The real geometry profile can be seen in figure 1.9. The
shape difference could explain the high deviation in the mean stress at

54
CHAPTER 5. DISCUSSION AND
CONCLUSIONS 5.2. FEM

Figure 5.1: The figure illustrates the geometry difference at the trailing
edge of the blade between the FEA model and the real runner. The
figure to the left represents the shape profile of the FEA model and the
one on the right is that of the real runner. It should be noted that the
figure is only an illustration of the difference and is in no way an exact
representation of the geometry.

SS15.

Other factors like hydraulic damping by the large body of water, un-
der which the runner operates, were not taken into account in the finite
element analysis of the thesis. Another important fact is the natural
frequency respond of the runner. The structures natural vibration could
have proven carousal for accurate prediction of the dynamic stress am-
plitudes.

The reason why these factors were not modeled was that the simula-
tions conducted in the thesis were used to try and understand the effect
of pressure fluctuation. The aim was to quantify by some means the
structural response of the runners to the pressure fluctuations due to
RSI. In the case of the geometrical differences, they were assumed to not
have an effect on the end results. The assumption helped to simplify the
model making it less CPU time consuming. One example is the fillet at
the blade - band/crown t-joint transition. Not modeling the fillet reduces
the geometry complexity at the area and the size of the mesh is therefor
smaller for such a model. The CPU time is in that way lessened. The
conclusions drawn in this section, by looking at the results, show though
that the statement regarding the geometrical assumption needed to be
further examined.

55
CHAPTER 5. DISCUSSION AND
5.3. FUTURE WORK CONCLUSIONS

5.3 Future work


Further work needs to be done for higher accuracy prediction of the dy-
namic stresses the runner is subjected to. The FEA model need further
work to more accurately represent the real geometry for further under-
standing of the effect of geometry deficiencies.

The technique Fluid Structural Interaction (FSI) is a method developed


during resent years and is being validated for various different appli-
cations. FSI is the study of the effect of fluid forces on a mechanical
structure and visa versa. The governing equations for the FEA are cou-
pled with the equations for the CFD calculations and the system is solved
simultaneously.

Their are two different classifications of the method. The first is one
way FSI, where the solution only goes one way. The fluid domain is
solved and all fluid forces are used for the the structural calculations, so
a solution is obtained. The other type is two way FSI. The fluid domain
is solved and the structural analysis is done in the same way as for one
way FSI, but here the structural solution is remeshed with the new de-
formed geometry and a new solution is obtain for the fluid domain. This
procedure is repeated till a satisfactory convergent solution i reached.

FSi has during the last years proven to be efficient and accurate, as
can be read in the report presented in 2007 by Christine Monette, An-
dré Coutu and Omprakash Velagandula, "Francis runner natural fre-
quency and mode shape prediction" , where highly accurate solutions
was reached using the method [20].

The damping effect of the large body of water and dynamic stress ampli-
fication factors need to be model and taken into account in the calcula-
tions, to accurately predict the dynamic stress amplitudes. The natural
frequency and mode shapes of the runner are known amplification fac-
tors in the interaction between the hydraulic and mechanical resonance.
These factor have proven to be important in these types of calculations
[20].

56
CHAPTER 5. DISCUSSION AND
CONCLUSIONS 5.3. FUTURE WORK

Bibliography

[1] Ansys 11 manual pages.

[2] Hydro power, part 3: Turbines. Rainpower Norway AS, documen-


tation.

[3] N.-E. W. Alf Samuelsson. Finita elementmetodens grunder. Stu-


dentlitteratur, 1988.

[4] H. Bjorndal. Tonstad, unit 1 and 2, turbine runner stress measure-


ments report. Norconsult report, 2004.

[5] J. s. Carl Nordling. Physics handbook, Elementary constants and


units, tables, formulas and diagrams and mathematical formulae.
Studentlitteratur, Bratt Institut fur Neues Lernen, Chartwell-Bratt,
1980.

[6] A. Coutu, H. Aunemo, B. Baddig, and O. Velagandula. Dynamic


assessment of hydraulic turbines. Technical report, GE Hydro, 2005.

[7] A. Coutu, D. Proulx, and S. Coulson. Dynamic assessment of hy-


draulic turbines. Technical report, GE Hydro, 2003.

57
CHAPTER 5. DISCUSSION AND
5.3. FUTURE WORK CONCLUSIONS

[8] A. Coutu, D. Proulx, S. Coulson, and A. Demers. Dynamic assess-


ment of hydraulic turbines. Technical report, GE Hydro, 2003.

[9] T. Dahlberg. Teknisk hållfastighetslära. Studentlitteratur, 2001.

[10] C. A. Felippa. Introduction to finite ele-


ment methods. University of Colorado, 2004.
http://www.colorado.edu/engineering/CAS/courses.d/IFEM.d/.

[11] J. H. Ferziger and M. Peric. Computational methods for fluid dy-


namics. Berlin : Springer, cop., 2002.

[12] H.-J. Hult. Fatigue design of hydraulic turbine runners. PhD thesis,
NTNU, 2005.

[13] J. Hult. Laster och brott. Almqvist & Wiksell Förlag AB, 1978.

[14] J. Hult. Spänningar och brott. Almqvist & Wiksell International,


1990.

[15] S. D. Knutsen. Dynamic blade load of high head francis and rpt.
Technical report, GE Hydro, 2007.

[16] G. Krivchenko. Hydraulic machines: Turbines and pumps. Lewis


publishers, 1994.

[17] B. W. Lennart Råde. Mathematics handbook, for science and engi-


neering. Studentlitteratur, 1998.

[18] E. Martinot. Renewables, global status report, 2006 update. Techni-


cal report, Renewable Energy Policy Network for the 21st Century,
2006. www.ren21.net.

[19] C. B. Moler. Numerical computing with MATLAB. Philadelphia:


Society for Industrial and Applied Mathematics, 2004.

[20] C. Monette, A. Coutu, and O. Velagandula. Francis runner natural


frequency and mode shape prediction. Technical report, GE Hydro,
2007.

[21] J. J. O’Connor and E. F. Robertson. Johann andreas von segner.


University of St Andrews, Scotland, 2004.

58
CHAPTER 5. DISCUSSION AND
CONCLUSIONS 5.3. FUTURE WORK

[22] J. N. Reddy. An introduction to the finite element method. McGrew-


Hill, 1993.

[23] E. Ullman. Materiallära. Liber AB, 2003.

[24] N.-E. Wiberg. Finita elementmetodens tillämpningar, inom olika


teknikområden. Studentlitteratur, 1980.

[25] J. M. C. Yunus A. Cengel. Fluid mechanics, Fundamentals and


application. McGraw-Hill, 2006.

59
CHAPTER 5. DISCUSSION AND
5.3. FUTURE WORK CONCLUSIONS

60
APPENDIX A. FIGURES

Appendix

Figures

A.1 Figure from site

61
A.1. FIGURE FROM SITE

62
APPENDIX A. FIGURES

Figure A.1: The figure shows the location of the strain gauges on blade 7. The Picture shows the locations of
the gauges on the pressure side. The image is a print from the site measurement report[4].
APPENDIX A. FIGURES

63
Figure A.2: The figure shows the location of the strain gauges on blade 7. The Picture shows the locations on
A.1. FIGURE FROM SITE

the suction side. The image is a print from the site measurement report[4].
A.2. FIGURES FROM THE FEA APPENDIX A. FIGURES

A.2 Figures from the FEA

64
APPENDIX A. FIGURES

65
Figure A.3: Figure showing the results obtained by the FEA calculation. The figure shows the maximum principle
stress for a given time step. The plotted contour is of σ/σmax .
A.2. FIGURES FROM THE FEA
A.2. FIGURES FROM THE FEA

66
Figure A.4: Figure showing the results obtained by the FEA calculation. The figure shows a zoom in for a better
picture of the trailing edge at the pressure side. The contour plot gives the distribution of the maximum principle
stresses for a given time step. The plotted contour is of σ/σmax .
APPENDIX A. FIGURES
APPENDIX A. FIGURES

67
Figure A.5: Figure showing the results obtained by the FEA calculation. The figure shows a zoom in for a better
picture of the trailing edge at the suction side. The contour plot gives the distribution of the maximum principle
stresses for a given time step. The plotted contour is of σ/σmax .
A.2. FIGURES FROM THE FEA
A.2. FIGURES FROM THE FEA APPENDIX A. FIGURES

Figure A.6: Figure showing the results obtained by the FEA calculation.
The contour plot gives the distribution of the maximum principle stresses
for a given time step. And the streamlines color code responds to the
local velocity of the fluid. The figure was only meant to serv as an
illustrative picture of the interaction between fluid and structure and
nothing else.

68
APPENDIX B. TABLES

Appendix

Tables

B.1 Site measurements

69
B.1. SITE MEASUREMENTS APPENDIX B. TABLES

Table B.1: The table presents the normalized mean stress measured. The
stresses are normalized with respect to the maximum measured mean
stress.
Servo SS15 SS25 PS15 PS25 PS200 PS300
(mm)
30 -0.060653 -0.10886 -0.19907 -0.22395 -0.33126 -0.30171
40 -0.066874 -0.13841 -0.073095 -0.11353 -0.25972 -0.24728
50 -0.068429 -0.14619 -0.048212 -0.079316 -0.24417 -0.24261
60 -0.065319 -0.1493 -0.010886 -0.032659 -0.20995 -0.21617
70 -0.059098 -0.14774 0.03888 0.010886 -0.17885 -0.18818
80 -0.057543 -0.15552 0.11042 0.066874 -0.15086 -0.14463
90 -0.055988 -0.16019 0.1633 0.10731 -0.13375 -0.1182
100 -0.054432 -0.16174 0.19285 0.13064 -0.12131 -0.10264
110 -0.054432 -0.16796 0.23173 0.16019 -0.10109 -0.087092
120 -0.052877 -0.17729 0.28149 0.19596 -0.07154 -0.07154
130 -0.051322 -0.17885 0.31726 0.22551 -0.048212 -0.060653
140 -0.051322 -0.18818 0.34837 0.25194 -0.031104 -0.048212
150 -0.051322 -0.19285 0.38569 0.28771 -0.0031104 -0.024883
160 -0.051322 -0.19596 0.41835 0.31726 0.020218 -0.0046656
170 -0.045101 -0.19285 0.41835 0.33437 0.043546 0.010886
180 -0.048212 -0.20373 0.47278 0.37481 0.07465 0.032659
190 -0.049767 -0.2084 0.51322 0.40435 0.10109 0.057543
200 -0.049767 -0.20995 0.54121 0.42457 0.1182 0.073095
210 -0.049767 -0.21462 0.58165 0.45257 0.14308 0.10264
220 -0.051322 -0.22084 0.62208 0.48834 0.16796 0.13219
230 -0.054432 -0.22551 0.66407 0.53344 0.19129 0.15708
240 -0.057543 -0.22706 0.71384 0.57854 0.2224 0.18507
250 -0.059098 -0.22706 0.76205 0.61275 0.24728 0.2084
260 -0.060653 -0.22706 0.81649 0.65163 0.27372 0.23328
270 -0.062208 -0.22706 0.87092 0.69518 0.29705 0.25661
280 -0.063764 -0.24106 0.93313 0.77605 0.3437 0.30949
290 -0.059098 -0.24728 1 0.86936 0.41369 0.37636

70
APPENDIX B. TABLES B.1. SITE MEASUREMENTS

Table B.2: The table presents the normalized stress amplitudes at the
guide vane passing frequency 150Hz. The amplitudes are normalized
with respect to the maximum measured stress amplitude.
Servo SS15 SS25 PS15 PS25 PS200 PS300
(mm)
30 0.02439 0.036585 0.12805 0.054878 0.054878 0.060976
40 0.036585 0.054878 0.14024 0.18293 0.097561 0.091463
50 0.036585 0.067073 0.14024 0.18293 0.10976 0.097561
60 0.042683 0.073171 0.18902 0.18902 0.11585 0.12195
70 0.054878 0.079268 0.30488 0.21951 0.12195 0.14634
80 0.060976 0.091463 0.42683 0.28049 0.11585 0.15854
90 0.073171 0.10366 0.4939 0.32927 0.091463 0.17073
100 0.079268 0.10976 0.52439 0.34756 0.091463 0.17683
110 0.091463 0.11585 0.57927 0.39024 0.15854 0.18293
120 0.10976 0.11585 0.64024 0.42683 0.2622 0.15854
130 0.12195 0.11585 0.68293 0.46341 0.31707 0.13415
140 0.13415 0.10976 0.71341 0.5061 0.34146 0.14024
150 0.15244 0.10976 0.81098 0.58537 0.36585 0.20732
160 0.16463 0.11585 0.87805 0.64024 0.36585 0.26829
170 0.14024 0.14634 0.78049 0.58537 0.31098 0.28659
180 0.15244 0.12195 0.86585 0.63415 0.31707 0.36585
190 0.15854 0.097561 0.89024 0.65244 0.33537 0.40244
200 0.16463 0.091463 0.93902 0.68902 0.35976 0.42683
210 0.17073 0.073171 0.95732 0.72561 0.37195 0.40854
220 0.17683 0.036585 0.93902 0.7439 0.37805 0.37195
230 0.17683 0.018293 0.93293 0.70122 0.37805 0.37195
240 0.18902 0.0060976 0.96341 0.68293 0.37195 0.37195
250 0.19512 0.0060976 1 0.67073 0.37195 0.37805
260 0.19512 0.012195 0.95732 0.68293 0.34756 0.36585
270 0.19512 0.0060976 0.93902 0.70732 0.31098 0.35976
280 0.17073 0.030488 0.90244 0.68902 0.28659 0.31707
290 0.14024 0.054878 0.83537 0.64634 0.34146 0.27439

71
B.2. FEA RESULTS APPENDIX B. TABLES

B.2 FEA results

72
APPENDIX B. TABLES B.2. FEA RESULTS

Table B.3: The table presents the normalized FEA principle stresses.
The stresses are normalized with respect to the maximum measured
mean stress on site. σ denotes the mean stress and ∆σ 2 stands for the
stress amplitude. It shoould be noted that the ∆σ
2 was normalized with
respect to the maximum measured stress amplitude.
Degree SS15 SS25 PS15 PS25 PS200 PS300
0.5 -0.12568 -0.19639 1.199 1.0428 0.38246 0.3118
1 -0.12475 -0.19734 1.2248 1.0662 0.40536 0.335
1.5 -0.12395 -0.19794 1.2509 1.0895 0.4271 0.35793
2 -0.1231 -0.19789 1.2781 1.1136 0.44732 0.37954
2.5 -0.12239 -0.19731 1.299 1.1318 0.46118 0.39451
3 -0.12192 -0.19633 1.3091 1.1403 0.46566 0.39938
3.5 -0.12182 -0.19521 1.3101 1.1405 0.46177 0.39502
4 -0.1219 -0.19395 1.3019 1.1326 0.45057 0.38285
4.5 -0.1223 -0.19279 1.2818 1.1143 0.43139 0.36266
5 -0.12277 -0.19174 1.2571 1.092 0.40974 0.3403
5.5 -0.12347 -0.19126 1.2302 1.068 0.38844 0.31887
6 -0.1246 -0.19163 1.2009 1.0422 0.36843 0.29952
6.5 -0.12547 -0.19242 1.1822 1.0261 0.35755 0.28896
7 -0.12581 -0.19353 1.1778 1.0228 0.35795 0.28876
7.5 -0.12606 -0.19512 1.1826 1.0277 0.36617 0.29611
8 -0.12568 -0.19639 1.199 1.0428 0.38246 0.3118
8.5 -0.12475 -0.19734 1.2248 1.0662 0.40536 0.335
9 -0.12395 -0.19794 1.2509 1.0895 0.4271 0.35793
9.5 -0.1231 -0.19789 1.2781 1.1136 0.44732 0.37954
10 -0.12239 -0.19731 1.299 1.1318 0.46118 0.39451
10.5 -0.12192 -0.19633 1.3091 1.1403 0.46566 0.39938
11 -0.12182 -0.19521 1.3101 1.1405 0.46177 0.39502
11.5 -0.1219 -0.19395 1.3019 1.1326 0.45057 0.38285
12 -0.1223 -0.19279 1.2818 1.1143 0.43139 0.36266
12.5 -0.12277 -0.19174 1.2571 1.092 0.40974 0.3403
13 -0.12347 -0.19126 1.2302 1.068 0.38844 0.31887
13.5 -0.1246 -0.19163 1.2009 1.0422 0.36843 0.29952
14 -0.12547 -0.19242 1.1822 1.0261 0.35755 0.28896
14.5 -0.12581 -0.19353 1.1778 1.0228 0.35795 0.28876
15 -0.12606 -0.19512 1.1826 1.0277 0.36617 0.29611
σ -0.13025 -0.22928 1.3267 1.1626 0.45092 0.37284
∆σ
2 0.0087521 0.015406 0.27611 0.24766 0.23192 0.23544
73
B.2. FEA RESULTS APPENDIX B. TABLES

74
APPENDIX C. MATLAB SOURCE CODE

Appendix

Matlab source code

C.1 Result presentation

%%% Translation of the finite element method analysis %%%%

%%%Reading the FEA stress results and stie measurements

S=dlmread(’RES_v4.txt’);
M=dlmread(’mean.txt’);
amp=dlmread(’amplitude.txt’);
s_max=max(M(:,4));
a_max=max(amp(:,4));
CFD_T=S(:,8)/(120*pi*160e6/375)*15;
FEA_T=S(:,7)/(120*pi*160e6/375)*15;

%%% calculating the principle stresses in the same direction


as the measured %%%
vinkel; % executing vinkel for calculating the deviating angles %
for i=1:6
S(:,i)=(S(:,i).*cos(delta_angle(i)*pi/180));

75
C.1. RESULT PRESENTATION
APPENDIX C. MATLAB SOURCE CODE

end

for i=1:1
S=[S;S];
end
x=0:(length(S)-1);

%%% calculating the mean stress and stress amplitude


Mean=ones(6,1);
MaxFluc=ones(6,1);
Amp=ones(6,1);

for i=1:6
Mean(i)=(mean(S(:,i)))/s_max;
MaxFluc(i)=max(S(:,i))-min(S(:,i));
Amp(i)=(MaxFluc(i)/2)/a_max;
end

%%% Normalization %%%


for i=1:6
S(:,i)=(S(:,i).*cos(delta_angle(i)*pi/180))/s_max;
j=i+1;
M(:,j)=M(:,j)/s_max;
amp(:,j)=amp(:,j)/a_max;
end
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

delta_mean=ones(6,1);
delta_amp=ones(6,1);
for i=1:6
j=i+1;
delta_mean(i)=abs((Mean(i)-M(27,j))/M(27,j))*100;
delta_amp(i)=abs((Amp(i)-amp(27,j))/amp(27,j))*100;
end

%%%%%%%%%%%%%%%%% Plot commands %%%%%%%%%%%%%%%%%


figure(1);
subplot(2,1,1),plot(x,S(:,1),’b’,’LineWidth’,2),grid on,hold;

76
APPENDIX C. MATLAB SOURCE C.1.
CODE RESULT PRESENTATION

legend(’SS15 FEA’,-1);
title(’Stress variation at suction side’);
subplot(2,1,2),plot(x,S(:,2),’r’,’LineWidth’,2),grid on;
xlabel(’Time step [degrees]’);
ylabel(’Relative stress’);
legend(’SS25 FEA’,-1);

figure(2);
subplot(4,1,1),plot(x,S(:,3),’g’,’LineWidth’,2),grid on;
legend(’PS15 FEA’,-1);
title(’Stress variation at pressure side’);
subplot(4,1,2),plot(x,S(:,4),’c’,’LineWidth’,2),grid on;
legend(’PS25 FEA’,-1);
subplot(4,1,3),plot(x,S(:,5),’m’,’LineWidth’,2),grid on;
legend(’PS200 FEA’,-1);
subplot(4,1,4),plot(x,S(:,6),’k’,’LineWidth’,2),grid on,hold;
legend(’PS300 FEA’,-1);
xlabel(’Time step [degrees]’);
ylabel(’Relative stress’);

figure(3);
subplot(2,1,1);
plot(M(:,1),M(:,2),’b’,’LineWidth’,2);
hold;
plot(290,Mean(1),’bo’);

plot(M(:,1),M(:,3),’r’,’LineWidth’,2);
plot(290,Mean(2),’ro’);

plot(M(:,1),M(:,4),’g’,’LineWidth’,2);
plot(290,Mean(3),’go’);

plot(M(:,1),M(:,5),’c’,’LineWidth’,2);
plot(290,Mean(4),’co’);

plot(M(:,1),M(:,6),’m’,’LineWidth’,2);
plot(290,Mean(5),’mo’);

plot(M(:,1),M(:,7),’k’,’LineWidth’,2);

77
C.1. RESULT PRESENTATION
APPENDIX C. MATLAB SOURCE CODE

plot(290,Mean(6),’ko’);
hold;
title(’Mean Stress’);
xlabel(’Servomotor stroke [mm]’);
ylabel(’Relative stress’);
legend(’SS15’,’SS15 FEA’,’SS25’,’SS25 FEA’,’PS15’,’PS15 FEA’,
’PS25’,’PS25 FEA’,’PS200’,’PS200 FEA’,’PS300’,’PS300 FEA’,-1);
GRID;

subplot(2,1,2);
plot(amp(:,1),amp(:,2),’b’,’LineWidth’,2);
hold;
plot(290,Amp(1),’bo’);

plot(amp(:,1),amp(:,3),’r’,’LineWidth’,2);
plot(290,Amp(2),’ro’);

plot(amp(:,1),amp(:,4),’g’,’LineWidth’,2);
plot(290,Amp(3),’go’);

plot(amp(:,1),amp(:,5),’c’,’LineWidth’,2);
plot(290,Amp(4),’co’);

plot(amp(:,1),amp(:,6),’m’,’LineWidth’,2);
plot(290,Amp(5),’mo’);

plot(amp(:,1),amp(:,7),’k’,’LineWidth’,2);
plot(290,Amp(6),’ko’);
hold;
title(’Stress amplitude at guide vane passing frequency (150Hz)’);
xlabel(’Servomotor stroke [mm]’);
ylabel(’Relative stress amplitude’);
legend(’SS15’,’SS15 FEA’,’SS25’,’SS25 FEA’,’PS15’,’PS15 FEA’,
’PS25’,’PS25 FEA’,’PS200’,’PS200 FEA’,’PS300’,’PS300 FEA’,-1);
GRID;

x_1=[15,25];
x_2=[15,25,200,300];

78
APPENDIX C. MATLAB
C.2.SOURCE
DEVIATION
CODEANGLE CALCULATIONS

figure(4);
subplot(2,1,1),plot(x_1,delta_mean(1:2),’r’,’LineWidth’,2),
grid on;
title(’Mean stress deviation between FEA and
site measurements’);
legend(’Suction side’,-1);
subplot(2,1,2),plot(x_2,delta_mean(3:6),’b’,’LineWidth’,2),
grid on;
legend(’Pressure side’,-1);
xlabel(’Distance from the blade trailing edge [mm]’);
ylabel(’Diviation [%]’);

figure(5);
subplot(2,1,1),plot(x_1,delta_amp(1:2),’r’,’LineWidth’,2),
grid on;
title(’Stress amplitude deviation between FEA
and site measurements’);
legend(’Suction side’,-1);
subplot(2,1,2),plot(x_2,delta_amp(3:6),’b’,’LineWidth’,2),
grid on;
legend(’Pressure side’,-1);
xlabel(’Distance from the blade trailing edge [mm]’);
ylabel(’Diviation []’);

x_3=0:14;
figure(6);
plot(x_3,CFD_T,’b’,’LineWidth’,2);
hold;
plot(x_3,FEA_T,’r’,’LineWidth’,2);

C.2 Deviation angle calculations


%%% Calculations of the deviating angle between site measurements
and FEA %%%

%%% input of the FEA priciple stresses %%%


sigma=[-9.561,7.808,0.11196;6.9943,-11.654,-1.4528;
13.277,72.724,13.484;18.107,63.464,10.737;
11.509,27.253,5.9124;11.164,20.653,6.2288];

79
C.2. DEVIATION ANGLE
APPENDIX
CALCULATIONS
C. MATLAB SOURCE CODE

Sigma=[14.065;6.9943;90.917;81.392;32.529;25.111];

%%% calculating the angle of the FEA principle stress%%%


global delta_angle;
theta=zeros(6,1);
phi=zeros(6,1);

for i=1:6
theta(i)=acos(sigma(i,3)/Sigma(i));
theta(i)=theta(i)*180/pi;
end

%%% calculating the angle difference between the labe and the
principle stress %%%
blade_norm_angle_z=[161.3360;160.1230;28.6518;
28.8518;34.428;37.2787];
blade_angle_z=abs(90-blade_norm_angle_z);
delta_angle=blade_angle_z-theta;

80

You might also like