You are on page 1of 152

Master Thesis of Science

Study of Design Load Cases for MultiMegawatt Onshore Vertical Axis Wind
Turbines
Christos Galinos

June 15, 2015

European Wind Energy Master - EWEM- Rotor design Track

Study of Design Load Cases for


Multi-Megawatt Onshore Vertical Axis Wind
Turbines

Master of Science Thesis

For obtaining the degree of Master of Science in Engineering Wind


Energy at Technical University of Denmark and in Aerospace
Engineering at Delft University of Technology.

Christos Galinos
June 15, 2015

European Wind Energy Master - EWEM


DUWIND - Delft University of Technology
RIS DTU - Technical University of Denmark
DNV GL - Classification Society

c Christos Galinos
Copyright
All rights reserved.

European Wind Energy Master - EWEM


Of
Rotor design Track

The undersigned hereby certify that they have read and recommend to the European
Wind Energy Master - EWEM for acceptance a thesis entitled Study of Design Load
Cases for Multi-Megawatt Onshore Vertical Axis Wind Turbines by Christos
Galinos in partial fulfillment of the requirements for the degree of Master of Science.

Dated: June 15, 2015

Supervisor:
Senior Scientist T.J. Larsen of DTU

Supervisor:
Prof. Dr. H.A. Madsen of DTU

Supervisor:
Assoc. Prof. Dr.ir. C.J. Simao Ferreira of TU Delft

Supervisor:
Dr. L. Vita of DNV GL

Reader:
Prof. Dr. G.J.W. van Bussel of TU Delft

Summary

In the last years, large scale Vertical Axis Wind Turbines (VAWTs) are re-gaining a
general interest in the wind energy sector as an alternative to Horizontal Axis Wind
Turbines (HAWTs). However, the existing standards and guidelines for wind turbine
certification do not include specific requirements for VAWT rotors.
This thesis aims to study the adequacy and applicability of the IEC 61400-1 ed.3, 2005
International Standard minimum design requirements [28], when it is applied on multimegawatt VAWTs. This is accomplished by simulating a VAWT aeroelastic model under
various Design Load Cases (DLCs) and analysing the results. The motivation comes from
the recent years scientific research and developments which revealed that a VAWT can be
beneficial in some types of applications in comparison with a HAWT [13, 74]. As a result,
a minimum engineering integrity proof for VAWTs is needed before the installation and
operation phase as in HAWT. The work is split into two main phases.
The first copes with the construction of a reliable large scale VAWT aeroelastic model
which serves as the basis for the analysis (chapter 2). A 5 MW 2-bladed VAWT aeroelastic
model is built based on the DeepWind concept [63, 80]. Since blade instabilities are
present [80, p. 21] the Sandia 34 m, 500 kW VAWT test bed [3] is up-scaled and used as
guidance for the modification of the blade structural properties. In addition, the selection
and validation of an aeroelastic code able to compute the VAWT aerodynamics and loads
under different environmental conditions and wind turbine states is investigated. The
study has been conducted through a special project by the author [20] as part of this
thesis. It was meant to study and apply an aeroelastic code by constructing a VAWT
aeroelastic model. The Sandia 34 m, 500 kW VAWT was modelled and simulated with
HAWC2 aeroelastic code [39], showing good agreement with experimental data published
by Sandia National Laboratories.
In the second part is studied the adequacy and applicability of the aforementioned international standard on VAWTs and recommendations are made by simulating several
DLCs using HAWC2 code (chapter 3). Ultimate and fatigue loads during normal operation of the wind turbine are extracted and sensitivity analyses of relevant parameters that
could influence the results are performed. These include the blade stiffness and structural

vi

Summary

damping, and the generator slip. Various wind conditions are applied including extreme
operating gust, extreme wind shear and extreme wind direction change. The results from
normal power production are compared with the NREL HAWT [35] which has the same
rated power as the VAWT model in order to give an insight on the load levels between
horizontal and vertical axis turbines. Design situations of parked turbine and emergency
shut down are also investigated.
The results (chapters 3 and 4) indicate that the loads emerging from the extreme wind
shear and wind direction change conditions are not critical for the VAWT, thus the corresponding design load cases could become optional. The case of parked turbine with
locked rotor at specific orientations relatively to the wind direction revealed blade instabilities when the turbine was simulated under 50 year recurrence period wind conditions.
The load comparison between the horizontal and the vertical axis wind turbines showed
that the loads at blade low root and turbine base bottom are higher in VAWT but it
should be noted that the NREL turbine is a 3 bladed configuration, while the developed
VAWT model has 2 blades. The case of supported tower with guy wires was analysed
and relevant load cases are proposed. In general regarding the loads it is concluded that
the inherent variation of the blade aerodynamic forces during every revolution results to
highly deterministic loads on the turbine reducing the effect of turbulence. This could
lead to lower partial safety factors if could be demonstrated in other VAWT types. Finally, some small modifications on the definitions of the standard are proposed in order
to avoid ambiguities when it is applied to VAWTs.

Acknowledgements

This work has been carried out in partial fulfilment of the Erasmus Mundus European
Wind Energy Master program. During my studies I met many people who inspired
and help me along these two years in Denmark and The Netherlands. I am pleased to
acknowledge everyone at the four participating Universities responsible for the creation
and support of this Master program, particularly Carlos J. Simao Ferreira from TU Delft.
I would like to thank my supervisors Torben J. Larsen and Helge A. Madsen from
Ris DTU, Luca Vita from DNV-GL and Carlos J. Simao Ferreira from TU Delft for
the constructive comments and triggering me with thoughts, which helped me to accomplish this study, especially Torben for the guidance through all the steps of the master
thesis.
I would also like to thank people at DNV-GL who provided me a pleasant environment every time I was studying there, which I truly enjoyed, among them Ole Kjr,
Erik R. Jrgensen, Gireesh K. Ramachandran and Johan Olaison.
Special thanks go to Uwe S. Paulsen who gave me the opportunity to meet the scientific
community of Wind Energy Division at Ris DTU, where I had the possibility to select
this interesting topic.
I wish to thank the secretary coordinators of the European Wind Energy Master in Delft
Linda and Zara for organizing the EWEM summer school at the end of the 2nd semester
bringing us, students and instructors closer, and meeting people from other universities
and industries related to wind energy. Last but not least, a big thank to my family for
their encouragement and support to my studies.

Roskilde, Denmark
June 15, 2015

Christos Galinos

vii

viii

Acknowledgements

Everything you can imagine is real.


By Pablo Picasso

Contents

Summary

Acknowledgements

vii

List of Figures

xvii

List of Tables

xx

Nomenclature

xxi

1 Introduction

1.1

Vertical Axis Wind Turbine Technology and Status . . . . . . . . . . . . .

1.2

Motivation of Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3

Thesis Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 Analysis Tools and VAWT Model

2.1

Numerical Aero-servo-elastic Codes . . . . . . . . . . . . . . . . . . . . . .

10

2.2

Vertical Axis Wind Turbine Model . . . . . . . . . . . . . . . . . . . . . .

11

2.2.1
2.2.2

Up-scaling the Sandia 500 kW VAWT and Comparison with the


DeepWind Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . .

15

Modification of the 5 MW DeepWind Rotor . . . . . . . . . . . . .

18

Blades and Tower

. . . . . . . . . . . . . . . . . . . . . . . . . . .

19

Guy Wires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

28

Generator model . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

Implementation in HAWC2 . . . . . . . . . . . . . . . . . . . . . .

31

ix

Contents

3 Design Load Cases


3.1
3.2

35

General Considerations on the Definitions of the Standard . . . . . . . . .


Normal Power Production . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Ultimate Load Analysis . . . . . . . . . . . . . . . . . . . . . . . .

36
38
42

Influence of Turbulence . . . . . . . . . . . . . . . . . . . . . . . .
Influence of Wind Shear . . . . . . . . . . . . . . . . . . . . . . . .
Generator Slip Sensitivity Analysis . . . . . . . . . . . . . . . . . .

48
49
50

Fatigue Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51

3.2.2

3.3
3.4
3.5

Influence of Turbulence . . . . . .
Influence of Wind Shear . . . . . .
Ultimate Loads under Extreme turbulence
Extreme Wind Shear Case . . . . . . . . .
Extreme Operating Gust Case . . . . . .

.
.
.
.
.

52
53
54
55
57

3.6

Extreme Wind Direction Change Case . . . . . . . . . . . . . . . . . . . .

62

3.7

Emergency Shut Down . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

64

3.8

Parked Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.8.1 Parked Idling Rotor - Ultimate Load Analysis . . . . . . . . . . . .

68
68

3.8.2

Parked Idling Rotor - Fatigue Analysis . . . . . . . . . . . . . . . .

69

3.8.3

Parked Standing Still Rotor - Ultimate Load Analysis . . . . . . .

71

Influence of Blade Structural Properties . . . . . . . . . . . . . . .

78

Guy wires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

79

3.9

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

4 Discussion of DLC Results and Recommendations

83

5 Conclusions and Recommendations for Further Research


5.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Recommendations for Further Research . . . . . . . . . . . . . . . . . . .

87
87
88

References

91

A Modules of the Aeroelastic Code HAWC2


A.1 Aerodynamics Module . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

97
97

A.1.1 2D Airfoil Aerodynamics

. . . . . . . . . . . . . . . . . . . . . . . 100

A.1.2 Aerodrag module . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103


A.2 Wind Module . . . . . . . . . . . . .
A.2.1 Tower shadow module . . . .
A.3 Structural Module . . . . . . . . . .
A.4 Generator . . . . . . . . . . . . . . .
A.5 Controller . . . . . . . . . . . . . . .
A.6 Hydrodynamics-Sea and Soil module

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

103
104
104
106
107
107

A.7 Dynamic Wires module . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

Contents

B VAWT Model-Complementary Data

xi

109

B.1 Blade Structural Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 109


B.2 Loads on the Modified DeepWind VAWT . . . . . . . . . . . . . . . . . . 111
C Airfoil Polars

113

D Complementary Results from the DLCs Analysis

115

D.1 Normal Power Production . . . . . . . . . . . . . . . . . . . . . . . . . . . 115


D.2 Extreme Operating Gust Case . . . . . . . . . . . . . . . . . . . . . . . . 118
D.3 Extreme Wind Direction Change Case . . . . . . . . . . . . . . . . . . . . 119
D.4 Parked Standing Still Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . 120
E NREL 5 MW HAWT Properties

123

xii

Contents

List of Figures

1.1

Vertical axis wind mills used by Persians at 7th century AD [1] . . . . . .

1.2

Savonius and Darrieus rotors [24] . . . . . . . . . . . . . . . . . . . . . . .

1.3

Different lift driven VAWT configurations [74] . . . . . . . . . . . . . . . .

1.4

Onshore Darrieus VAWT schematic diagram [74] . . . . . . . . . . . . . .

1.5

The Sandia 34 m Test-Bed, a 500 kW Darrieus VAWT [56] . . . . . . .

1.6

Vestas 9 kW bi-blade and EOLE 4 MW Darrieus VAWTs [22] . . . . . . .

1.7

VAWT-850 H-rotor [52] and FloWind Darrieus type turbine [22] . . . . .

1.8

DeepWind 5 MW [63, 65] and Vertiwind 2 MW [54] VAWT concepts. . .

2.1

Schematic diagram of aero-servo-elasticity [18] . . . . . . . . . . . . . . . .

10

2.2
2.3
2.4
2.5
2.6

Definition of global and blade coordinate systems . . . . . . . . . . . . . .


Definition of local coordinate systems . . . . . . . . . . . . . . . . . . . .
Time series of blade edgewise position at equator station . . . . . . . . . .
DeepWind and Sandia scaled blade center-lines layout . . . . . . . . . . .
Original and modified DeepWind and Sandia scaled blade mass per length
distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Original and modified DeepWind and Sandia scaled blade flatwise and
edgewise stiffness distribution . . . . . . . . . . . . . . . . . . . . . . . . .
Original and modified DeepWind and Sandia scaled blade torsional and
longitudinal stiffness distribution . . . . . . . . . . . . . . . . . . . . . . .
Original DeepWind blade cross-section layout for NACA 0018 and NACA 0025
profiles [80] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

13
13
14
17

2.7
2.8
2.9

19
20
20
20

2.10 Modified DeepWind blade cross-section layout for NACA 0018 and NACA 0025
profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.11 Centrifugal and gravitational forces acting to a rotating Darrieus VAWT
blade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.12 Flatwise moment and stress on the blade due to gravity and centrifugal
loads at rated rotor speed . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

xiii

xiv

List of Figures

2.13 Comparison of flatwise moment and stress on the blade due to gravity with
the turbine standing-still . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23

2.14 Original and modified tower mass per length and bending stiffness distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.15 Original and modified tower torsional and longitudinal stiffness distributions 25
2.16 Campbell diagram of the developed VAWT . . . . . . . . . . . . . . . . .

27

2.17 Campbell diagram of the developed VAWT with guy wires . . . . . . . . .

30

2.18 Static torque slip curve of induction generator [38] . . . . . . . . . . . . .

31

2.19 Schematic of the 2-bladed vertical axis wind turbine . . . . . . . . . . . .

33

3.1

Rotor power for vertical and horizontal axis wind turbines . . . . . . . . .

39

3.2
3.3

Time series of VAWT rotor power and torque at 13.5 m/s under turbulent
wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Time series of VAWT rotor power and torque at 13.5 m/s for steady wind

40
40

3.4

Rotor torque and speed for vertical and horizontal axis wind turbines . .

41

3.5

Turbine base bottom bending moment and blade root bending moment for
vertical and horizontal axis wind turbines . . . . . . . . . . . . . . . . . .
Time series of angle of attack at 13.5 m/s wind speed for the blade section
at equator height . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

42

3.7

Mean and STD of blade upper and low root bending moments for VAWT

42

3.8

Extrapolated 50 year return period extreme values of turbine base bottom


bending moment for vertical and horizontal axis wind turbines . . . . . .

43

3.6

Extrapolated 50 year return period extreme values of blade root bending


moment for vertical and horizontal axis wind turbines . . . . . . . . . . .
3.10 Support struts and strengthened blade root solutions for stress reduction
of the VAWT blade roots [22, 56] . . . . . . . . . . . . . . . . . . . . . . .

41

3.9

44
44

3.11 VAWT maximum base bottom moments from simulations with and without
turbulence and 50 year extrapolated values . . . . . . . . . . . . . . . . .

45

3.12 VAWT maximum blade low root bending moments from simulations with
and without turbulence and 50 year extrapolated values . . . . . . . . . .

45

3.13 Time series of VAWT base bottom bending moments . . . . . . . . . . . .

46

3.14 Time series of VAWT blade low root bending moments . . . . . . . . . . .

47

3.15 Turbine base bottom bending moment Mx and My spectra at 13.5 m/s

47

3.16 Blade low root flatwise and edgewise bending moments spectra at 13.5 m/s 47
3.17 Extrapolated 50 year return period extreme values of the VAWT base bottom bending moments for different turbulence levels . . . . . . . . . . . .

48

3.18 Extrapolated 50 year return period extreme values of the VAWT blade low
root bending moments for different turbulence levels . . . . . . . . . . . .

48

3.19 Extrapolated 50 year return period extreme values of the VAWT base bottom bending moments for different wind shear exponents . . . . . . . . .

49

3.20 Extrapolated 50 year return period extreme values of the VAWT blade low
root bending moments for different wind shear exponents . . . . . . . . .

49

3.21 Maximum base bottom and blade low root bending moments for different
generator slip values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

List of Figures

xv

3.22 Equivalent 1 Hz fatigue base bottom bending moments for vertical and
horizontal axis wind turbines . . . . . . . . . . . . . . . . . . . . . . . . .
3.23 Equivalent 1 Hz fatigue blade root bending moments for vertical and horizontal axis wind turbines . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.24 VAWT equivalent 1 Hz fatigue base bottom bending moments for different
turbulence levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.25 VAWT equivalent 1 Hz fatigue blade root bending moments for different
turbulence levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.26 VAWT equivalent 1 Hz fatigue base bottom bending moments for different
shear exponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

52

3.27 VAWT equivalent 1 Hz fatigue blade low root bending moments for different shear exponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

54

3.28 Comparison of ultimate base bottom bending moments for DLC 1.3 . . .

55

3.29 Comparison of ultimate blade low root bending moments for DLC 1.3 . .

55

3.30 Ultimate loads emerging from vertical and horizontal EWS for the turbine
base bottom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.31 Ultimate loads emerging from vertical and horizontal EWS for the blade
low root . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.32 Time series of VAWT base bottom bending moments during EOG at rated
wind speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51
52

53
53

56
57
59

3.33 Time series of VAWT blade-1 low root bending moments during EOG at
rated wind speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

60

3.34 Time series of VAWT rotor speed and torque during EOG at rated wind
speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

61

3.35 Ultimate loads emerging from positive and negative EDC for the turbine
base bottom bending moment . . . . . . . . . . . . . . . . . . . . . . . . .

63

3.36 Ultimate loads emerging from positive and negative EDC for the blade low
root bending moment . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

63

3.37 Time series of the turbine base bottom bending moments during emergency
shut down at 220 sec time stamp . . . . . . . . . . . . . . . . . . . . . . .

65

3.38 Time series of the blade low root bending moments during emergency shut
down at 220 sec time stamp . . . . . . . . . . . . . . . . . . . . . . . . . .

66

3.39 Maximum turbine base bottom bending moments for different brake time
constants and brake torques . . . . . . . . . . . . . . . . . . . . . . . . . .

67

3.40 Maximum blade low root bending moments for different brake time constants and maximum brake torques . . . . . . . . . . . . . . . . . . . . . .

67

3.41 VAWT equivalent 1 Hz fatigue base bottom bending moments for different
idling rotor speeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

70

3.42 VAWT equivalent 1 Hz fatigue blade root bending moments for different
idling rotor speeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

70

3.43 Rotor cross-section at equator height, and orientation of the blades at =

71

3.44 Maximum and STD of the turbine bottom loads with the turbine parked
in standing still mode under extreme winds . . . . . . . . . . . . . . . . .

72

3.45 Maximum and STD of blades low root moments with the turbine parked
in standing still mode under extreme winds . . . . . . . . . . . . . . . . .

73

3.46 Mean and STD of blades low roots bending moments with the turbine
parked standing still under extreme winds . . . . . . . . . . . . . . . . . .

73

xvi

List of Figures

3.47 Maximum and STD of the blade low root loads with rigid turbine except
the blades in parked-standing still mode . . . . . . . . . . . . . . . . . . .

74

3.48 Time series of blade-1 station position at equator height with the turbine
in parked-standing still mode under 50 year recurrence period extreme winds 75
3.49 Motion of blade-1 station position at equator height in parked standing
still mode under extreme winds . . . . . . . . . . . . . . . . . . . . . . . .
3.50 Position of the blade-1 station at equator height, angle of attack and lift
coefficient for one period of instability . . . . . . . . . . . . . . . . . . . .
3.51 Maximum and STD of the turbine bottom and blade root loads with in
standing still mode under 1 year recurrence period extreme winds . . . . .

76
76
77

3.52 Sensitivity analysis of the blade stiffness on blade low root bending moment 78
3.53 Sensitivity analysis of blade structural damping influence on blade one low
root maximum bending moment . . . . . . . . . . . . . . . . . . . . . . .

79

3.54 Turbine base bottom bending moment and blade low root loads for the
VAWT with and without guy wires . . . . . . . . . . . . . . . . . . . . . .

80

3.55 Mean and STD of the turbine base bottom bending moments between the
VAWT with guy wires and the HAWT . . . . . . . . . . . . . . . . . . . .

80

4.1

Comparison of the absolute maximum turbine base bottom and blade low
root bending moments for each DLC . . . . . . . . . . . . . . . . . . . . .

84

A.1 Discretisation of the rotor area into a polar grid in HAWC2, [83] . . . . .

98

A.2 The actuator cylinder flow model [45] . . . . . . . . . . . . . . . . . . . . 100


A.3 Sandias 500 kW VAWT blade relative velocity fluctuation at equator over
one revolution [20] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
A.4 Sandias 500 kW VAWT blade angle of attack variation at equator over
one revolution [20] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
A.5 Sandias 500 kW VAWT rotor power with and without the Stig ye dynamic stall model implementation [20] . . . . . . . . . . . . . . . . . . . . 102
A.6 Sandias 500 kW VAWT blade lift coefficient at equator over one revolution
with Stig ye dynamic stall model implementation [20] . . . . . . . . . . 103
A.7 Flow curvature effects on an airfoil in curved flow [40] . . . . . . . . . . . 103
A.8 Turbulent wind field representations in HAWC2 [83]. . . . . . . . . . . . . 104
A.9 Main body formulation in HAWC2, consisting by bodies and elements [83] 105
A.10 Main bodies in HAWC2, consisting by sub-bodies with Timoshenko beam
elements [83] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
A.11 The blade formulation, consisting by sub-bodies with Timoshenko beam
elements [83] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
B.1 DeepWind and Sandia scaled blade chord and mass per length distributions 109
B.2 DeepWind and Sandia scaled blade flatwise and edgewise stiffness distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
B.3 DeepWind and Sandia scaled blade torsional and longitudinal stiffness distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
B.4 Contours of blade-1 low root maximum flatwise bending moment for the
modified DeepWind turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 111

List of Figures

xvii

B.5 Contours of blade-1 low root maximum edgewise bending moment for the
modified DeepWind turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 111
B.6 Contours of turbine base bottom maximum Mx bending moment for the
modified DeepWind turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 112
B.7 Contours of turbine base bottom maximum My bending moment for the
modified DeepWind turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 112
C.1 Airfoil lift and drag coefficients of NACA 0018 and NACA 0025 [80] . . . 113
C.2 Airfoil lift and drag coefficients of NACA 0018 and NACA 0025 from = 0
to 40 [80] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
C.3 Airfoil moment coefficient of NACA 0018 and NACA 0025 as a function of
angle of attack at Re=1.2x107 [80] . . . . . . . . . . . . . . . . . . . . . 114
D.1 Mean rotor thrust for vertical and horizontal axis wind turbines . . . . . . 115
D.2 Time series of VAWT base bottom bending moments . . . . . . . . . . . . 116
D.3 Time series of VAWT blade low root bending moments . . . . . . . . . . . 116
D.4 Equivalent 1 Hz fatigue blade upper and low root bending moments as a
function of wind speed for the VAWT . . . . . . . . . . . . . . . . . . . . 117
D.5 Equivalent 1 Hz fatigue base bottom bending moments for vertical and
horizontal wind turbines with and without turbulence . . . . . . . . . . . 117
D.6 Equivalent 1 Hz fatigue base bottom bending moments for vertical and
horizontal wind turbines with and without turbulence . . . . . . . . . . . 117
D.7 Time series of VAWT base bottom bending moments during the EOG at
19.5 m/s wind speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
D.8 Time series of VAWT blade-1 low root bending moments during the EOG
at 19.5 m/s wind speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
D.9 Ultimate loads emerging from positive and negative EDC . . . . . . . . . 119
D.10 Maximum and STD of the braking torque with the turbine parked in standing still mode and exerted to 50 year recurrence period extreme winds . . 120
D.11 Maximum and STD of the turbine bottom bending moments with the turbine parked in standing still mode exerted to 50 year recurrence period
extreme winds (Dynamic stall model is deactivated) . . . . . . . . . . . . 120
D.12 Maximum and STD of blade one and two low roots bending moments with
the turbine parked in standing still mode exerted to 50 year recurrence
period extreme winds (Dynamic stall model is deactivated) . . . . . . . . 121
D.13 Maximum and STD of the turbine bottom and blade root loads with the
turbine parked in standing still mode exerted to 1 year recurrence period
extreme winds (Dynamic stall model is deactivated) . . . . . . . . . . . . 121
E.1 NREL 5 MW, 3-bladed horizontal axis wind turbine schematic . . . . . . 124
E.2 Definition of coordinate system for the HAWT [83] . . . . . . . . . . . . . 124

xviii

List of Figures

List of Tables

2.1

Operational and geometric data of Sandia and DeepWind VAWT rotors. .

12

2.2

DeepWind blade basic properties [80]. . . . . . . . . . . . . . . . . . . . .

12

2.3
2.4

Scale dependencies of rotor properties applying aerodynamic and geometric


similarities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Natural frequencies for the DeepWind and Sandia scaled rotors. . . . . . .

17
18

2.5

Natural frequencies and structural damping of the modified blade. . . . .

23

2.6

Scale dependencies of the tower structural properties.

. . . . . . . . . . .

24

2.7

Geometric properties of the wind turbine tower. . . . . . . . . . . . . . . .

26

2.8

Natural frequencies and structural damping for the DeepWind modified


rotor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Wire properties of Sandia 34 m VAWT [72]. . . . . . . . . . . . . . . . . .

28
28

2.10 Scale dependencies of the guy wire structural properties. . . . . . . . . . .

29

2.11 Natural frequencies and structural damping for the DeepWind modified
rotor with guy wires. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

29

2.12 Operational and geometric data of the modified DeepWind Rotor which is
used on the remaining of this study and will refer as VAWT. . . . . . . .

32

3.1

Wind turbine classes [28]

. . . . . . . . . . . . . . . . . . . . . . . . . . .

36

3.2

Site turbulence characteristics [28] . . . . . . . . . . . . . . . . . . . . . .

36

3.3

Design load cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

37

3.4

Maximum wind turbine base bottom bending moments. Comparison of


loads emerging from gust (EOG) and normal power production at different
wind speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

59

Maximum wind turbine blade-1 low root bending moments. Comparison of


loads emerging from gust (EOG) and normal power production at different
wind speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

60

Loads and deflections of wind turbines with idling rotor under 50 year
recurrence period wind conditions. . . . . . . . . . . . . . . . . . . . . . .

69

2.9

3.5

3.6

xix

xx

List of Tables

3.7
3.8

Loads and deflections of wind turbines with idling rotor under 1 year recurrence period wind conditions. . . . . . . . . . . . . . . . . . . . . . . .

69

Loads and deflections of wind turbines with idling rotor under 50 year
recurrence period wind conditions. . . . . . . . . . . . . . . . . . . . . . .

77

E.1 Operational and geometric data of the NREL 5 MW horizontal axis wind
turbine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

Nomenclature

Latin Symbols
[m2 ]

Area

Wind shear exponent

[]

Cd

Drag coefficient

[]

Cl

Lift coefficient

[]

Cm

Moment coefficient

[]

Extreme turbulence model constant

ca

Blade chord

[m]

Rotor diameter

[m]

Youngs modulus

Fb

Braking force

[N ]

Fc

Centrifugal force

[N ]

Fg

Gravity force

[N ]

FN

Blade cross section internal normal force

[N ]

Fk

Brake clamping force

[N ]

Fp

Guy wire pretension force

[N ]

Fn

Aerodynamic blade force normal to the chord

[N ]

Ft

Aerodynamic blade force tangent to the chord

[N ]

Shear modulus

Rotor height

Area moment of inertia

[m4 ]

Ix

Area moment of inertia around x-axis

[m4 ]

[m/s]

[N/m2 ]

[N/m2 ]
[m]

xxi

xxii

Nomenclature

Iref

Expected value of hub height turbulence intensity at a 10 min average wind


speed of 15 m/s
[]

VAWT rotor mass moment of inertia

Guy wire length

Generator torque at HSS

[N m]

Mn

Generator torque at rated power expressed at LSS

[N m]

Mx

Flatwise blade cross section internal bending moment

[N m]

My

Edgewise blade cross section internal bending moment

[N m]

Mz

Torsional blade cross section internal bending moment

[N m]

W
ohler exponent

[]

Number of blades

[]

Per revolution frequency

Pnmech

Nominal rotor power

Pr

Propeller mode

[]

Qn

Blade normal force distributed on actuator cylinder surface

[N ]

Qt

Blade tangential force distributed on actuator cylinder surface

[N ]

Rotor radius

[m]

Re

Reynolds number

[]

Distance from an axis

[m]

rb

Mechanical disc brake radius

[m]

Generator slip

[]

sf

Scaling factor

[]

Characteristic times of gust, extreme wind shear, extreme wind direction


change
[sec]

Ta

Aerodynamic torque

Time

Tb

Braking torque

[N m]

Tj

Inertia torque

[N m]

ta

Airfoil thickness

tc

Brake time constant

tt

Tower thickness

Wind velocity

Volume

V1

Extreme wind speed averaged over 10 min with 1 year recurrence period[m/s]

V50

Extreme wind speed averaged over 10 min with 50 year recurrence period[m/s]

Vave

Annual average wind speed at hub height

Ve1

Expected extreme wind speed averaged over 3 sec with 1 year recurrence
period
[m/s]

Vgust

Largest gust magnitude with an expected return period of 50 years

[kgm2 ]
[m]

[Hz]
[W ]

[N m]
[sec]

[m]
[sec]
[m]
[m/s]
[m3 ]

[m/s]

[m/s]

Nomenclature

xxiii

Vhub

Hub height wind velocity

[m/s]

Vref

Reference wind speed average over 10 min

[m/s]

Vr

Rated wind speed

[m/s]

zhub

Wind turbine hub height

[m]

Greek Symbols

Angle of attack

[deg]

Parameter of the extreme wind shear model

[]

Logarithmic damping decrement

[]

Damping ratio

[]

Azimuth angle

[deg]

Tip speed ratio

Turbulence scale parameter

[]
[m]
[m2 /s]

Kinematic viscosity of air

Density

[kg/m3 ]

Solidity

[]

Longitudinal turbulence standard deviation

Normal stress due to normal force

[P a]

Normal stress due to flatwise bending moment

[P a]

Inflow angle

[deg]

Rotor speed

[rad/s]

Nominal generator speed

[rad/s]

Actual generator speed

[rad/s]

Synchronous generator speed

[rad/s]

Subscripts
DW

DeepWind model

hub

Hub height

Sand

Sandia model

sc

Scaled value

Superscripts

[m/s]

xxiv

sc

Nomenclature

Scaled value

Abbreviations
AC

Actuator Cylinder model

ATEFlap

Active Trailing Edge Flap model

BEM

Blade Element Momentum model

BE

Blade Edgewise

BM

Bending Moment

CFD

Computational Fluid Dynamics

DLC

Design Load Case

DLL

Dynamic Link Library

DMST

Double Multiple Stream Tube model

DNV

Det Norske Veritas

DTU

Technical University of Denmark

ECN

Energy Research Centre of the Netherlands

EDC

Extreme Direction Change Model

EOG

Extreme Operating Gust

ETM

Extreme Turbulence Model

EWS

Extreme Wind Shear Model

FA

Flatwise Anti-symmetric

FS

Flatwise Symmetric

GL

Germanischer Lloyd

GM

Global Maxima

HAWT

Horizontal Axis Wind Turbine

HSS

High Speed Shaft

IEA

International Energy Agency

IEC

International Electrotechnical Community

LSS

Low Speed Shaft

MDAO

Multi-disciplinary Design, Analysis and Optimization

MST

Multiple Stream Tube model

NREL

National Renewable Energy Laboratory

NTM

Normal Turbulence Model

NWP

Normal Wind Profile model

PSD

Power Spectra Density

SNL

Sandia National Laboratories

SST

Single Stream Tube model

STD

Standard Deviation

Nomenclature

TI

Tower In-plane

TO

Tower Out-of-plane

TU Delft

Delft University of Technology

VAWT

Vertical Axis Wind Turbine

xxv

xxvi

Nomenclature

Chapter 1
Introduction

Wind turbines play a major role in the limitation of electrical energy production from fossil
fuels towards a sustainable world. Nowadays, horizontal axis wind turbines (HAWTs)
have been developed and installed all over the world converting a fraction of the kinetic
energy of the wind to electricity. Other than the HAWT configuration there is another
large category, the vertical axis wind turbines (VAWTs). In the latter, the rotor axis is
oriented perpendicular to the wind direction, while HAWT rotors are parallel to the wind
direction.

1.1

Vertical Axis Wind Turbine Technology and Status

VAWTs can be classified into two categories, lift and drag driven machines. This concept
has been used from the ancient years to convert wind energy to mechanical power (vertical
axis wind mills) for flour grinding and water pumping (figure 1.1 on the following page).
The wind mills exploit the aerodynamic drag which is created in the rotor to produce
torque (drag driven devices). This configuration is used today in some applications like
the cup anemometers or small turbines known as Savonius rotors (figure 1.2 on the
next page). They produce high torque at low wind speeds and they have low efficiency.
In addition, the quantity of the material used in the rotor is considerably higher in
comparison with lift driven machines, thus is not an option for construction of large
scale wind turbines.
The French engineer Georges Jean Marie Darrieus in 1925 came up with a design of
VAWT where the rotor torque is produced by the aerodynamic lift created on the blades
(Darrieus wind turbine or eggbeater). The blades in this concept have a Troposkien shape
(figure 1.2 on the following page). The power coefficient can reach high values similar with
the ones achieved in HAWTs. Many alterations of lift driven devices have been conceived
as it can be seen in figure 1.3 on page 3 with the most common after the Darrieus, the
H-rotor proposed by Peter Musgrove in the 70s.

Introduction

In figure 1.4 on the facing page a typical schematic of Darrieus turbine is shown. The main
parts of the turbine are the blades with the tower (rotating column) comprising the rotor
and the gearbox with the generator located at the bottom of the structure. Depending of
the design, guy wires can connect the tower top with the ground for stabilisation purposes.
Sandia National Laboratories (SNL) in the 70s started intensive research on VAWT
mainly focusing in the Darrieus configuration constructing small and medium size prototypes with 5, 17 and 34 m equator diameter. The 34 m Test-Bed [3] a 500 kW
experimental VAWT was successfully in operation for more than ten years (figure 1.5 on
page 4). At that time similar studies were conducted by the National Research Council of
Canada and in Denmark at Ris National Laboratory for Sustainable Energy and Aalborg
University.
This research around the world led to the construction of prototypes and commercial
products by private companies in the 80s and 90s. In Denmark Vestas built a 9 kW
Darrieus prototype turbine with two double blades for increased stiffness (figure 1.6 left
on page 5). In 1984 at Canada was erected a 4 MW prototype called Eole [67] the largest
VAWT ever built but after failure of one of the large bearings its operation terminated
(figure 1.6 right on page 5). In the UK designs were focused on straight blade VAWTs
or H-rotor [52], utilizing the shorter blades and simpler rotor design, for instance the
VAWT-850 a 500 kW turbine (figure 1.7 left on page 5), built by VAWT Ltd. FloWind

Figure 1.1: Vertical axis wind mills used by Persians at 7th century AD, [1].

Figure 1.2: Savonius and Darrieus rotors, left and right respectively, [24, p. 66].

1.1 Vertical Axis Wind Turbine Technology and Status

Figure 1.3: Different lift driven VAWT configurations: Darrieus (a),H-rotor (b),V-rotor
(c), Delta-rotor (d), Diamond-rotor (e) and the Giromill (f ), [74].

Figure 1.4: Onshore Darrieus VAWT schematic diagram, [74].

a company from United States commercialized two Darrieus wind turbine models, one of
142 kW and the other 250 kW rated power in the mid 80s (figure 1.7 right on page 5) and
produced hundreds of them with a total capacity of 95 MW at the end of 1985. Similar

Introduction

Figure 1.5: The Sandia 34 m Test-Bed, a 500 kW Darrieus VAWT, [56].

size VAWTs were produced by other companies such as the Adecon 150 kW [29], the
DAF-Indal 50 kW [61, p. 40] and the VawtPower with 200 kW rated power [74].
After a successful course the first years of wide installations commercial VAWTs started
to fail costing too much for maintenance. The main failure originated from the blades.
Due to the rotor rotation the blades exposed to highly cyclic loading, more than 108
cycles for 20 years design lifetime, causing fatigue in the blade parts and in the blade to
tower connections. The problem was partially originated due to the aluminium material
that most of the manufacturers were using to construct the blades, but it is known its
moderate fatigue characteristics especially when parts are constructed through extrusion
method [74]. For this reason, in combination with the fact that the blades of a Darrieus
VAWT have to be curved, which was usually done during installation introducing extra
stresses, deteriorated the lifetime of the blades. Another factor which accelerates the blade
fatigue is the turbulence of the wind, where the wide low frequency spectrum excites the
blades causing small deformations and as a result add to the fatigue. On the other hand,
the brake system was also reported critical and prone to failure for one widely installed
VAWT. The FloWind turbines braking system was the most costly component of the
turbine regarding maintenance [74]. Concurrently, the developments in the counterpart
HAWTs and the possibility to produce cheaper energy led the VAWT industry to abandon
the production of medium and large scale VAWTs around 1995.
Darrieus configuration was preferred in the multi-kilo watt VAWT prototypes and commercial products as was described above, because the curved blade shape minimises the
flatwise bending stresses. The ideal shape called troposkien and it is the shape that
a rope takes when is rotated along an axis with a defined speed under the effect of centrifugal and gravitational forces. In this case the rope is loaded only in tension. This is
preferred in VAWT designs and is also beneficial when fibre reinforced composite materials are used, because they have superior strength along the fibre direction in comparison
with the transverse direction.

1.1 Vertical Axis Wind Turbine Technology and Status

Figure 1.6: Vestas 9 kW bi-blade (left) and EOLE 4 MW (right), both Darrieus VAWTs, [22].

Figure 1.7: VAWT-850 a 500 kW H-rotor VAWT in UK in the 90s (left), [52]. FloWind
Darrieus type turbines installed at Tehachapi Pass, California in 90s (right), [22].

Nowadays, new scientific research is ongoing in many universities, among them the TU
Delft in The Netherlands and DTU in Denmark, and a new generation of VAWT designs
emerge such as the DeepWind EU project [63, 65] a 5 MW VAWT concept (figure 1.8
left on page 6) and the Nenuphars Vertiwind a 2 MW VAWT [54] (figure 1.8 right on
page 6). This shows that VAWTs could be advantageous in some cases for wind energy
exploitation especially in offshore deep water installation places.
Some facts in terms of advantages and disadvantages between VAWTs and HAWTs are
given below.
VAWT can capture the wind from different directions without the need of yaw
mechanism in contrast to HAWT.
Complex and heavy equipment such as the gearbox, the generator and the inverter

Introduction

Figure 1.8: DeepWind 5 MW concept [63, 65] and a small scale prototype installed at
Roskilde-fjord in Denmark (left) and Vertiwind 2 MW turbine with a 35 kW prototype installed
in France [54], both VAWTs.

are located to the bottom of the turbine and essentially lay in the ground without
burden the rotor with extra weight. Thus, the maintenance of these parts is also
easier due to their location. In HAWT these components are located on the tower
top.
The rotor is simpler for a VAWT because does not have pitch mechanism in contrast
to a contemporary HAWT which has. On the other hand, if cyclic pitch is utilised
in a VAWT could increase both power performance and optimum design tip speed
ratio and result in lower drive-train costs as stated in [32, p. 221].
The aerodynamic noise produced in VAWTs can be less as the higher relative blade
velocities are smaller compared to the velocities of the blade tips in HAWTs.
VAWT studies [15, 16] have shown that a skewed inflow can increase the energy
captured by the turbine as the downwind part of the rotor is exposed both to the
wake and the new incoming wind. In the case of floating VAWTs the motion
of the platform creates small relative velocities between the wind and the rotor in
all directions that seen by the blades as skewed inflow, thus increasing the energy
extraction.
A drawback related to the mean wind is that in HAWT the rotor center is usually
located higher compared to the equivalent height in a VAWT configuration. So,
there is a small benefit in the energy yield for a HAWT of the same rated power as
it is exposed to higher mean wind speeds.
A main disadvantage of Darrieus design is the blade length which is around double
compared to an equivalent size HAWT. This could be alleviated from the simpler
blade design without twist and the commonly used one or two airfoil profiles with
constant or step chord length along the span. In addition, the VAWT blades are
attached from both ends, while HAWT blades are cantilevered from the hub.
Finally, the cyclic aerodynamic loading due to blade rotation and crossing of the
wind leads to higher fatigue in VAWT blades and should be taken into account during design in comparison with HAWTs, which have constant aerodynamic loading
for a given velocity and small fluctuations due to turbulence. However, VAWTs
loading due to gravity and wind shear is steady.

1.2 Motivation of Thesis

1.2

Motivation of Thesis

As was presented above, VAWTs have advantages of usage in comparison with HAWTs
which make them desirable in some applications. This is the reason why new concepts and
prototypes for VAWTs both in academia and industry level start to emerge. This means
that there will be a need of proper certification before companies could introduce to the
market multi-megawatt VAWTs, as in the case of HAWTs. The certification is based on
standards and for wind turbines the International Electrotechnical Community IEC-61400
series has become acceptable worldwide by the wind energy sector. These series cover a
wide range of aspects of wind turbines, from the requirements for site assessment to proper
measurement of electrical properties, and the minimum design requirements regarding
their structural integrity. The latter is covered by the series 1 (IEC 61400-1 ed.3:2005,
[28]) in the case of large scale onshore wind turbines. In that document different external
conditions together with various operating states of the turbine are defined creating the so
called design load cases (DLCs). Even though they refer in general on wind turbines, they
were developed taking implicitly into account the concept of HAWT and the experience
gained from realized wind projects with them. This means that in principle could be
applicable to VAWT, as well. On the other hand, inherent differences of VAWTs from
HAWTs, as summarized in the previous section, may lead to other factors in the case
of VAWTs that could be critical, and changes in the current DLCs or additions to the
existing ones may be necessary. Furthermore, given definitions in the document about a
wind turbine could need to be altered or introduce new, accordingly.
The interest of the author on VAWTs together with the current uncertainty in the adequacy and applicability of the IEC 61400-1 Ed.3 DLCs for the structural integrity of
multi-megawatt VAWTs was the motivation for conducting this study. The uncertainty
refers to the lack of relevant studies regarding DLCs for VAWTs. In addition, the existence of a standard that covers the VAWT certification could help the new generation of
VAWT concepts to be realised.

1.3

Thesis Objectives

In this work the adequacy and applicability of the IEC 61400-1 standard is studied.
In particular the wind turbine minimum design requirements as described in the latest
edition of series 1, and recommendations are for the case of onshore VAWT rotors. This
is performed by the construction of a VAWT aeroelastic model and the simulation of
several design load cases described in the standard with an aeroelastic code. It should
be noted that the analysis is focused in onshore VAWTs even though it is more likely as
described above, to see sooner applications with multi-megawatt VAWTs for offshore wind
exploitation. This is done because as in the case of HAWTs, first the turbine is tested on
land applying the IEC 61400-1 ed.3 standard and if this phase is succeeded then analysis
is conducted according to offshore wind turbine requirements [31]. The objectives of the
study are summarized below:
1. Construction of a VAWT aeroelastic model. This model consists the base where the
standard with the various load cases are applied. More specifically:

Introduction

Have an onshore VAWT model with details of aerodynamics, structural and


generator properties able to be simulated in an available aeroelastic code.
The turbine to be of a similar rated power with a conventional horizontal axis
wind turbine in order to do relevant comparisons.
2. Investigation of the standard for its applicability and adequacy in the case of VAWT
rotor, by the selection of the external conditions and wind turbine operation modes
(DLCs) through numerous simulations and sensitivity analyses. Finally, draw conclusions and recommendations upon the definitions of the standard and the DLCs
by analysing the results from the simulations.

Chapter 2
Analysis Tools and VAWT Model

Utility scale wind turbines used for electricity production are complex large machines
which need to be designed and simulated thoroughly before costly prototypes are built
and commercialisation take place. High fidelity numerical tools and corresponding wind
turbine models are therefore developed. This is done in order to assure safe operation
during the design lifetime but also to minimise the cost of energy produced from the wind
exploitation.
Concerning the mechanical parts, large quantities of steel and composite materials are
required which are expensive and decisive of the final price of the turbine. Especially
the blades and the tower of the turbine are constructed from large amounts of these
materials, thus are designed to have the least amount of weight. This leads to flexible
structures where the forces acting to the turbine create deformations and in turn change
the behaviour of the turbine structurally and aerodynamically. In other words there is an
interaction between the wind (air) and the structure known as fluid-structure interaction.
On top of that the coupling of the generator and the controller of the turbine influences the
response of the structure. The combined result called aero-servo-elasticity and requires
concurrent simulations of all the interactions. A schematic of it is given in figure 2.1 on
the following page (upper part of the hexahedron) as proposed by Friedmann [18]. Inertial
forces are composed of the gravity and the centrifugal forces. Aerodynamics include the
lift and drag that acts to the structure through the wind interaction and elastic forces
caused due to the elastic deformations of the structure itself. Thermal effects are not so
important and usually are not included in aero-servo-elastic simulations.

10

Analysis Tools and VAWT Model

Figure 2.1: Schematic diagram of aero-servo-elasticity (upper part of the hexahedron), [18].

2.1

Numerical Aero-servo-elastic Codes

Different theories and models that can capture the aerodynamics and structural dynamics
are combined in wind turbine simulations. These numerical analyses are carried out with
sophisticated codes and can predict the loads and other parameters taking into account
the turbine state and external conditions. Many universities, research institutes and
companies have developed such tools. Some state of the art aeroelastic codes according
to relevant literature [47, 62] are given below:
Phatas and TURBU from Energy Research Centre of the Netherlands (ECN) [43].
HAWC2 and Flex5 from Technical University of Denmark (DTU) - Ris National
Laboratory for Sustainable Energy [39].
Fast from National Renewable Energy Laboratory of US (NREL) [34].
Bladed from Classification Society DNV-GL [9, 10].
OWENS from Sandia National Laboratories (SNL) [57].
BHawC from Siemens Wind Power.
For this study HAWC2 code is selected as the most suitable numerical tool because it is
available to the DTU students and researchers as an in-house program. In addition, direct
contact with the scientific personnel who develop the HAWC2 tool (one of the universities
where the author carries out his study) in the case of implementation problems is a clear
advantage. The same code was used in the DeepWind VAWT project [63, 82], and is also
validated concerning the part of the aerodynamic prediction of VAWTs specifically, which
is important for correct load predictions [17].
Furthermore, HAWC2 was used to model and study the structural and aerodynamic
response of the Sandia 34 m test-bed, a 500 kW Darrieus VAWT by the author [20] where

2.2 Vertical Axis Wind Turbine Model

11

the results compared with the published measured data [2, 72, 79]. At that work a good
agreement was found in the natural frequencies, the aerodynamic performance and the
loads between the aeroelastic model and the measurements from SNL. This means that
not only it is possible to model a VAWT in HAWC2 but also the simulated response of
the turbine will be close to a real one.
It should be mentioned that the faculty of Aerospace Engineering of TU Delft also develops state of the art modules for analysing VAWTs such as the unsteady two-dimensional
vorticity model (U2DiVA) [14] concerning the aerodynamics. But due to the need of coupling it with a solver for the structural dynamics interaction and the high computational
requirements, this was considered less suitable for this study.

2.2

Vertical Axis Wind Turbine Model

In the previous section the available simulation tools were reviewed and the most appropriate was chosen for this study. Next arise the question of what configuration and size
VAWT should be used as a baseline for this analysis.
Regarding the size of the wind turbine in terms of rated power; nowadays a lot of interest
is focused on utility scale applications such as wind farms. VAWTs have potential as an
alternative to multi-megawatt HAWTs on cost reduction of electricity. This leads to the
conclusion that a few MW rated power VAWT would be representative and shed light
on the weak and strong points of the turbine in combination with the application of the
relevant guidelines.
Large scale VAWT designs and prototypes built in the last decades with rated power at
least 0.5 MW are few. The SNL 500 kW test-bed, the VAWT-850 a 500 kW H-rotor,
the Eole 4 MW and the 5 MW DeepWind Darrieus VAWTs are found in this range.
Structural and aerodynamic details for the Eole and VAWT-850 turbines do not exist
in the open literature. On the other hand, Sandia National Laboratories has published
many papers from their VAWT research program and as was mentioned earlier, the 34 m
test-bed has already been modelled and studied by the author. In addition, the recent
Europeans Union DeepWind project for the design and optimisation of a VAWT for
offshore installations is also documented in detail in the literature [63, 64, 80, 82] and
belongs in the desired range of power output. However, the DeepWind rotor is reported
to have large blade edgewise deflections [80, p. 21] which make it unsuitable and unrealistic
to use in the analysis without modifications.
Taking into account the available sources with their limitations it is decided to construct
an aeroelastic model of 5 MW rated electrical power based on the 5 MW DeepWind rotor
and the Sandia 500 kW, 34 m test-bed. On the remaining of this study when Sandia
VAWT is mentioned will always refer to the Sandia 500 kW, 34 m test-bed VAWT.
The DeepWind model is a floating turbine thus, only the rotor (blades and tower) will be
used and a base will be added in order to simulate it as an onshore turbine. The main
properties of the 5 MW DeepWind [80] and Sandia 500 kW [4] rotors are summarized
at table 2.1 on the next page. The blades of the DeepWind rotor are composed by
two different profiles the NACA 0018 and NACA 0025 while the Sandia blades utilise
the NACA 0021 and the SNLA 0018/50 airfoils. In table 2.2 on the following page the

12

Analysis Tools and VAWT Model

DeepWind basic blade properties are given. The airfoil polars of the DeepWind blades
are used as given in the DeepWind report [80, p. 11] and can be also found in appendix
C on page 113 for completeness.
Table 2.1: Operational and geometric data of Sandia and DeepWind VAWT rotors.

Property
Rated electrical power
Rated rotational speed
Rated wind speed
Cut-in wind speed
Cut-out wind speed
Number of blades, N
Rotor diameter, D
Rotor height to diameter ratio
Blade chord, ca
Solidity, = NRca
Swept area

DeepWind

Sandia

5
5.95
0.62
14
4
25
2
121
1.18
5
0.165
11996

0.5
37.50
3.93
12.5
2
34
1.25
0.91/1.07/1.22
0.107
955

[MW]
[rpm]
[rad/s]
[m/s]
[m/s]
[m/s]
[-]
[m]
[-]
[m]
[-]
[m2 ]

Next, the rotor model from the DeepWind project is copied and a low shaft with a rigid
base are added. The low shaft initially is assumed rigid in order to evaluate the response
of the rotor and especially the blades. The upper and lower roots of the blades are fixed
to the tower both in translation and rotation.
Table 2.2: DeepWind blade basic properties [80].

Property
Blade length
Chord length
Airfoil

Value
Low root
29.44
5
NACA 0025

Intermediate
149.29
5
NACA 0018

Upper root
21.70
5
NACA 0025

[m]
[m]
[-]

Three types of coordinate systems are used throughout this study to express the results.
The global (figure 2.2 on the facing page) which is fixed to rotation and translation and
is located at the turbine base bottom, the blade 1 and 2 coordinate systems located at
blade 1 and 2 low roots respectively and local coordinate systems. Each blade coordinate
system (figure 2.2 on the next page) is attached to the blade and as a result, when
the rotor rotates it also rotates. Each blade coordinate system origin is located to the
blade low root cross section elastic center. In figure 2.2 the lines illustrate the center-line
of the turbine blades, the tower, the shaft and the base. It is also depicted the blade
equator station where the distance from the tower is the largest (60.5 m). The wind
direction is towards yglobal and the turbine base bottom point coincides with the origin of
global coordinate system. The local coordinate systems are used to express the loads on
the blades (figure 2.3 on the facing page). In the figure a blade part is illustrated (blade
center-line) together with the corresponding structural elements and nodes. Each element
has a local coordinate system with its z-axis pointing from node i to node i+1. The origin

2.2 Vertical Axis Wind Turbine Model

13

of each local system is found at the respective blade cross section elastic center, where
x-axis is aligned with the first elastic axis and y-axis with the second (see figure 2.10 on
page 21 for elastic axes orientation). In this study when blade loads are given are always
expressed to local coordinate system. For instance, the blade low root moments Mx ,
My and Mz correspond to the in-plane (flatwise), out of plane (edgewise) and torsional
moments respectively. Similarly, the blade coordinate systems are only used to express
relative position of blade sections, and never used to express blade loads.

Figure 2.2: Definition of coordinate systems. The global coordinate system is fixed to the
ground, while blade-1 and 2 coordinate systems are located at blade-1 and blade-2 lower roots
respectively and rotate with the blades. Positive wind direction is towards yglobal .

Figure 2.3: Definition of local coordinate system for an element. Each element has a local
coordinate system with its z-axis pointing from node i to node i+1.

Initially, basic aeroelastic simulations were performed applying a turbulent wind speed
ramp from 4 m/s to 15 m/s without wind shear at various rotor speeds (0.62, 0.60, 0.58
and 0.56 rad/s) and the response was evaluated. From the results of various signals was
found that the blades have large oscillations around their equilibrium position. These

14

Analysis Tools and VAWT Model

were seen for all the different rotor speeds even below 13 m/s wind speed where the
mean rotor power is below rated. The range of the relative edgewise (out-of-plane, xdirection on blade coordinate system) blade motion of equator station exceeds the 7 m
at 13 m/s winds for every revolution of the rotor and is not acceptable. Not only was
considered unrealistic but also the ultimate and fatigue damage due to this motion would
be tremendous. In figure 2.4 are plotted the time series of blade-1 edgewise position at
equator station expressed in blade-1 coordinates for 0.62 and 0.58 rad/s rotor speeds (If
the blade was rigid the deflection would be zero).The blade edgewise motion is large with
similar range for the different rotor speeds. The range of the relative blade motion in
the y and z-directions is less than 3 m. The simulations were also performed with rigid
tower in order to focus on the blade response and exclude any possible coupling between
the blade and the tower modes. The results showed that the large blade oscillations
are still present. This indicated that the blade stiffness is relatively low and should be
modified. Similar study in ([80, p. 23]) emphasizes the large blade edgewise deflections
during different operating conditions.

Position x [m]

15
0.62 rad/s Rotor speed

10
5
0
-5
-10
-15
200

400

600

800

1000

1200

1400

Position x [m]

15
0.58 rad/s Rotor speed

10
5
0
-5
-10
-15
200

400

600

400

600

800

1000

1200

1400

800

1000

1200

1400

Wind Speed [m/s]

15

10

0
200

time [sec]

Figure 2.4: Time series of blade edgewise (out-of-plane) position at equator station expressed
in blade coordinate system (upper and middle). The simulations are performed applying a wind
speed ramp (down).

The Sandia 0.5 MW VAWT aeroelastic model will be up-scaled and used as a reference.
This will give an insight of how to modify the DeepWind rotor which will form the baseline
for the investigation of various design load cases. Thus, a case study research strategy will
be followed where one aeroelastic numerical model will be developed taking into account

2.2 Vertical Axis Wind Turbine Model

15

relevant data from the literature. Then, thorough numerical simulations of the model
and interpretation of the results will lead to recommendations on the DLCs of the current
IEC standard.

2.2.1

Up-scaling the Sandia 500 kW VAWT and Comparison with the


DeepWind Rotor

In this section are described the laws which are used to up-scale the Sandia VAWT. The
up-scaling is based on geometric and aerodynamic similarities [70]. This implies that
the tip speed (R) and the solidity (equation 2.1) between the original and the scaled
turbines are equal. The blade profiles and the number of blades are kept the same, thus
the flow geometry is similar regarding the rotor at equivalent operating points, neglecting
second-order aerodynamic effects. As a result, the tip speed ratio (equation 2.2) is
maintained.

N ca
R

(2.1)

R
u

(2.2)

Where N is the number of blades, ca is the blade chord length, R the turbine radius,
the rotor speed and u the wind speed.
The selected parameter for up-scaling is the rotor swept area and the scaling factor sf is
denoted as the square root of the swept areas ratio (equation 2.3). This means that the
lengths will be linearly proportional to the scaling factor ( sf 1 ).
r
sf =

ADW
ASand

(2.3)

Where ADW and ASand are the DeepWind and Sandia rotor swept areas respectively.
The resultant scaling factor is equal to 3.54 [-] and leads to a wind turbine with similar
rated power as the DeepWind rotor (5 MW).
Keeping the same rotor height to diameter ratio (H/D) it is implied that the scaled
rotor height Hsc and diameter Dsc are increased proportionally with the scaling factor
(equations 2.4 and 2.5).

Hsc = H sf 1

(2.4)

(2.5)

Rsc = R sf

As the tip speed and solidity are kept the same, the scaled chord csc
a and rotor speed sc
are given by equations 2.6 and 2.7 below.

16

Analysis Tools and VAWT Model

1
csc
a = ca f s

(2.6)

(2.7)

sc = sf

The airfoil profiles are preserved, thus the airfoil chord to thickness ratio ca /ta (relative
thickness) does not change after scaling, and because the chord is sf 1 the same holds
for the blade thickness (equation 2.8).
1
tsc
a = ta sf

(2.8)

It should be mentioned that the Reynolds number (Re) is higher in the up-scaled turbine
as the chord length is increased. However, this effect is small for Re numbers which
already (equation 2.9) exceed the one million like the Sandia VAWT when calculated at
equator station and as a first approximation Re effects are neglected.
urel ca

Re =

(2.9)

Where urel is the relative velocity and the kinematic viscosity of air.
The structural properties of the scaled Sandia VAWT can now be computed, having the
basic expressions for up-scaling derived. The area moment of inertia Iaa of an arbitrary
cross-section A about an arbitrary axis aa is given by equation 2.10.
Z
Iaa =

r2 dA

(2.10)

Where dA is the differential area of the cross-section and r is the distance of the dA from
the aa axis. It can be deduced from the above equation that the area moment of inertia
has units of length to the forth power. Consequently, applying geometric similarity, the
area moment of inertia of a scaled cross-section will be proportional with the scaling factor
to the forth power ( sf 4 ). Similarly, the cross-section area A scales with the square of
the scaling factor ( sf 2 ).
The mass m of a component is the product of its volume V with the density so applying
pure geometric similarity and assuming constant density the mass will be proportional to
sf 3 since the volume is proportional to sf 3 . The Youngs (E) and shear (G) moduli
are intensive properties of the material, so are independent of the size of the structure.
This also holds for composite parts assuming linear structural behaviour and keeping the
layout of the layers unchanged.
The scale dependencies of the rotor as a result of aerodynamic and geometric similarities
are summarized at table 2.3 on the next page.
It should be noted that pure geometric similarity results in blade weights which follow
the cubic law. However, data from manufacturers reveal that the mass of larger blades
can be less than the one anticipated by the cubic law [32, p. 86].
Applying the above relations the geometric and structural properties of the Sandia VAWT
rotor are computed. The up-scaled blade structural properties are compared with the

2.2 Vertical Axis Wind Turbine Model

17

Table 2.3: Scale dependencies of rotor properties applying aerodynamic and geometric
similarities

Property

Scale dependency
sf 0
sf 0
sf 0
sf 1
sf 1
sf 1
sf 1
sf 1
sf 2
sf 4
sf 3
sf 0
sf 0
sf 1

Solidity
Tip speed R
Tip speed ratio
Rotor speed
Rotor radius R
Rotor height H
Blade chord ca
Blade thickness ta
Blade cross-section area A
Blade area moments of inertia I
Blade mass m
Youngs modulus E
Shear modulus G
Reynolds number Re

DeepWind blades. In figure 2.5 are shown the rotor layouts (blade center-lines) for the two
models. The blade chord, mass per length, flatwise, edgewise, torsional and longitudinal
stiffness along the blade span are plotted in figures B.1, B.2 and B.3 in appendix B on
page 109.

Position z [m]

150

100

DeepWind rotor
Sandia scaled rotor
50

0
-80

-60

-40

-20

20

40

60

80

Position y [m]

Figure 2.5: DeepWind and Sandia scaled blade center-lines layout.

It can be seen from the plots of blade stiffness that the DeepWind model has more flexible
blades in all directions (flatwise, edgewise, torsional and longitudinal) compared with the
Sandia scaled blades, although the DeepWind blade chord is larger throughout the span
(figure B.1 on page 109). There is also a large difference in mass per length (figure B.1
on page 109) but as stressed before the mass in real designs does not follow the cubic
rule, so the scaled Sandia blade mass per length values are conservative. In addition, the
material that was used in Sandia blades was aluminium and in DeepWind are composites
which have less density.

18

Analysis Tools and VAWT Model

An eigenvalue analysis of the DeepWind rotor and the up-scaled version of the Sandia
34 m rotor is performed using HAWC2 code. The natural frequencies are summarized
at table 2.4. The DeepWind blade natural frequencies are significantly lower from their
equivalent of the Sandia scaled model. This happens due to much lower blade stiffness
of the DeepWind blades and can be connected with the high blade deflections during
operation as were found in the previous section.
Table 2.4: Natural frequencies for the DeepWind and Sandia scaled rotors.

Mode shapes
1st Flatwise antisym.
1st Flatwise symmetric
1st Propeller
1st Butterfly
2nd Flatwise antisym.
2nd Flatwise symmetric
1st Tower in-plane
1st Tower out-of-plane
3rd Flatwise antisym.
3rd Flatwise symmetric
2nd Propeller

2.2.2

Sandia scaled model

DeepWind model

Natural frequency [Hz]

Natural frequency [Hz]

0.32
0.32
0.42
0.49
0.59
0.61
0.70
0.72
0.99
1.00
1.05

0.2
0.2
0.26
0.26
0.44
0.44
0.31
0.35
0.77
0.77
0.68

Modification of the 5 MW DeepWind Rotor

The onshore VAWT aeroelastic model that is developed is based on the DeepWind rotor
with some modifications, consulting properties of the scaled Sandia VAWT, such as the
blade stiffness. Regarding the electric generator; a direct drive one was used in the
DeepWind project with integrated shell as part of the floater which is not applicable in
this onshore model. Thus, the NREL 5 MW HAWT [35, p. 175] generator properties
are used which has the same rated power. Moreover, the NREL 5 MW wind turbine
aeroelastic model in HAWC2 is available and is used in order to make some comparisons
of loads with the VAWT later. For this purpose the base length of the developed turbine
is adjusted such that the hub height of the VAWT to be equal with the HAWT one. The
definition of the hub height for the VAWT is given in section 3.1 on page 36 in the general
considerations on the definitions of the standard. The base of the VAWT is assumed very
stiff compared to the blade and tower stiffness as it is relatively short and it is usually
the case in VAWTs such as the Sandia 34 m test-bed [72, p. 61]. It is stressed that in
this study are not considered aspects on manufacturability, neither taken into account
details in the electrical sub-systems of the turbine, the electrical grid side and the control
strategies. Finally, it is noticed that the case of a 3-bladed rotor was considered in the
beginning of the study in order to be used as a baseline instead of the 2 bladed. The
conversion from 2 to 3 blades rotor was based on solidity similarity keeping the same tip
speed ratio. Then according to [75] for the same turbine radius the chord length is 2/3 of
the 2 bladed turbine. As a first approximation the blade weight remains the same while

2.2 Vertical Axis Wind Turbine Model

19

2
the area moments of inertia are considerably reduced (I3B = 23 I2B ). This led to a more
flexible rotor and the blades due to their own weight experienced high deformations. The
latter also made it impossible to run simulations without extensive modifications of the
rotor and was decided to continue with a 2 bladed rotor. The investigation of DLCs for
a similar rated power VAWT with 3 blades is recommended as future work.
Blades and Tower
The original blade and tower stiffness as given in the DeepWind report [80, p. 175] are
changed in order to reduce the very large blade edgewise deflections in the equator and
find operational rotor speed ranges using the Campbell diagram. The airfoils and the
chord length are not changed as this will alter the whole rotor model. In other words
during this iterative procedure it is not main target to design an optimum VAWT as this
requires multi-disciplinary design, analysis and optimization (MDAO) framework which
goes beyond this study.
Due to large blade edgewise deflections during operation found previously, the edgewise
stiffness is increased by changing the cross-section layout. The existing shear webs are set
closer to the leading and trailing edge of the blade section and four new are added in order
to increase primarily the edgewise stiffness. In addition, carbon-fibre composite material is
used to keep the mass close to the original value. This is important because from the blade
mass are dependent the centrifugal and gravitational loads which in turn determine the
Troposkien shape and in this case was kept the same for simplicity. An MDAO technique
could probably result in a design using glass-fibres together with a modified chord length
and Troposkien shape. The new blade cross-section beam properties that used in the
HAWC2 model are calculated using the cross section analysis software BECAS [8]. A
comparison between the original DeepWind, the modified DeepWind and the Sandia
scaled blade mass per length and stiffness is performed in figures 2.6, 2.7 and 2.8 below.
The new blade mass is 52.4 tons in contrast to the 47.4 tons of the original DeepWind
blades.
1400

DeepWind blade
Modified DeepWind blade
Sandia scaled blade

Mass per Length [kg/m]

1200

1000

800

600

400

200
0

10

20

30

40

50

60

70

80

90

100

Distance along blade from bottom to top [%]

Figure 2.6: Original and modified DeepWind and Sandia scaled blade mass per length
distribution.

20

Analysis Tools and VAWT Model

10 9

5
DeepWind blade
Modified DeepWind blade
Sandia scaled blade

10 10

3.5

Edgewise Stiffness (EI y) [Nm ]

Flatwise Stiffness (EI x) [Nm ]

2.5
2

1.5
1

0.5
0

0
0

20

40

60

80

100

Distance along blade from bottom to top [%]

20

40

60

80

100

Distance along blade from bottom to top [%]

Figure 2.7: Original and modified DeepWind and Sandia scaled blade flatwise (left) and
edgewise (right) stiffness distribution.
10 9

3.5

Longitudinal Stiffness (EA) [N]

Torsional Stiffness (GJ) [Nm ]

DeepWind blade
Modified DeepWind blade
Sandia scaled blade

3.5
3
2.5
2
1.5
1
0.5
0

10 10

3
2.5
2
1.5
1
0.5
0

20

40

60

80

100

Distance along blade from bottom to top [%]

20

40

60

80

100

Distance along blade from bottom to top [%]

Figure 2.8: Original and modified DeepWind and Sandia scaled blade torsional (left) and
longitudinal (right) stiffness distribution.

A final parameter that has to be set is the structural damping of the blades which is done
through the Rayleigh damping input parameter in HAWC2 main file. This tuned so as
the logarithmic damping decrement to be between 2% and 3% for the first eigenmodes.
The original and the modified cross-section layouts are depicted in figures 2.9 and 2.10.

Figure 2.9: Original DeepWind blade cross-section layout for NACA 0018 (up) and
NACA 0025 (down) profiles [80, p. 17].

Next the internal loads and stresses for the modified blades are computed at turbines

2.2 Vertical Axis Wind Turbine Model

21

Figure 2.10: Modified DeepWind blade cross-section layout for NACA 0018 (up) and
NACA 0025 (down) profiles.

Troposkien rotor speed (0.62 rad/s) due to gravity and centrifugal forces excluding the
aerodynamics and compared with the original. This is done in order to see the difference
between the original and modified blade because the flatwise stiffness was increased together with a small change in the mass per length distribution but the Troposkien shape
was kept the same. In figure 2.11 on the following page are illustrated the gravitational
and centrifugal forces acting to the rotating Darrieus VAWT blade. The centrifugal forces
create a constant loading in the blade. The gravity in this case also creates a constant
load to the blade in contrast to the HAWT where it has sinusoidal form due to rotation
around a horizontal and not the vertical axis relatively to the ground. The internal blade
loads are the normal force FN and the internal flatwise Mx , edgewise My and torsional
Mz moments. The blades are connected to the tower on the elastic center and the blade
centreline is the line that goes through the elastic center of blade cross sections. As a
result, the stresses from the edgewise and torsional moments are practically zero and only
the normal stresses due to the internal normal force and flatwise bending moments (equations 2.11 and 2.12 on page 21) are presented. It should be mentioned that the DeepWind
blade center-line curve was such to minimize the blade flatwise stresses (F + M ) at the
Troposkien rotor speed.
F =

M =

FN
A

(2.11)

Mx y
Ix

(2.12)

The results are plotted in figure 2.12 on the next page where tension at the outer side
of the blade is defined as positive. It can be seen that there is a very small difference in
the flatwise moment and stress distribution between the old and the new blade except

22

Analysis Tools and VAWT Model

Figure 2.11: Centrifugal and gravitational loads acting to a rotating Darrieus VAWT blade.

at around 125 m blade region. In figure 2.13 on the facing page are plotted the flatwise
moment and stress when the turbine is standing-still and exposed only to the gravitational
field. The flatwise moment distribution for the two blades is the same except the two
peaks at around 60 and 130 m span position where the moment on the modified blade
due to slightly higher weight is larger. Both stresses on the other hand are practically
the same and have an abrupt change around 178 m blade span where the transition from
NACA 0018 to NACA 0025 cross-section is present.

Figure 2.12: Comparison of flatwise moment (left) and stress (right) on the blade due to
gravity and centrifugal loads between the original and the modified DeepWind blade at rated rotor
speed (0.62 rad/s).

A modal analysis of the modified blade was performed using HAWC2s eigenvalue solver
and the first couple of natural frequencies with the corresponding structural logarithmic
damping decrement are given in table 2.5 on the next page. The same boundary conditions
with the two ends of the blade fixed rigidly to a rigid frame are applied as in the case of
the blade-tower connection.

2.2 Vertical Axis Wind Turbine Model

23

Figure 2.13: Comparison of flatwise moment (left) and stress (right) on the blade due to
gravity between the original and the modified DeepWind blade with the turbine standing-still.
Table 2.5: Natural frequencies and structural damping of the modified blade.

Mode shape
1st Flatwise
1st Edgewise
2nd Flatwise
2nd Edgewise
3rd Flatwise
4th Flatwise
3rd Edgewise

Nat. frequency [Hz]

Log. damping decrement [%]

0.35
0.43
0.78
1.08
1.34
2.01
2.40

2.7
2.9
6.0
7.4
10.3
15.4
16.3

The tower can be excited from the aerodynamic loads of the passing blades and any
uneven distribution of mass around the rotor axis. The latter source of dynamic loading
is ignored assuming that no unbalanced mass in the rotor structure exists. The tower
stiffness is increased using the Campbell diagram as guidance. The reason is to create
areas that the rotor can rotate at variable speed without the need to pass often through
crosses that can lead either to resonances or increased fatigue damage. Critical frequencies
for a 2-bladed VAWT wind turbine occur when the first couple tower natural frequencies
cross the two per revolution frequency 2P and its first multiples (2P n, n = 1, 2, ...) or
when the blade lower natural frequencies cross the per revolution frequency P and its
multiples. Resonances from the gearbox and the generator can also occur and should be
avoided or damped but in the context of this study these components are assumed rigid so
no such critical frequencies exist. The tower (rotating column) cylindrical tapered shape
and length (143 m) are kept constant and the towers diameter is scaled gradually in order
to achieve the previous defined objective. The material of the tower is construction steel
with density of 7850 kg/m3 . The scaling factor sf is given in equation 2.13 below as the

24

Analysis Tools and VAWT Model

ratio of the scaled to the original tower diameter. As a result, the thickness, the stiffness,
the mass and other structural properties can be scaled similarly like in the case of the
blades and are summarized at table 2.6.
sf =

Dsc
D

(2.13)

Table 2.6: Scale dependencies of the tower structural properties.

Property
Tower diameter D
Tower length H
Tower thickness tt
Tower cross-section area A
Tower area moments of inertia I
Tower mass m
Youngs modulus E
Shear modulus G

Scale dependency
sf 1
sf 0
sf 1
sf 2
sf 4
s3
sf 0
sf 0

The scaling factor was increased gradually and a Campbell diagram was created every
time. Then was decided if a new increment of the tower strength was necessary or not
by inspection of the line crossings in the diagram. As there is not a built-in future in
HAWC2 for the creation of a Campbell diagram the following procedure was developed:
1. An external impulse force is applied to a component of the wind turbine in order to
excite it.
2. The response of a components node from the simulation such as a deflection or
acceleration is written to a file. It should be noted that a long simulation time after
the force excitation is preserved (ideally infinite) such that the vibration due to
excitation to be close to zero at the end of the simulation. This is needed in order
the analysis of the next step to give correct results.
3. Spectral analysis is performed in the captured signal and the natural frequencies of
the component are identified.
4. This procedure is repeated for different rotor speeds from zero up to rated. In
addition, at a given speed the application of the external force is applied more
than one time in order to excite different parts or even different modes of the same
component like the blade flatwise and edgewise modes.
5. Finally, the natural frequencies are plotted as a function of rotor speed together with
the lines of the per revolution frequency and its first multiples (2P n, n = 1, 2, ...)
creating the turbines Campbell diagram.
For the creation of the Campbell diagram the wind was not taken into account in order to
evaluate only the structural behaviour of the wind turbine without any effect on damping

2.2 Vertical Axis Wind Turbine Model

25

due to aerodynamics. An appropriate wind turbine operational range was found when
the tower was scaled with sf = 1.20. The tower mass increased to 511 tons from the
initial value of 355 tons. A comparison between the original and the modified DeepWind
tower mass per length and stiffness is given in figures 2.14 and 2.15. The low shaft is
assumed 6 m long after reviewing the scaled length of the same part from the Sandia
34 m test-bed. The structural properties of the low shaft are taken equal with the tower
bottom cylindrical section properties. Finally, the base is 21 m long and was assumed
much stiffer compared to the tower because this was also the case for the Sandia VAWT.
In table 2.7 on the next page are summarised the geometric properties of the original and
the modified tower.
5500
DeepWind tower
Modified tower

Mass per Length [kg/m]

5000
4500
4000
3500
3000
2500
2000

Bending Stiffness (EI x,y) [Nm2]

10

1500

10 11

9
8
7
6
5
4
3
2
1

50

100

50

100

Distance from bottom to top [m]

Distance from bottom to top [m]

Figure 2.14: Original and modified tower mass per length (left) and bending stiffness (right)
distributions.
10 11

14
DeepWind tower
Modified tower

7
6
5
4
3
2
1
0

Longitudinal Stiffness (EA) [N]

Torsional Stiffness (GJ) [Nm2]

10 10

12

10

4
0

50

100

Distance from bottom to top [m]

50

100

Distance from bottom to top [m]

Figure 2.15: Original and modified tower torsional and longitudinal stiffness distributions left
and right respectively.

26

Analysis Tools and VAWT Model

Table 2.7: Geometric properties of the wind turbine tower.

Profile
Cylindrical
(step tapered)
section
section
section
section

1 (bottom)
2
3
4 (top)

Length
[m]
35.75
35.75
35.75
35.75

Outer diameter

Thickness

[m]
Original Modified

[mm]
Original Modified

6.40
5.84
5.32
4.62

7.67
7.00
6.40
5.54

21.90
18.00
17.00
15.00

26.40
21.70
20.50
18.10

The Campbell diagram of the modified VAWT is plotted in figure 2.16 on the facing page
where the natural frequencies of the blades are found from the PSD analysis of the blade
deflections expressed in the rotating frame located on the blade roots, while the tower
natural frequencies from the PSD analysis of the corresponding signals expressed in the
global fixed frame located on the turbine base bottom. Consequently, critical rotational
speeds for the tower occur when its natural frequencies cross the two per revolution and
its multiplies (2P n, n = 1, 2, ...). For the blades, crossings with 1P and all the multiplies
(1P n, n = 1, 2, ...) can be potentially critical but as stated in [5, p. 10] odd modes
will couple with odd per-revolution frequencies (n = 1, 3, ...) and even modes with even
per-revolution frequencies (n = 2, 4, ...). More details about the critical per-revolution
excitation frequencies expressed in rotating or fixed frames for VAWTs can be found in
reference [58, p. 5].
The rotor speed under normal power production varies from around 0.4 rad/s up to the
rated 0.62 rad/s. The 1st blade flatwise symmetric and anti-symmetric natural frequencies
(1F S, 1F A) cross the 4P and 5P frequencies at 0.446 rad/s and 0.559 rad/s respectively.
Similarly, the 1st propeller frequency 1P r crosses the 5P and 6P but as is it even mode,
the crossing with the 6P is expected more critical. The 1st blade edgewise 1BE (butterfly
mode) natural frequency crosses the 5P , 6P and 7P where the 5th and 7th per revolution
frequency crossings, at 0.61 and 0.43 rad/s respectively, should be avoided primarily as
BE mode is odd. The 1st tower in and out of plane natural frequencies (1T I, 1T O)
cross only one multiple of 2P inside the operational rotor speed range namely the 4P at
around 0.588 rad/s. It is stressed that resonance only occurs if the turbine operates in
critical or close to critical frequencies and the damping is not high enough.
An investigation was done in order to confirm the results from the Campbell diagram
by simulating the wind turbine from cut-in up to cut-out wind speed without turbulence
and rotor speeds from 0.4 to 0.62 rad/s with a step of 0.05 rad/s. Then the maximum
bending moments of the blade-1 low roots and turbine base bottom bending moments
were extracted and plotted in appendix B.2 on page 111 (figures B.4, B.5, B.6 and B.7).
From the results can be identified the critical frequencies that should be avoided by finding
the peaks in the contours. These are:
At 0.43 rad/s due to crossing of the 1P r-6P and 1BE-7P and correspond to strong
resonances since the loads in that region increase a lot.
At 0.53 rad/s due to crossing of the blade 1P r with 5P which is of medium strength
resonance and

2.2 Vertical Axis Wind Turbine Model

27

Around 0.575 rad/s where the 1st tower in plane and out of plane modes cross the
4P which is of medium strength resonance.
The natural frequencies of the modified standing still turbine, as well as, the logarithmic
damping decrement in percentage are given in table 2.8 on the following page.
8P

0.55

7P

6P
5P

0.5

1BE

0.45

Frequency [Hz]

1Pr

4P

0.4
1TI, 1TO
0.35
1FA, 1FS

3P

0.3

0.25

2P

0.2
0

0.1

0.2

0.3

0.4

0.5

0.6

Rotor speed [rad/s]

Figure 2.16: Campbell diagram of the modified DeepWind VAWT. F S, F A denote the blade
flatwise symmetric and anti-symmetric modes, T I, T O the tower in-plane and out-of-plane
modes, P r the Propeller mode and BE the blade edgewise or butterfly mode.

It is stressed again that the main criterion was to modify the blade and tower properties
in order to confine the blade deflections during normal operation and have a working
model for the next phase of the study and not to optimise structurally the rotor.

28

Analysis Tools and VAWT Model

Table 2.8: Natural frequencies and structural damping for the DeepWind modified rotor.

Mode shape

Nat. frequency
[Hz]

Log. damping decrement


[%]

0.35
0.35
0.36
0.37
0.40
0.47
0.78
0.78
1.04
1.10
1.34
1.34

2.6
2.7
2.1
2.1
2.7
3.0
6.0
6.0
7.1
7.4
10.2
10.3

1st Flatwise antisym. (1FA)


1st Flatwise symmetric (1FS)
1st Tower in-plane (1TI)
1st Tower out-of-plane (1TO)
1st Propeller (1Pr)
1st Butterfly (1BE)
2nd Flatwise antisym.
2nd Flatwise symmetric
2nd Propeller
2nd Blade edgewise
3rd Flatwise antisym.
3rd Flatwise symmetric
Guy Wires

The modified DeepWind rotor as described above without guy wires is used as a baseline
for the design load cases study. Guy wires have been used in the past as an option to
support the tower top of onshore VAWTs and alleviate the bending moments on the root
of the rotor. Thus, a complementary VAWT aeroelastic model with wires is described in
this paragraph and a comparison of loads in normal power production between the two
models is also performed in the next chapter. The wire properties of Sandia 34 m turbine
(table 2.9) are up-scaled and implemented as a first approximation.
Table 2.9: Wire properties of Sandia 34 m VAWT [72].

Property

Value

No of wires
Ground-wire angle
Pretension per wire
Wire axial stiffness

3 sets of 2
35 [deg]
413.83 [kN]
7883.39 [kN/m]

Instead of 2 wires per set one equivalent is used. A 10 m shaft is added in the tower-top
of the modified DeepWind rotor to avoid wire collision with the blades upper part. In
addition, between the upper shaft and the tower top a bearing connection is placed to
allow the tower to rotate relatively to tower top. The angle between the ground and the
cables is kept at 35 which results to a minimum downward force of the tower for a given
stiffness [12, p. 11]. The shaft cross section structural properties assumed equal with
the tower upper cross section properties. The scaling factor sf is kept the same as was
defined in the Sandia 34 m up-scaling analysis. In table 2.10 on the facing page the scale
dependencies of the wire properties are given.
The guy wires are modelled in HAWC2 as non linear beam elements without bending
stiffness. Pretension of the wires is applied by reducing the actual wire length in the

2.2 Vertical Axis Wind Turbine Model

29

Table 2.10: Scale dependencies of the guy wire structural properties.

Property

Scale dependency

Youngs modulus E
Cross section area A
Longitudinal stiffness EA
Axial stiffness EA/l
Pretension force Fp

sf 0
sf 2
sf 2
sf 1
sf 2

model. The wires create a downward force to the tower which is equivalent of three
times the rotor mass (616 tons). The natural frequencies of the wind turbine with the
wires are summarized in table 2.11. Compared with the ones without the guy wires the
most important difference is found in the 1st tower natural frequencies which have been
increased. The blade flatwise natural frequencies remain the same, while the propeller
and butterfly frequencies have a minor change. The Campbell diagram is depicted in
figure 2.17 on the following page. Since the tower natural frequencies have been increased
there is more space for the turbine to operate at variable speed with less crossings in
between.
Table 2.11: Natural frequencies and structural damping for the DeepWind modified rotor with
guy wires.

Mode shape
1st Flatwise antis. (1FA) coupled with tower
1st Flatwise symmetric (1FS)
1st Propeller (1Pr)
1st Butterfly (1BE)
2nd Flatwise antisym. coupled with tower
2nd Flatwise symmetric
2nd Propeller
2nd Blade edgewise
1st Tower in-plane (1TI) coupled with blades
1st Tower out-of-plane (1TO) coupled with blades

Nat. freq.
[Hz]

Log. damping
decrement [%]

0.35
0.35
0.41
0.43
0.78
0.78
1.07
1.07
1.19
1.25

2.6
2.6
2.7
2.8
5.9
6.0
7.3
7.0
6.5
7.0

30

Analysis Tools and VAWT Model

8P

7P

6P

0.55

5P
0.5

0.45

Frequency [Hz]

1BE

4P

1Pr

0.4

0.35

1FA, 1FS

3P

0.3

0.25

2P

0.2
0

0.1

0.2

0.3

0.4

0.5

0.6

Rotor speed [rad/s]

Figure 2.17: Campbell diagram of the modified DeepWind VAWT with guy wires. F S, F A
denote the blade flatwise symmetric and anti-symmetric modes, P r the Propeller mode and BE
the blade edgewise or butterfly mode.

Generator model
Next a simplified generator model is implemented in the aeroelastic model with an external
DLL according to the notes in reference [38]. The model represents the dynamics of an
induction generator which are approximated by a first order differential equation. The slip
s is given by equation 2.14, where s and r are the synchronous and the actual generator
speed respectively. The static torque characteristic as a function of slip is approximated
with a 1st order polynomial given by equation 2.15 on the facing page where n is the
nominal generator speed, Pnmech is the nominal rotor power expressed at the generator
side (high speed shaft) and sn the nominal slip. In figure 2.18 on the next page with
red colour is depicted a typical static torque slip curve of an induction generator while
with blue colour the first order approximation. The solid black lines which create a cross
depict the nominal point of operation where the slip is equal to 0.6 %. The slip affects
how stiff the generator is, or in other words higher slip values enable the generator speed
(equivalently the rotor speed) to vary more around the nominal value if a fluctuation
in rotor torque is present. On the other hand, higher slip increases the losses of the
generator. In order to achieve variable speed operation with this model it is assumed that
an inverter is used which changes the generator synchronous speed according to the rotor
speed. In addition, the generators maximum torque (pull-out torque) is assumed higher
than the equivalent maximum rotor torque during normal power production from cut-in
to cut-out wind speeds.

s=

r s
r

(2.14)

2.2 Vertical Axis Wind Turbine Model

M (s) =

Pnmech s
n sn

31

(2.15)

Figure 2.18: Static torque slip curve of an induction generator with a nominal slip of 0.6% [38].

The generator is located on the ground like the Sandia 34 m test-bed and the generator
mass moment of inertia is taken equal to the NREL 5 MW wind turbines value [35, p. 14]
which is 534 kgm2 expressed at high speed shaft HSS. The gearbox ratio is equal to 87:1
and for the baseline model the slip is taken equal to 0.8%.
Implementation in HAWC2
The structural properties of the VAWT components are inserted in HAWC2 as external
files and are modelled as Timoshenko beam elements. For the aerodynamic computations
the actuator cylinder model [44] is used together with the Stig ye dynamic stall model
[59]. This dynamic stall model is not appropriate when deep stall occurs which could be
the case for Darrieus rotors close to the blade roots. In addition, after initial simulations,
lift coefficients of more than 3 [-] where identified close to the blade roots for high wind
speeds, which are considered rather unrealistic, thus the dynamic stall model was decided
to be deactivated very close to the blade roots. An implementation of the dynamic stall
model in HAWC2 able to capture the high rate of change of angle of attack and predict
reliable the lift coefficient is referred to future work. The drag of the cylindrical tower is
computed using the components drag coefficient and the projected area normal to the
mean wind velocity direction. The influence of the tower in the wind field downstream is
also captured, based on the boundary layer solution of an axisymmetric jet stream (jet
model). In table 2.12 on the following page are summarized the main characteristics and
properties of the modified DeepWind VAWT aeroelastic model which on the remaining
study is referred as VAWT. Lastly, in figure 2.19 on page 33 is illustrated the turbine
geometric layout with some basic parameters.

32

Analysis Tools and VAWT Model

Table 2.12: Operational and geometric data of the modified DeepWind Rotor which is used on
the remaining of this study and will refer as VAWT.

Property
Rated electrical power
Rated rotor speed
Rated wind speed
cut-in wind speed
cut-out wind speed
Rated tip speed
Hub height
Number of blades
Rotor diameter
Rotor height to diameter ratio
Blade chord
Solidity
Swept area
Rotor inertia on LSS
Generator inertia on HSS
Gearbox ratio
Rotor mass
Blade mass
Total height
Rotor height

Value
5
5.95
0.62
13.5
4
25
37.5
90
2
121
1.18
5
0.165
11996
184.2x106
534
87:1
615800
52400
170
143

[MW]
[rpm]
[rad/s]
[m/s]
[m/s]
[m/s]
[m/s]
[m]
[-]
[m]
[-]
[m]
[-]
[m2 ]
[kgm2 ]
[kgm2 ]
[-]
[kg]
[kg]
[m]
[m]

2.2 Vertical Axis Wind Turbine Model

Figure 2.19: Schematic of the 2-bladed vertical axis wind turbine.

33

34

Analysis Tools and VAWT Model

Chapter 3
Design Load Cases

First time that was tried systematically to set rules and recommendations for the structural integrity of large wind turbines was by the International Energy Agency (IEA) in
1984 [26] and in Germany by Hu and Hau in 1986 [25] even though certification of wind
turbines was initiated in Denmark as early as 1979 according to reference [53]. Directives
started to take a form of national standards in Denmark [30], The Netherlands [55] and
in Germany [19]. Classification societies also composed their first guidelines for wind turbines like the Germanischer Lloyd (GL) Regulation for the Certification of Wind Energy
Conversion Systems Vol. IV Non-Marine Technology, Part 1 - Wind Energy [68] and
Det Norske Veritas (DNV) with Ris National Laboratories for Sustainable Energy the
Guidelines for Design of Wind Turbines [81]. Parallel the International Electrotechnical
Community (IEC) introduced the first standard for the safety requirements of wind turbine generator systems in 1994 [27] and the European certification bodies initiated the
JOULE project [53] in order to harmonise the standardisation of wind turbines within
Europe according to the IEC 61400 series standard substituting national norms.
The latest international standard that covers the minimum design requirements of onshore wind turbines is the IEC 61400-1 3rd edition of 2005 [28]. It was revised last time
before ten years having in focus the HAWTs as the predecessor rules and norms, and sets
the minimum structural requirements of wind turbines in general. This means that the
design requirements for a VAWT in principle are covered by this, even though has not
been examined in detail its applicability and adequacy so far for the vertical axis wind
turbines. Thus, the study of VAWTs structural integrity through this standard will reveal
shortcomings and points that possibly need to change for the case of VAWTs. In the rest
of this study when IEC 61400-1 or standard is stated it will refer to the latest edition
of IEC 61400-1 (3rd edition of 2005).

35

36

3.1

Design Load Cases

General Considerations on the Definitions of the Standard

The DLCs as proposed by the IEC 61400-1 are a combination of external conditions and
wind turbine states. Electrical power network, wind and other environmental conditions
including the soil parameters compose the external conditions which are further divided
into normal and extreme. The latter refer to rare external states like 50 year recurrence
period wind.
Wind is the main environmental parameter for wind turbines and in the IEC 61400-1
three standard wind turbine classes are defined representing different wind characteristics
in order to cover many sites. There is also a special class that can be applied in sites
with cyclones, typhoons or other harsh climate conditions. The normal classes are characterized by a reference wind speed Vref averaged over 10 minutes at hub height with
a corresponding annual average wind speed at hub height Vave equal to 0.2Vref and are
summarized at table 3.1, together with an expected turbulence intensity Iref at a 10 min
average wind speed of 15 m/s (table 3.2).
Table 3.1: Wind turbine classes [28]

Wind turbine class

Vref [m/s]

Vave [m/s]

50
42.5
37.5

10
8.5
7.5

I
II
III

Table 3.2: Site turbulence characteristics [28]

Turbulence category
Site with high turbulence characteristics
Site with medium turbulence characteristics
Site with low turbulence characteristics

A
B
C

Iref [-]
0.16
0.14
0.12

In the context of this study wind turbine class II is selected as it is the median between
the three. IEC 61400-1 predefines 22 load cases combining different wind conditions and
turbine states, 17 of them concerning ultimate loading and 5 fatigue damage. Site specific
conditions should be taken into account as mentioned at clause 11 of the standard, like
earthquakes, extreme temperature ranges, icing, soil characteristics, wake effects from
neighbouring wind turbines in combination with electrical network conditions. Cases
dependent on control strategies such as start up and normal shut down are not studied,
since the results are highly dependent of the turbine control algorithm. Are also not
studied the loads that emerge from transport, assembly, maintenance and repair of wind
turbine components(DLC 8.x). The DLCs considered in this study are summarized at
table 3.3 on the facing page.
The main outputs that are extracted from the simulations are the blade upper and lower
root moments, the wind turbine base bottom moments, the blade deflection and angle of
attack at equator height and the rotor torque, speed and power. In the remaining analysis

3.1 General Considerations on the Definitions of the Standard

37

Table 3.3: Design load cases

Design situation

DLC according
to IEC 61400-1

Wind condition

Type of
analysis

Power production
Power production
Power production
Power production
Power production
Power production
Power product. VAWT
with guy-wires
Emergency shut down
Parked-standing still
rotor
Parked-idling rotor

1.1
1.2
1.3
1.5
2.3
-

NTM
NTM
ETM
EWS
EOG
EDC
NTM

Ultimate
Fatigue
Ultimate
Ultimate
Ultimate
Ultimate
Ultimate

5.1
6.1/6.2/6.3/7.1

Ultimate
Ultimate

6.1/6.2/6.3/7.1

Parked-idling rotor

6.4

NTM
EWM
return
EWM
return
NTM

1 and 50 year
periods
1 and 50 year
periods

Ultimate
Fatigue

the turbine base bottom loads are always expressed in the global fixed coordinate system
located in the base of the turbine. The blade root loads to the local element coordinate
systems and the blade deflections to the blade coordinate systems (see figures 2.2 and 2.3
on page 13).
As was written in the introduction the IEC 61400-1 even though refers to large wind
turbines in general, it was written focusing mostly on HAWTs. Thus, some definitions
may not be clear when VAWTs are screened. Below are stressed these aspects in order to
be taken into account on a future revision of the standard. Furthermore, the assumptions
on these aspects that are made in this study are stated.
Firstly, about the height where the wind reference values are applied such as the
mean wind at hub height. In HAWTs the hub height is the rotor center which
corresponds to the centre of the swept area. In the DeepWind project the equator
height was taken as the hub height for the simulations, where equator is defined as
the point of the blade with the largest distance (radius) from the tower center-line.
A horizontal line passing at that height cuts the rotor swept area in two unequal
parts. Forty percent of the rotor area is below that point when the rotor swept
area is analysed with the blades at Troposkien shape. In addition, the equator
height definition becomes more inconsistent in V-rotor or Delta-rotor VAWT
configurations. In the case of DeepWind rotor the equator is located 9 m lower from
the point where the rotor swept area is halved. In this study as hub height is taken
the point where a horizontal line passing that point cuts the rotor swept area into
two equal pieces.
Secondly, the rotor diameter D is needed in some equations of the standard for
the definition of wind characteristics such as the gust magnitude. For the case of

38

Design Load Cases

VAWT the diameter can be a function of rotor height like in the case of Darrieus
configurations. In this work the largest diameter of the turbine is used.
In paragraph 7.4.1 of the standard it is mentioned that the statistical analysis of
DLC 1.1 simulation shall include the tip deflection. For the case of VAWT, deflected
blades are not expected to collide with any other part of the turbine in contrast to
upwind HAWT where the blade tip comes close to the tower due to aerodynamic
loading and collision may occur. Thus calculation of blade loads should be required
and blade deflections could be optional.

3.2

Normal Power Production

After setting the aeroelastic model the turbine is subjected to simulations from cut-in to
cut-out wind speeds at hub height (Vhub ) with normal turbulence model (NTM) according
to the IEC 61400-1. The NTM is based on the Mann uniform shear turbulence model with
isotropic von Karman energy spectrum modified to account for anisotropic atmospheric
conditions using rapid distortion theory [46]. The applied longitudinal standard deviation
(90% quantile) of the turbulence 1 is given by equation 3.1 and the normal wind profile
(NWP) model which is a power law with exponent (wind shear) of 0.2 [-] by equation 3.2.
1 = Iref (0.75Vhub + 5.6)

V (z) = Vhub

z
zhub

(3.1)

0.2
(3.2)

Where z is the height measured from the ground level. The subscript hub refers to hub
height.
Turbulence category B (Iref = 0.14 [-]) is used as basis in the simulations and the statistics
such as absolute maximum, mean and standard deviation of different outputs are investigated. Then ultimate and fatigue load analyses are performed followed by sensitivity
analysis of main parameters.
In addition, a direct comparison of outputs with the NREL 5 MW, 3 bladed HAWT [35]
is made when it is applicable in order to see the trends and the main differences between
the two configurations. The basic properties of the NREL wind turbine are given in
appendix E on page 123. Both turbines are designed for a rated electrical power output
of 5 MW. The cut-in wind speeds are 3 m/s and 4 m/s for the HAWT and the VAWT
respectively, while the cut-out is the same and equal to 25 m/s. The common operation
wind speed range is compared through the division in wind speed bins of 1m/s and the
simulations are performed at the center of each bin. The applied wind climate is the same
for both turbines and the wind speed nominal values refer to the hub height where it is
located 90 m above ground level.
In figure 3.1 on the facing page the mean and standard deviation (STD) of the rotor
power between the two turbines is compared. The mean power output above rated speed
for the VAWT is not totally constant as a controller is not implemented. The rotor

3.2 Normal Power Production

39

speed for every wind speed was defined manually through the generator synchronous
frequency in order to have the optimum power production below rated winds and a rated
rotor power of around 5.8 MW above rated wind speeds. This rotor power with the
mechanical and electrical losses of all parts of the turbine gives a rated electrical power
of around 5 MW which was the target. On top of that constrain the critical crossings
from Campbell diagram were taken into account to avoid resonances as was analysed in
the previous chapter. It is important to notice the much higher standard deviation of
the power output for the VAWT. This is inherent in VAWTs especially for two bladed
configurations where the angle of attack varies a lot during one revolution and in turn the
aerodynamic loads on the blades which produce the rotor torque vary highly, as well. In
figure D.1 in appendix D on page 115 is plotted the mean rotor thrust for the horizontal
and vertical axis wind turbines, where it can be seen that up to rated wind speed the
thrusts are almost equal. Then for the VAWT the thrust continues to increase while for
the HAWT due to blade pitching decreases. In figures 3.2 and 3.3 on page 40 are plotted
the time series of the rotor power and torque at 13.5 m/s wind speed simulations with and
without turbulence, where it can be seen the high variation during a full rotation of the
rotor. These fluctuations overshadow the small changes between the steady and turbulent
wind simulations and indicate that turbulence does not truly affect the rotor power and
torque. The number of blades should be taken into account when one designs a VAWT
as the variation of loads is highly dependent of the total number of blades. The mean
and STD values of the rotor torque and speed as a function of wind speed are plotted
in figure 3.4 on page 41. The torque for the VAWT is increased until the cut-out speed
in contrast to the HAWT which after rated remains constant. This happens because the
power regulation in the VAWT is done varying the rotor speed, where at high winds the
rotor speed is decreased in order to confine the power output and as a result the torque
increases.
7000
Mean VAWT
Mean HAWT
STD VAWT
STD HAWT

6000

Rotor Power [kW]

5000

4000

3000

2000

1000

0
0

10

15

20

25

Wind Speed [m/s]

Figure 3.1: Rotor power as a function of wind speed for the vertical and the horizontal axis
wind turbines.

Next the mean and STD values of the blade roots and turbine base bottom internal
bending moment magnitudes (equation 3.3 on page 41) are compared in figure 3.5 on
page 41. It can be seen that the mean values of the base bottom bending moment for
the VAWT are close with the ones of the HAWT at low wind speeds but deviate at high
wind speeds. This is because the thrust on HAWT is decreased after rated speed due
to pitch of the blades and consequently the base bottom moments are decreased. In

40

Design Load Cases

12

300

10

250

200

150

100

50

0
200
10 4
2

Rotor Torque [kNm]

350

205

210

215

220

225

230

235

0
240
400

Torque
Azimuth

1.5

300

200

0.5

100

0
200

Rotor Azimuth [deg]

400
Power
Azimuth

14

205

210

215

220

225

230

235

Rotor Azimuth [deg]

Rotor Power [MW]

16

0
240

Time [sec]

Rotor Power [MW]

16
14

400
Power
Azimuth 350

12

300

10

250

200

150

100

50

0
200
10 4
2

205

210

215

220

225

230

235

Rotor Azimuth [deg]

Figure 3.2: Time series of VAWT rotor power and torque at 13.5 m/s wind speed simulations
including turbulence.

0
240
400

300

200

0.5

100

Rotor Torque [kNm]

1.5

0
200

205

210

215

220

225

230

235

Rotor Azimuth [deg]

Torque
Azimuth

0
240

Time [sec]

Figure 3.3: Time series of VAWT rotor power and torque at 13.5 m/s wind speed simulations
without turbulence.

addition, there is an abrupt peak at 15.5 m/s wind speed for the VAWT results where
the rotor mean speed is 0.575 rad/s and is close to the crossing of the 1st tower natural
frequency with the 4P . It means that the selection of this rotor speed operation point
was not far enough from the crossing point where resonance happens. Nevertheless, in the
remaining analysis the same setting on the generator synchronous frequency at 15.5 m/s
wind speed is used and can be seen the influence on loads when the turbine operates
close to the resonance. The standard deviation for the VAWT is much higher along the
total wind speed operation range and the reason is the high variation of angle of attack
per revolution causing the aerodynamic loads to vary, as well. The fluctuation of angle

3.2 Normal Power Production

41

14000

1.4
Mean VAWT
Mean HAWT
STD VAWT
STD HAWT

10000

1.2

Rotor Speed [rad/s]

Rotor Torque [kNm]

12000

8000
6000
4000
2000

1
0.8
0.6
0.4
0.2

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.4: Rotor torque (left) and rotor speed (right) as a function of wind speed for vertical
and horizontal axis wind turbines.

of attack during every revolution for the blade section at equator height for 13.5 m/s
wind speed simulation without and with turbulence is plotted in figure 3.6 on the next
page. The effect of turbulence in angle of attack variation is small. On the other hand the
mean values of the HAWT blade root bending moments (right figure 3.5) are considerable
higher from cut-in up to 15 m/s where after that range the values for the VAWT become
larger. The standard deviations are of similar magnitude up to 11 m/s. After that point
the VAWT values are much higher while for the HAWT are almost constant. It should
be mentioned again that the HAWT has 3 blades while the VAWT two. It is also visible
the influence of centrifugal forces to the VAWT blade root bending moments at low wind
speeds and the importance to run the turbine close to the Troposkien rpm. Up to
7.5 m/s wind speed the turbines rotor speed is less than 0.5 rad/s, which is not close to
the Troposkien rpm (0.62 rad/s) but at 8.5 m/s the rotor speed becomes 0.6 rad/s and
the blade root bending moments decrease. The blade low root of the VAWT was selected
for the comparison with the HAWT blade root as its magnitude is higher from the upper
root at high wind speeds as it can be seen in figure 3.7 on the next page.
||Mx2 + My2 || =
10 4

Turbine Base ||M2x +M 2y || [kNm]

12

(3.3)

10000
Mean VAWT
Mean HAWT
STD VAWT
STD HAWT

9000
8000

Blade Root ||M2x +M 2y || [kNm]

14

q
Mx2 + My2

10

7000
6000
5000
4000
3000
2000

2
1000
0

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.5: Turbine base bottom bending moment (left) and blade root (low root for VAWT)
bending moment (right) as a function of wind speed for vertical and horizontal axis wind turbines.

42

Design Load Cases

30

360

240

180

-10

120

-20

60
205

210

215

220

225

230

235

0
240

30

360

20

300

10

240

180

-10

120

-20

60

-30
200

205

210

215

220

225

230

235

Rotor Azimuth [deg]

[deg]

10

-30
200

[deg]

300

Azimuth

Rotor Azimuth [deg]

20

0
240

Time [sec]

Figure 3.6: Time series of angle of attack at 13.5 m/s wind speed simulation without and
with turbulence for the blade section at equator height (VAWT).
7000
Mean bottom root
Mean upper root
STD bottom root
STD upper root

6000

||M 2x +M 2y || [kNm]

5000

4000

3000

2000

1000

0
0

10

15

20

25

Wind Speed [m/s]

Figure 3.7: Mean and STD of blade upper and low root bending moments as a function of wind
speed for the vertical axis wind turbine.

3.2.1

Ultimate Load Analysis

In this paragraph the ultimate loads for 50 year recurrence period winds (averaged over
three seconds) are calculated through the extrapolation of maximum values for normal
operation of wind turbine with NTM wind. No further safety factors are applied in
the analysis which corresponds to the DLC 1.1 of the international standard. At least
six turbulent seed realisations are simulated at every wind speed and case as proposed
in the standard for robust statistical analysis, and different methods are applied for the
extrapolation. The first is proposed by the GL guidelines where the ultimate loads for the
50 year return period are assumed that do not exceed the maximum value multiplied by
a factor of 1.35 [21]. The second method uses the Gumbel distribution for the statistical
extrapolation and proposed by the current standard. The extreme values which used for
the analysis are found from the time series of the signals applying two methods. The
peak over threshold (POT) and the global maxima (GM). The POT proposed as the

3.2 Normal Power Production

43

most appropriate for wind turbine load extrapolation [51, 66] with a threshold equal to
the mean value plus 1.4 times the standard deviation of the load. It is worthy to note that
the above references refer explicitly to HAWTs, thus the results for the VAWT should be
assessed carefully.
In figures 3.8 and 3.9 a comparison is made between the HAWT and VAWT 50 year
extrapolated loads of the turbine base and blade roots bending moments as a function of
wind speed respectively. The extrapolation is done through the Gumbel distribution and
the points for the extrapolation are found with the POT method in this case. The base
root bending moments on the VAWT are around 3 times higher compared to the HAWT
except at low wind speeds up to 7.5 m/s. A way to reduce these high loads would be to
support the tower with guy wires, a solution that was widely applied in VAWTs in the
past. It is recalled that at 15.5 m/s wind speed the turbine operates close to resonance
this is why there is the peak in the load, but in practice this operational rotor speed
should be avoided.

Base Bottom ||M 2x +M 2y || [kNm]

10 5
VAWT
HAWT

0
0

10

15

20

25

Wind Speed [m/s]

Figure 3.8: Extrapolated 50 year return period extreme values of turbine base bottom bending
moment as a function of wind speed for vertical and horizontal axis wind turbines.

On the other hand, the blade upper root bending moments of VAWT (figure 3.9 on the
following page) are less than the HAWT, while the loads at VAWT blade low root are
much higher after 11.5 m/s winds. It should be stressed that even though the blade low
root bending moments are higher for the VAWT the stresses can be kept relatively low by
strengthening that section with more material. This is not expected to be complicated as
the blade roots of the VAWT are fixed unlike the HAWT that have to rotate relatively to
the hub in order to pitch and extra weight at that region is not desirable. Another way
could be to add support struts very close to the roots. Two pictures from VAWT designs
that adopted such solutions in the past are shown in figure 3.10 on the next page.
As was noticed above the extrapolation methods used are based on HAWTs studies and
experience. Furthermore, in reference [50] it is concluded that the turbulence level affects
significantly the distribution of maximum loads when extrapolation methods are used for
HAWT load predictions. Next a comparison is made for the VAWT between the maximum
values as extracted from the simulations for the different turbulence seeds (black dots),
the extrapolated 50 year return period loads using the Gumbel distribution with POT

44

Design Load Cases

Blade Root ||M2x +M 2y || [kNm]

10 4
VAWT low-root
VAWT upper-root
HAWT

0
0

10

15

20

25

Wind Speed [m/s]

Figure 3.9: Extrapolated 50 year return period extreme values of blade root bending moment as
a function of wind speed for vertical and horizontal axis wind turbines.

Figure 3.10: Support struts and strengthened blade root solutions used in the past in order to
reduce the stresses of the VAWT blade roots [22, 56].

(red crosses) and GM methods (black squares), the GL guidelines (blue diamonds) and
the maximum loads from the equivalent simulations without turbulence (red circles).
Twelve different seeds per wind speed are used in this specific analysis for more reliable
results. The turbine base root bending moments Mx and My are plotted in figure 3.11
on the facing page where it can be seen that the 50 year extrapolated loads with GM
method are almost everywhere higher compared to the extrapolation with POT method
and the GL guidelines. The same trend is followed for the blade low root flatwise Mx and
edgewise My moments (figure 3.12 on the next page). The maximum values extracted
from the different turbulence seeds have small dispersion. The ultimate loads predicted
by the Gumbel extrapolation with POT method and the GL guidelines are very close.
Furthermore, the maximum loads extracted with turbulence are around 1.5 times higher
compared to the ones without turbulence.
In addition, the time series of the VAWT base bottom and blade low root bending moments with and without turbulence are plotted for 2 different wind speeds in figures 3.13
and 3.14 on page 46 for 13.5 m/s, and in appendix D on page 115 for 7.5 m/s (figures D.2
and D.3). It can be seen that the loads are relatively deterministic even with turbulent
wind. Next an analysis of the different load sources is conducted and is identified which

3.2 Normal Power Production

10 5

Bending Moment |Mx | [kNm]

45

Turbulence seeds (Max. values)


Without turbulence (Max. Values)
50-year extrapol. POT method
GL guidelines (Max. Values x 1.35)
50-year extrapol. GM method

Bending Moment |My | [kNm]

0
10 5

10

10

15

20

25

15

20

25

Wind Speed [m/s]

Figure 3.11: VAWT maximum base bottom moments (Mx and My ) from simulations with and
without turbulence and 50 year extrapolated values.

Bending Moment |Mx | [kNm]

10 4
Turbulence seeds (Max. values)
Without turbulence (Max. values)
50-year extrapol. POT method
GL guidelines (Max .values x 1.35)
50-year extrapol. GM method

Bending Moment |My | [kNm]

0
10 4

10

10

15

20

25

15

20

25

Wind Speed [m/s]

Figure 3.12: VAWT maximum blade low root flatwise Mx and edgewise My moments from
simulations with and without turbulence and 50 year extrapolated values.

blade loads are affected more from turbulence. The blade under operation is exposed to
the loads given below.

46

Design Load Cases

Gravitational loads,
Centrifugal loads and
Aerodynamic loads
The first two create a steady loading in the wind turbine blade as it rotates to a given
speed. The aerodynamic loads can be split to harmonic due to blade rotation and random
due to turbulence. Since the variation of angle of attack and relative velocity are inherently
high at every revolution of VAWT (see for example figure A.3 on page 101 and figure A.4
on page 102 in appendix A) the harmonic aerodynamic loading is much larger compared to
the random aerodynamic loading due to turbulence. This is also evident in the time series
from the highly deterministic load response. Power spectra density results for the turbine
base bottom and blade low root bending moments at 13.5 m/s wind speed operation are
plotted in figures 3.15 and 3.16 on page 47. The first 3 peaks in the turbine base moments
correspond to the 2P, 4P and 6P respectively. The energy content in 2P, which is the
highest, is exactly the same for the simulation results with and without the turbulence.
The same holds for the 1P, 2P and 3P frequencies which correspond to the first 3 peaks
in the blade PSD plots. Thus, it can be said that the blade aerodynamic loads are highly
deterministic for VAWTs even under turbulent wind conditions. This leads to highly
deterministic internal loads for the rotor the shaft, the gearbox, and the turbine base.
Regarding the IEC 61400, a further analysis on this topic is proposed. If the deterministic
nature of the loads could be demonstrated and extended to other VAWT types, some kind
of reduction to the safety factors could eventually be applied. In the following paragraph
a sensitivity analysis is performed for different turbulence intensities for the VAWT.

Mx [kNm]

10 5
With turbulence

My [kNm]

-2
200
10 5
4

Wind Speed [m/s]

Without turbulence

300

400

500

600

700

800

300

400

500

600

700

800

300

400

500

600

700

800

2
0
-2
200
25
20
15
10
5
200

Time [sec]

Figure 3.13: Time series of VAWT base bottom bending moments (Mx and My ) for 13.5 m/s
mean wind speed with and without turbulence.

3.2 Normal Power Production

47

Mx [kNm]

5000
0
-5000
-10000
With turbulence

Wind Speed [m/s]

My [kNm]

-15000
200
10 4
2

Without turbulence

300

400

500

600

700

800

300

400

500

600

700

800

300

400

500

600

700

800

1
0
-1
200
25
20
15
10
5
200

Time [sec]

Figure 3.14: Time series of VAWT blade low root flatwise Mx and edgewise My moments for
13.5 m/s mean wind speed with and without turbulence.
12

Power Spectra Density of My [ log( kNm2 s ) ]

Power Spectra Density of Mx [ log( kNm2 s ) ]

11
Without turbulence
With turbulence

10
9
8
7
6
5
4

11
10

9
8
7
6
5
4
3

Frequency [Hz]

Frequency [Hz]

Figure 3.15: Turbine base bottom bending moment spectra at 13.5 m/s.
9

Power Spectra Density of My [ log( kNm2 s ) ]

Power Spectra Density of Mx [ log( kNm2 s ) ]

9
Without turbulence
With turbulence

8
7
6
5
4
3
2
1
0

8
7
6
5
4
3
2
1
0

Frequency [Hz]

Frequency [Hz]

Figure 3.16: Blade low root flatwise Mx and edgewise My bending moment spectra at 13.5 m/s.

48

Design Load Cases

Influence of Turbulence
The influence of turbulence level in the extreme loads (extrapolated 50 year return period)
for the VAWT is investigated through the simulation of the three turbulence categories
of the standard A,B and C with Iref equal to 0.16, 0.14 and 0.12 [-] respectively.
The 50 year return period extrapolated base bottom and blade low root bending moments
for different turbulence intensities are plotted in figures 3.17 and 3.18. As the turbulence
increases the loads increase but the difference is not so large. The Gumbel distribution was
applied for the statistical extrapolation and the selected points for the analysis are picked
up with the POT method. It is suggested that for VAWTs two turbulence categories
instead of the three proposed by the current standard suffice for the analysis. Again the
peak at 15.5 m/s wind speed for the turbine bottom bending moments is due to operation
close to a resonance frequency as described on page 39.
6

10 5

10 5

Iref = 0.16 [-]


Iref = 0.14 [-]

Bending Moment |My | [kNm]

Bending Moment |Mx | [kNm]

Iref = 0.12 [-]


4

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.17: Extrapolated 50 year return period extreme values of the VAWT base bottom
fore-aft (Mx ) and side-side (My ) bending moments as a function of wind speed for different
turbulence levels Iref .
10 4

4.5
Iref = 0.16 [-]

Bending Moment |Mx | [kNm]

10 4

Iref = 0.14 [-]

3.5

Bending Moment |My | [kNm]

4.5

Iref = 0.12 [-]

3
2.5
2
1.5
1
0.5

3.5
3
2.5
2
1.5
1
0.5

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.18: Extrapolated 50 year return period extreme values of the VAWT blade low root
flatwise (Mx ) and edgewise (My ) bending moments as a function of wind speed for different
turbulence levels Iref .

3.2 Normal Power Production

49

Influence of Wind Shear


The influence of wind shear in the extreme loads (extrapolated 50 year return period)
for the VAWT is also investigated through the simulation of three different wind shear
exponents a equal to 0.20, 0.17 and 0.14 [-]. The latter is proposed in the IEC 61400-3
standard [31] for offshore wind turbines. The base bottom and blade low root bending
moments are plotted in figures 3.19 and 3.20. The results from all the cases reveal that
the turbine is not affected by moderate changes of the wind shear. This can be explained
from the position of the blades during rotation. Always a blade station remains at the
same level from the ground during rotation on a given mean rotor speed, while on a
HAWT a blade station during every revolution sees the wind at considerably different
heights and due to wind shear also different wind speeds.
10 5

10 5

5
a = 0.20 [-]
a = 0.17 [-]
a = 0.14 [-]

4.5

Bending Moment |My | [kNm]

Bending Moment |Mx | [kNm]

4.5

3.5
3
2.5
2
1.5
1
0.5

4
3.5
3
2.5
2
1.5
1
0.5

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.19: Extrapolated 50 year return period extreme values of the VAWT base bottom Mx
and My bending moments as a function of wind speed for different wind shear exponents a.
10 4

4.5
a = 0.20 [-]
a = 0.17 [-]
a = 0.14 [-]

Bending Moment |Mx | [kNm]

4
3.5

10 4

Bending Moment |My | [kNm]

4.5

3
2.5
2
1.5
1
0.5

3.5
3
2.5
2
1.5
1
0.5

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.20: Extrapolated 50 year return period extreme values of the VAWT blade low root
flatwise (Mx ) and edgewise (My ) bending moments as a function of wind speed for different wind
shear exponents a.

50

Design Load Cases

Generator Slip Sensitivity Analysis

Another factor that may affect the ultimate loads during power production is the generator
slip. Higher values of slip make the generator torque curve less steep. This means that
applying a higher torque than the actual at the equilibrium point will accelerate the rotor
easier compared with a generator with a lower slip. Because of the high variation of rotor
torque especially in a 2-bladed VAWT this could mitigate the loads through the possibility
of the rotor to accelerate/decelerate acting as a flywheel. Ten minute simulations are
performed with six seeds per wind speed from cut-in to cut-out for various generator
slips. Then the maximum loads on different parts of turbine are compared with the
default loads where the slip is 0.8%. Extrapolation to 50 year recurrence period is not
performed in order to see the net infuence on loads.
In figure 3.21 are plotted the turbine base bottom and blade low root bending moments
left and right respectively, for slips of 0.8, 4 and 8%. At low wind speeds the difference in
loads is negligible while for higher wind velocities there is a small reduction in loads for
the blade root when the slip is increased. For the turbine bottom base moment between
9.5 to 16.5 m/s the loads decreased but then increase again for higher slips. As a result,
the generator slip could be beneficial for some load reduction in the blades. On the other
hand, the default value of 0.8% that is based the analysis of all other cases in the standard
is not expected to alter the general conclusions, since the influence of the slip is deemed
small.

10 5

4
(s = 0.8%)
(s = 4.0%)
(s = 8.0%)

Bending Moment ||M2x +M 2y || [kNm]

3.5

10 4

3.5

2.5

2.5

1.5

1.5

0.5

0.5

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

Wind Speed [m/s]

Figure 3.21: Maximum base bottom (left) and blade low root (right) bending moments as a
function of wind speed for different generator slip values (Normal power production).

25

3.2 Normal Power Production

3.2.2

51

Fatigue Analysis

Failure on a turbine can also occur from the fluctuating loads on the components, which
cause fatigue during normal power production. In wind turbines during 20 years of
operation the number of load cycles exceeds the 106 [-]. The equivalent 1 Hz fatigue loads
are calculated from the 10 min simulations using the Rainflow counting method. The
analysis corresponds to the DLC 1.2 of IEC 61400-1.
A comparison between the HAWT and the VAWT is conducted, as well as sensitivity
analysis for the case of VAWT. A Wohler exponent m equal to 3 [-] is used for the
tower which is constructed from steel. For the composite blades is assumed an exponent
m = 10 [-] as proposed in [11, p. 530].
The equivalent 1 Hz fatigue loads for the turbine base bottom bending moments Mx and
My are plotted in figure 3.22. A large difference is observed for both bending moments
between the vertical and horizontal wind turbines, where the loads for the VAWT are
much larger. This happens because the load fluctuations for every revolution are large.
In figure 3.23 on the next page the equivalent 1 Hz fatigue blade root bending moments are
compared. The VAWT low root loads are presented as their magnitude is larger compared
to the upper root (see appendix D, figure D.4 on page 117). The Mx corresponds to
flatwise while the My to edgewise moment for the VAWT. For the HAWT Mx and My
are the blade flapwise and edgewise moments respectively. The equivalent loads between
the HAWT and the VAWT have similar values up to around 13 m/s wind speeds. In
the next paragraphs sensitivity analysis is conducted regarding the turbulence and wind
shear influence on the fatigue equivalent 1 Hz loads.
2

10 5

2
1.8

VAWT
HAWT

1.6

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.8

10 5

1.4
1.2
1
0.8
0.6
0.4
0.2

1.6
1.4
1.2
1
0.8
0.6
0.4
0.2

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.22: Equivalent 1 Hz fatigue base bottom bending moments Mx and My as a function
of wind speed for vertical and horizontal axis wind turbines.

52

Design Load Cases

10 4

10 4
VAWT
HAWT

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.5

0.5

1.5

0.5

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.23: Equivalent 1 Hz fatigue blade root bending moments Mx and My as a function of
wind speed for vertical and horizontal axis wind turbines.The low root loads for the VAWT are
compared.

Influence of Turbulence
The influence of turbulence to the 1 Hz equivalent fatigue loads for the VAWT is studied below. The results are extracted from the simulations of three different turbulence
intensities and 6 seeds at every wind speed bin. The loads for the base bottom and blade
low root are given in figures 3.24 and 3.25. For both the low blade root and base bottom
bending moments a small increase in loads occurs as the turbulence level is increased.
Furthermore, a comparison is made between the loads with turbulence (Iref = 0.14 [-])
and without for both the vertical and horizontal axis wind turbines. The plots can be
found in appendix D in figures D.5 and D.6 on page 117. It can be seen that the fatigue
loads for the VAWT does not differ so much (as in the case of the blade edgewise bending
moment in HAWT) which implies that are not highly dependent from turbulence.
2

10 5

2
Iref = 0.16 [-]

1.8

10 5

1.8

1.6

Iref = 0.12 [-]

Bending Moment M y [kNm]

Bending Moment M x [kNm]

Iref = 0.14 [-]

1.4
1.2
1
0.8
0.6

1.6
1.4
1.2
1
0.8
0.6

0.4

0.4

0.2

0.2

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.24: VAWT equivalent 1 Hz fatigue base bottom Mx and My bending moments as a
function of wind speed for different turbulence levels.

3.2 Normal Power Production

53

10 4

10 4

Iref = 0.16 [-]

Iref = 0.14 [-]

Bending Moment M y [kNm]

Bending Moment M x [kNm]

Iref = 0.12 [-]


1.5

0.5

1.5

0.5

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.25: VAWT equivalent 1 Hz fatigue blade root Mx and My bending moments as a
function of wind speed for different turbulence levels.

Influence of Wind Shear


A sensitivity analysis like the previous paragraph is performed on the influence of wind
shear in the 1 Hz equivalent fatigue loads for the vertical axis wind turbine. In figures 3.26
and 3.27 are illustrated the results for the base bottom and blade low root moments. It
can be seen that the loads are almost constant and independent of the wind shear at a
given wind speed for both the blade roots and turbine base bottom. The same trend was
obtained in the influence of wind shear on the extrapolated 50 year ultimate loads.
2

10 5

2
a = 0.20 [-]
a = 0.17 [-]
a = 0.14 [-]

1.6

1.8

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.8

10 5

1.4
1.2
1
0.8
0.6

1.6
1.4
1.2
1
0.8
0.6

0.4

0.4

0.2

0.2

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.26: VAWT equivalent 1 Hz fatigue base bottom Mx and My bending moments as a
function of wind speed for different shear exponents.

54

Design Load Cases

10 4

10 4

a = 0.20 [-]
a = 0.17 [-]
a = 0.14 [-]

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.5

0.5

1.5

0.5

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.27: VAWT equivalent 1 Hz fatigue blade low root Mx and My bending moments as a
function of wind speed for different shear exponents.

3.3

Ultimate Loads under Extreme turbulence

The IEC 61400-1 under the normal power production case proposes the calculation of
ultimate loads applying an Extreme Turbulence Model ETM directly (DLC 1.3), without
the need of statistical extrapolation as in the DLC 1.1. Then, if the computed loads from
DLC 1.1 simulations are smaller than the DLC 1.3 results the latter should be considered
as design values. The standard deviation 1 of the ETM is given by the equation 3.4,
where c is a constant equal to 2 m/s. According to the standard, if the extrapolated loads
from DLC 1.1 are higher the c factor may be increased up to the point where the resulting
loads from DLC 1.3 to become larger than the DLC 1.1 extrapolated values. This actually
substitutes the need of extrapolation of the DLC 1.1 results for the prediction of extreme
50 year return period loads by increasing the c factor. It could be said that this model is
calibrated for HAWTs to give the extreme values in a simpler way than the extrapolation
methods, thus most probably this model should be re-tuned for the case of VAWTs.
For this study the default value of c = 2 m/s and a second value of c = 3 m/s are
used to calculate and compare the extreme maximum loads with the extrapolated results
of subsection 3.2.1. In figures 3.28 and 3.29 on the next page are compared the results
from the simulations. The extrapolated values from the DLC 1.1 are always larger at
high wind speeds when compared with the results from DLC 1.3 even with the increased
c = 3 m/s. This holds for the base bottom and blade low root bending moments. As
a result, can be concluded that the estimated ultimate loads from the 50 year statistical
extrapolation methods may be overestimated in the case of VAWT or that the ETM needs
to be readjusted.





Vave
Vhub
1 = c Iref 0.072
+3
4 + 10
c
c

(3.4)

3.4 Extreme Wind Shear Case

55

10 5
5

DLC 1.1 50-years


DLC 1.3 c = 2 m/s
DLC 1.3 c = 3 m/s

Bending Moment |My | [kNm]

4.5

Bending Moment |Mx | [kNm]

10 5

4.5

4
3.5
3
2.5
2
1.5

3.5
3
2.5
2
1.5
1

0.5

0.5
0

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.28: Ultimate (50 year return period) base bottom Mx and My bending moments as a
function of wind speed. Comparison of DLC 1.3 with DLC 1.1.
10 4

Bending Moment |Mx | [kNm]

4.5

3.5

10 4

DLC 1.1 50-years


DLC 1.3 c = 2 m/s
DLC 1.3 c = 3 m/s

Bending Moment |My | [kNm]

4.5

3
2.5
2
1.5
1
0.5

3.5
3
2.5
2
1.5
1
0.5

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.29: Ultimate (50 year return period) blade low root Mx and My bending moments as
a function of wind speed. Comparison of DLC 1.3 with DLC 1.1.

3.4

Extreme Wind Shear Case

One extreme event that is proposed by the standard as potentially critical during the
normal operation of the wind turbine is the transient extreme wind shear (EWS), which
corresponds to DLC 1.5. It refers to vertical and horizontal wind shear (wind veer). The
positive or negative transient vertical and horizontal EWS are given by the equations 3.5
and 3.6 on the next page. The transient time T is set to 12 sec and is a parameter
equal to 6.4 [-]. The turbulence standard deviation 1 is given by equation 3.1 on page 38.
D is the rotor diameter where for the VAWT is set equal to the wind turbine diameter
at equator height and 1 is the longitudinal turbulence scale parameter at hub height
(equation 3.10 on page 58). z and y refer to the vertical and horizontal position in global

56

Design Load Cases

coordinates.

V (z, t) =

0.2


hub zhub

zzhub
D

V (y, z, t) =


 0.25 
2.5 + 0.21 D1
1 cos
Vhub

0.2


hub zhub

y
D

0.2

zhub

Vhub

0.2

zhub



, 0tT

t>T
(3.5)


 0.25 
2.5 + 0.21 D1
1 cos

2t
T

2t
T



, 0tT

t>T
(3.6)

The extreme loads from these simulations are compared with the 50 year return period
values as were extrapolated from the normal power production case using the Gumbel
distribution (see subsection 3.2.1 on page 42). The turbine base bottom bending moments
Mx and My are plotted in figure 3.30. It can be seen that the extrapolated loads (DLC 1.1)
are higher throughout the operational wind speed range. The same holds for the loads
in the blade roots. In figure 3.31 on the facing page are depicted the low root bending
moments, since their magnitude is higher than the upper root loads.
The results from the vertical and the horizontal shears are almost the same, thus the
turbine is affected slightly from transient extreme wind shears. Furthermore, due to
much larger loads arising from the DLC 1.1 case can be concluded that this design load
case is practically redundant. It is recalled that a sensitivity analysis regarding the wind
shear under normal power production, which was done in paragraph 3.2.1 on page 49, has
indicated that moderate changes in wind shear does not influence the loads on VAWT.
10 5

4.5

4.5

Bending Moment |My | [kNm]

Bending Moment |Mx | [kNm]

10 5

4
3.5
3
2.5
2
1.5

4
3.5
3
2.5
2
1.5

0.5

0.5

50-year extrapol. POT method


Extreme wind shear
Extreme wind veer

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.30: Comparison of ultimate loads emerging from vertical and horizontal EWS with
the extrapolated 50 year return period maximum values (DLC 1.1) for the turbine base bottom
bending moment as a function of wind speed.

3.5 Extreme Operating Gust Case

57

10 4

3.5

3.5

Bending Moment |My | [kNm]

Bending Moment |Mx | [kNm]

10 4

3
2.5
2
1.5
1
0.5

50-year extrapol. POT method


Extreme wind shear
Extreme wind veer

3
2.5
2
1.5
1
0.5

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.31: Comparison of ultimate loads emerging from vertical and horizontal EWS with
the extrapolated 50 year return period maximum values (DLC 1.1) for the blade low root bending
moment as a function of wind speed.

3.5

Extreme Operating Gust Case

In the standard the power production design situation is also combined with the occurrence of fault in the control or electrical systems (DLC 2.x). Three DLCs refer to normal
wind conditions while one refers to the worst extreme operating gust (EOG) transient
event expected in 50 year recurrence period (DLC 2.3). This is combined with an external or internal electrical fault triggered by the EOG. It is particularly important in
pitch regulated HAWTs when operating around rated wind velocities where loads are
high. The turbine is not able to pitch the blades quickly and avoid the even higher loads
emerging from the passing of the gust. This also results in a generator over-speed which
in turn activates the emergency stop where the loads are severe. In this study the EOG
is applied and the ultimate loads are extracted without taking into account any electrical
fault. The extreme wind speed change is modelled as a deterministic sinusoidal signal
(Mexican hat) even though in reality has a more arbitrary shape.
The gust velocity V (z, t) as a function of height z and time t is given in equation 3.7.
The gust value Vgust (equation 3.8 on the next page) refers to the largest magnitude
that is expected on 50 year return period. The time of the transient phenomenon is set
T=10.5 sec according to the IEC. The turbulence longitudinal standard deviation 1 is
given by equation 3.1 on page 38. D is the rotor diameter and 1 is the longitudinal
turbulence scale parameter at hub height (equation 3.10 on the next page). In addition,
for the calculation of Vgust , Ve1 wind is used which refers to the extreme wind speed of
1 year recurrence period averaged over 3 sec (equation 3.9 on the following page). z is
the position expressed on the global coordinate system.


3t
V
(z)

0.37V
sin
1 cos
gust

V (z, t) =

0.2

Vhub zhub

2t
T



f or 0 t T
(3.7)
f or t > T

58

Design Load Cases

Vgust

1
 
= min 1.35 (Ve1 Vhub ) ; 3.3

1 + 0.1 D

(3.8)


Ve1 (z) = 1.12Vref

1 =

0.11

zhub

0.7z f or z 6 60 m
42 m f or z > 60 m

(3.9)

(3.10)

IEC proposes to apply the gust around rated wind speed Vr 2 m/s and at cut-out where
it is expected to appear high loads for HAWTs, but here the simulations are conducted
at various wind speeds including the aforementioned speeds in order to identify the most
critical case. Furthermore, the Taylor hypothesis is applied as it is suggested in [40, p. 9].
The simulations for each wind speed are performed applying the gust peak at six different
time stamps which correspond to different rotor orientations. This is done in order to
analyse the influence of the position of the frontal passage of the gust with respect to the
blade relative positions.
In figure 3.32 on the next page are plotted the time series of the turbine base bottom
bending moments Mx and My when the gust passes through the rotor for three of the
six different simulations. Before and after the transient the wind velocity is set to rated.
It can be seen that the loads are dependent on the orientation of the rotor both in x
and y directions. The blade-1 low root bending moments Mx and My are depicted in
figure 3.33 on page 60 where changes in the loads are apparent especially for the edgewise
moment My (see sub-plot on 3rd row, right). The rotor speed and torque are presented
in figure 3.34 on page 61. The torque follows the variations in the wind depending on
the relative position of the blades, while the rotor speed changes slightly due to the large
inertia of the rotor.
The time series of the turbine bottom and blade-1 low root bending moments for a mean
wind of 19.5 m/s and a peak due to gust 25 m/s are given in appendix D.2 on page 118.
The loads vary depending on the rotor orientation and the frontal passage location of the
gust as in the case of 13.5 m/s mean wind speed. The maximum of turbine base bottom
and blade-1 low root bending moments due to the gust are compared with the DLC 1.1
(normal turbulence model) equivalent maximums (not the extrapolated to 50 year return
period) at tables 3.4 and 3.5 on next page at 4 different mean wind speeds (8.5, 13.5,
19.5 and 24.5 m/s). It should be mentioned that when a maximum bending moment
appears in one direction it is not necessarily maximum in the other direction at the same
time. This is why the maximum magnitude of both bending moments is also given. The
maximum loads from DLC 1.1 are for all wind speed levels higher than the loads emerging
from the gust. It can be concluded that the gust passage does not lead to very high loads
compared to the ones with the NTM model for the vertical axis wind turbine.

3.5 Extreme Operating Gust Case

59

V hub

15

[m/s]

230

240

250

10
270

260

10 5

-1
220

230

240

250

10
270

260

10 5

20

20

15

15

[m/s]

[m/s]

[kNm]

-1
220

230

240

250

10
270

260

10 5

-1
220

250

10
270

260

20

[m/s]

15

[kNm]

15

240

250

10
270

260

-1
220

Rotor Azimuth Angle [deg]

230

300

200

100

0
220

240

10 5

20

-1
220

230

[m/s]

[kNm]

15

-1
220

[kNm]

V hub

Rotor Azimuth Angle [deg]

20
My

[kNm]

[kNm]

10 5

20
Mx

[m/s]

10 5
2

230

240

250

260

230

240

250

260

10
270

230

240

250

260

270

300

200

100

0
220

270

Time [sec]

Time [sec]

Figure 3.32: Time series of VAWT base bottom bending moments Mx (left) and My (right)
during the EOG at rated wind speed. The gust frontal passage passes the rotor plane at 3
different time stamps (corresponding to different rotor orientations), each one represented at the
1st , 2nd and 3rd row of sub-plots.

Table 3.4: Maximum wind turbine base bottom bending moments. Comparison of loads
emerging from gust (EOG) and normal power production at different wind speeds.

Mean wind speed [m/s]


DLC
Max. BM Mx [104 kNm]
Max. BM My [104 kNm]
Max. ||Mx2 + My2 || [104 kNm]

8.5
EOG
12.68
9.44
13.60

NTM
18.92
15.08
17.77

13.5
EOG
20.24
16.32
21.68

NTM
25.49
24.06
24.11

19.5
EOG
13.84
17.68
21.57

NTM
22.27
24.96
26.92

24.5
EOG
18.89
18.63
24.22

NTM
29.69
27.53
29.49

60

Design Load Cases

20

15

[m/s]

[kNm]

20
12000

V hub

[kNm]

Mx

-5000

-10000
220

My

V hub

7000
15

[m/s]

5000

2000

230

240

250

260

10
270

5000

-3000
220

230

240

250

10
270

260

20

20

-5000

-10000
220

7000
15

[m/s]

15

[m/s]

[kNm]

[kNm]

12000

2000

230

240

250

260

10
270

5000

-3000
220

230

240

250

10
270

260

20

20

-5000

230

240

250

260

10
270

-3000
220

300

200

100

0
220

15
2000

Rotor Azimuth Angle [deg]

Rotor Azimuth Angle [deg]

-10000
220

7000

[m/s]

15

[m/s]

[kNm]

[kNm]

12000

230

240

250

260

230

240

250

260

10
270

230

240

250

260

270

300

200

100

0
220

270

Time [sec]

Time [sec]

Figure 3.33: Time series of VAWT blade-1 low root bending moments Mx (left) and My
(right) during the EOG at rated wind speed. The gust frontal passage passes the rotor plane at 3
different time stamps (corresponding to different rotor orientations), each one represented at the
1st , 2nd and 3rd row of sub-plots.

Table 3.5: Maximum wind turbine blade-1 low root bending moments. Comparison of loads
emerging from gust (EOG) and normal power production at different wind speeds.

Mean wind speed [m/s]


DLC
Max. BM Mx [104 kNm]
Max. BM My [104 kNm]
Max. ||Mx2 + My2 || [104 kNm]

8.5
EOG
0.48
0.43
0.63

NTM
0.82
0.93
0.96

13.5
EOG
1.01
1.36
1.69

NTM
1.43
2.06
2.13

19.5
EOG
1.46
1.77
2.35

NTM
1.94
2.47
2.80

24.5
EOG
1.93
2.47
3.13

NTM
2.64
3.08
3.66

3.5 Extreme Operating Gust Case

240

250

260

0.65

[m/s]

15

230

20

10
270

V hub

Rotor torque

[kNm]

[rad/s]

0.625

0.6
220

10 4

20
V hub

Rotor speed

15

0
220

230

240

250

260

10 4

20

[m/s]

0.65

61

10
270
20

0.6
220

230

240

250

260

0.65

10
270

15

0
220

230

240

250

260

10 4

20

[m/s]

15

[m/s]

0.625

[kNm]

[rad/s]

10
270
20

Rotor Azimuth Angle [deg]

0.6
220

230

240

250

260

10
270

300

200

100

0
220

230

240

250

Time [sec]

260

270

15

0
220

230

240

250

260

10
270

230

240

250

260

270

[m/s]

[m/s]

15

[kNm]

0.625

Rotor Azimuth Angle [deg]

[rad/s]

300

200

100

0
220

Time [sec]

Figure 3.34: Time series of VAWT rotor speed (left) and rotor torque (right) during the EOG
at rated wind speed. The gust frontal passage passes the rotor plane at 3 different time stamps
(corresponding to different rotor orientations), each one represented at the 1st , 2nd and 3rd row
of sub-plots.

62

3.6

Design Load Cases

Extreme Wind Direction Change Case

In the beginning of this chapter was stated that cases referring to start up and normal
shut down are not investigated as the emerging loads will vary from the selected control
strategy. This also means that there is the possibility to adjust the turbine response in
order the loads to be below the desired limits. In the current standard under the start
up design situation there is the case of extreme direction case (EDC) wind condition.
Below an analysis is made applying the EDC wind transient phenomenon during normal
operation of wind turbine. This could indicate how critical such an extreme event is. The
magnitude of the extreme direction change e is given by equation equation 3.11. The
wind direction changes as a function of time according to equation 3.12. The duration
of the transient phenomenon T is 6 seconds. Simulations are set for every wind speed bin
from cut-in up to cut-out, considering positive as well as, negative values of e . The wind
speed follows the normal wind profile (equation 3.2 on page 38) including turbulence. The
total simulation time is set to 300 sec and at 200 sec the EDC takes place, in order to
avoid any initial transients at the beginning of the simulation.

e = 0.54arctan

(t) = e 1 cos

1


Vhub 1 + 0.1

t
T



D
1



f or t < 0
f or 0 t T
f or t > T

(3.11)

(3.12)

The bending moments on the turbine base bottom and the blade low root are compared
with the ultimate 50 year extrapolated values (DLC 1.1) in figures 3.35 and 3.36 on the
following page. The results reveal that the loads emerging from the extreme transient
wind direction change are much lower than the ultimate loads during 50 year return
period winds. The magnitude of the velocity during the wind direction change remains
the same and only the velocity components in x and y direction alter magnitudes. Since
the VAWT rotor is omnidirectional is almost unaffected from the quick change in wind
direction in contrast to a HAWT which would operate under skew inflow. The maximum
bending moments of EDC were also compared with the maximum loads during normal
power production without extrapolation and there are small differences due to the random
turbulence seeds (see appendix D.3, figure D.9 on page 119).

3.6 Extreme Wind Direction Change Case

63

10 5

4.5

4.5

Bending Moment |My | [kNm]

Bending Moment |Mx | [kNm]

10 5

3.5
3
2.5
2
1.5

3.5
3
2.5
2
1.5

0.5

0.5

50-year extrapol. POT method


Positive wind direction change
Negative wind direction change

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.35: Comparison of ultimate loads emerging from positive and negative EDC with the
extrapolated 50 year return period maximum values (DLC 1.1) for the turbine base bottom
bending moment as a function of wind speed.

10 4

4.5

3.5

3.5

Bending Moment |My | [kNm]

Bending Moment |Mx | [kNm]

4.5

3
2.5
2
1.5

50-year extrapol. POT method


Positive wind direction change
Negative wind direction change

3
2.5
2
1.5

0.5

0.5

10 4

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.36: Comparison of ultimate loads emerging from positive and negative EDC with the
extrapolated 50 year return period maximum values (DLC 1.1) for the blade low root bending
moment as a function of wind speed.

64

3.7

Design Load Cases

Emergency Shut Down

A VAWT rotor can be decelerated with a brake system under normal or emergency conditions by different ways. These include mechanical and aerodynamic brakes or deceleration
with generator, as well as blade pitching for a H-rotor. In this section the loads arising
from an emergency shut down of the Darrieus VAWT are investigated using only a mechanical brake system assuming that the aero-brakes if any, and the generator brake have
failed. This case is equivalent to DLC 5.1 of the IEC standard.
The normal turbulence model with Iref = 0.14 [-] is used in the simulations and the
turbine operates normally until the emergency shut down event. To avoid any transient
effects the mechanical brake is applied after the initial 200 seconds. In addition, it is
assumed that 0.5 sec before the activation of the mechanical brake, the generator torque
is lost for instance due to electrical grid failure. This small time delay was selected after
reviewing reference [33] where a delay of 0.1 sec was used for HAWT emergency shut
down simulations. The applied mechanical brake torque is modelled as a solution of a
first order ordinary differential equation in step response (equation 3.13). Tbmax is the
maximum braking torque and tc the brake time constant. The simulations are done at
various wind speeds including the rated and cut-out speeds and sensitivity analyses of the
two parameters of the brake system tc and Tbmax are performed. The latter is expressed
as multiple of the generator LSS torque at rated power (Mn ) by equation 3.14, where
is the rotor angular velocity in rad/s.


t
Tb = Tbmax 1 e tc

Mn =

5 [M W ]
[rad/s]

(3.13)

(3.14)

In figure 3.37 on the next page are plotted the rotor and brake torque (upper plot), the
turbine base bottom bending moments (Mx , My ) in the middle and the rotor azimuth
angle and speed down. The wind turbine operates at rated wind speed and at 220 sec the
mechanical brake is activated. Half second before the generator is deactivated therefore
the rotor accelerates for a short period as it can be seen in the lower plot. In this
simulation the maximum brake torque is 3 times the generator torque Mn (Tbmax = 3Mn )
and the brake time constant tc is set to 0.3 sec. The turbine decelerates to a rotor speed
of 10% of rated within 9 sec from the moment of brake activation (6 sec are needed if
Tbmax = 4Mn ). Regarding the turbine base bottom loads it is seen from the simulation
results that there is not any pick during the first seconds of brake activation but there is
a small fluctuation of loads even after 50 sec of brake activation due to dynamics of the
rotor.
From the same simulation are extracted the blade low root bending moments and the blade
station positions x and y at equator height (figure 3.38 on page 66). Similar behaviour for
the blade low root bending moments is observed as for the turbine base bottom, except
that the fluctuation of loads after the first 10 sec is more dominant for the out of plane
My bending moment. The blade station positions (x, y) at equator height show that the
blade does not deflect more during the deceleration phase.

3.7 Emergency Shut Down

1.5

1.5

0.5

0.5

210

220

230

240

250

260

0
280

270

10 5
2

10 5

Bending Moment M y [kNm]

Bending Moment M x [kNm]

Brake Torque [kNm]

Rotor Torque
Brake Torque

0
200

Mx
My

-1
200
360

Rotor Azimuth Angle [deg]

10 4
2.5

10 4

210

220

230

240

250

260

270

-1
280
0.8

Rotor Azimuth Angle


Rotor Speed
270

0.6

180

0.4

90

0.2

0
200

210

220

230

240

250

260

270

Rotor Speed [rad/s]

Rotor Torque [kNm]

2.5

65

0
280

Time [sec]

Figure 3.37: Time series of VAWT simulation during emergency shut down at 220 sec time
stamp. The turbine operates at 13.5 m/s turbulent wind. The rotor and brake torque are shown
in the upper plot, the turbine base bottom bending moments expressed in global coordinate system
in the middle and the rotor azimuth angle and speed down (Tbmax = 3Mn , tc = 0.3 sec).

The above brake configuration (maximum torque and time constant) and wind conditions
simulation results reveal that the turbine base and blade ultimate loads do not increase
during the emergency stop. This could be expected for a 2 bladed VAWT since under
normal operation the torque ripple is high at every revolution, thus the opposing brake
torque does not increase the loads further.
The maximum clamping force that is needed to apply to the disc brake and the dissipated
energy during the stop are also calculated. The brake is assumed to be located on the
low speed shaft side in order to be able to stop the rotor even if a failure in the gearbox
happens. From Newtons 2nd law the braking torque Tb should be equal to the rotor
inertia Tj = J plus the aerodynamic torque Ta neglecting the friction torque in the
bearings (equation 3.15).
Tb = Tj + Ta

(3.15)

The disc brake radius and thickness used in Sandia 34 m VAWT were 2 m and 2.54 cm

66

Design Load Cases

10 4
1.5

10 4
Mx

0.5

0.5

-0.5
-1
200

-0.5

210

220

230

240

250

260

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.5

-1
280

270

62
x
y

Pozition y [m]

Pozition x [m]

2
61
1
60
0

210

220

230

240

250

260

270

59
280
0.8

Rotor Azimuth Angle


Rotor Speed
270

0.6

180

0.4

90

0.2

0
200

210

220

230

240

250

260

270

Rotor Speed [rad/s]

Rotor Azimuth Angle [deg]

-1
200
360

0
280

Time [sec]

Figure 3.38: Time series of VAWT simulation during emergency shut down at 220 sec time
stamp. The turbine operates at 13.5 m/s turbulent wind. The blade low root bending moments
expressed in local coordinate system are given in the upper plot, the blade station position at
equator height expressed in blade coordinate system in the middle and rotor azimuth angle and
speed down (Tbmax = 3Mn , tc = 0.3 sec).

respectively [72, p. 66]. Up-scaling the dimensions with the same scaling factor (sf =
3.54 [-]) as was defined in subsection 2.2.1 on page 15 results to a 3.5 m radius and 9 cm
thickness. However, the radius is relatively large and is reduced to rb = 2.5 m here.
Since the braking torque in this case is predefined, the braking force Fb for one caliper is
computed by equation 3.16 and the clamping force Fk by equation 3.17 on the facing page.
Fb is the tangential force between the brake disc and pads, while Fk in the normal force
that the caliper should apply to the disc in both sides. For a mean friction coefficient of
= 0.4 [-] according to reference [77] and the case described above, a maximum clamping
force of 12.09 MN is needed if one caliper is used. In addition, integration of the brake
power from the moment of activation up to the point where the rotor speed goes below
5% of rated results to an energy dissipation of 71 MJ.

Fb =

Tb
rb

(3.16)

3.7 Emergency Shut Down

67

Fk =

Fb
2

(3.17)

Sensitivity analysis of the maximum loads for the turbine base bottom and blade roots
for different wind speeds and brake parameters is conducted below. The maximum brake
torque is set to 3 and 4 times the Mn and the brake time constant to 0.3, 0.4 and
1 sec. In figures 3.39 and 3.40 are plotted the maximum turbine bottom and blade
low root bending moments. The base bottom moments in general increase as the brake
time constant increases, while for the blade root moments the correlation is opposite. In
addition, the loads emerging from the different brake torques are almost the same for the
turbine base bottom but for the blade roots the loads increase as the maximum brake
torque increases.

10 5

10 5

Bending Moment ||M2x +M 2y || [kNm]

tc = 0.3 sec
tc = 0.5 sec

2.5

2.5

tc = 1 sec

1.5

1.5

0.5

0.5

0
8.5

13.5

18.5

24.5

8.5

Wind Speed [m/s]

13.5

18.5

24.5

Wind Speed [m/s]

Figure 3.39: Maximum turbine base bottom bending moments for different brake time
constants tc and brake torques Tbmax . At the left Tbmax = 3Mn and at right Tbmax = 4Mn .
T max
=3Mn
b

10 4

tc = 0.3 sec

Bending Moment ||M2x +M 2y || [kNm]

2.5

T max
=4Mn
b

10 4
2.5

tc = 0.5 sec
tc = 1 sec
2

1.5

1.5

0.5

0.5

0
8.5

13.5

18.5

Wind Speed [m/s]

24.5

8.5

13.5

18.5

24.5

Wind Speed [m/s]

Figure 3.40: Maximum blade low root bending moments for different brake time constants tc
and maximum brake torques Tbmax . At the left Tbmax = 3Mn and at right Tbmax = 4Mn .

68

3.8

Design Load Cases

Parked Rotor

In this section the parked standing still or idling situations are investigated where the
turbine does not produce energy as the wind velocity exceeds the turbines cut-out speed.
These correspond to DLCs 6.1 up to 6.4 and 7.1 of the IEC 61400-1. The moments on
the turbine bottom and blade roots are analysed, as well as the deflection of the blade at
equator height. Relevant results are compared with the NREL HAWT which is simulated
in idling mode with zero and eight degrees yaw misalignment.
The standard proposes to use for the simulations a steady extreme wind model (EWM)
or a turbulent EWM. In the case of the steady EWM effects of resonant response must
be taken into account separately, thus it is selected to use the turbulent EWM in this
analysis. Two different recurrence periods are used depending on the DLC, 50 years and
1 year, where the corresponding 10 min average extreme wind speeds V50 and V1 are given
by equations 3.18 and 3.19 respectively. The turbulence intensity at hub height is set to
0.11 [-] and the wind shear exponent of the power law wind profile to 0.11 [-] for both
recurrence period cases.
V50 (z) = Vref (

z
zhub

)0.11

V1 (z) = 0.8V50 (z)

(3.18)

(3.19)

Where z is the height from the ground and zhub the hub height.

3.8.1

Parked Idling Rotor - Ultimate Load Analysis

The case of idling rotor is analysed first for the 50 year return period turbulent EWM
(DLC 6.1), where the rotor is set to rotate at a prescribed very low speed. Initially, the
rotor was set totally free but it accelerates beyond rated rotor speed without reaching any
equilibrium point. The results at two different idling rotor speeds equal to 5% and 10%
of the rated (0.62 rad/s) for the VAWT are examined and are compared with the HAWT.
At table 3.6 on the next page are summarized the maximum values of critical loads and
deflections.
The current standard proposes to examine also the case of extreme wind conditions with
1 year return period together with extreme yaw misalignment (DLC 6.3). In the case
of the VAWT which is insensitive of wind direction yaw misalignment is not applicable.
Furthermore, the loads emerging from the 1 year return period wind are not expected
to be higher than the loads emerging from the 50 year recurrence period extreme wind
conditions. Nevertheless, the simulations are done and the loads are compared with
the HAWT under 20 yaw misalignment. As it is seen from the results (table 3.7 on
the facing page) the maximum loads and blade deflections for the VAWT are less in
comparison with the results from the 50 year return period simulations (table 3.6 on the
next page). As a conclusion from the above comparison it is stated that DLC 6.3 is
redundant for the VAWT case.

3.8 Parked Rotor

69

Table 3.6: Loads and deflections of wind turbines with idling rotor under 50 year recurrence
period wind conditions.

Maximum absolute values

VAWT
Rated rotor
speed [%]
5
10

Tower bottom BM ||Mx2 + My2 || [kNm]


Rotor torque [kNm]
Blade low root BM ||Mx2 + My2 || [kNm]
Blade upper root BM ||Mx2 + My2 || [kNm]
Blade root BM ||Mx2 + My2 || [kNm]
Blade deflection x at equator height [m]
Blade deflection y at equator height [m]
Blade deflection z at equator height [m]

30.5x104
2.2x104
2.6x104
1.8x104
3.7
3.5
3.6

29.8x104
2.4x104
2.7x104
1.8x104
3.6
3.6
4.0

HAWT
Yaw
misalign. [deg]
0
8
6.2x104
0.6x104
0.6x104
-

8.0x104
0.6x104
1.0x104
-

The results between the VAWT and the HAWT tower bottom and blade root bending
moments show that the maximum values are lower for the HAWT. In should be noted
that this is due to the extra degree of freedom for the HAWT blades, in other words the
ability to pitch and reduce the aerodynamic loading when it is needed as in this case,
where the pitch angle is set to 90 .
Table 3.7: Loads and deflections of wind turbines with idling rotor under 1 year recurrence
period wind conditions.

Maximum absolute values

VAWT
Rated rotor
speed [%]
5
10

Tower bottom BM ||Mx2 + My2 || [kNm]


Rotor torque [kNm]
Blade low root BM ||Mx2 + My2 || [kNm]
Blade upper root BM ||Mx2 + My2 || [kNm]
Blade root BM ||Mx2 + My2 || [kNm]
Blade deflection x at equator height [m]
Blade deflection y at equator height [m]
Blade deflection z at equator height [m]

3.8.2

19.1x104
1.4x104
1.6x104
1.3x104
2.2
3.0
2.4

19.3x104
1.6x104
1.7x104
1.2x104
2.4
3.1
2.6

HAWT
Yaw
misalign. [deg]
-20
20
5.6x104
0.3x104
4
0.9x10
-

6.1x104
0.4x104
4
0.8x10
-

Parked Idling Rotor - Fatigue Analysis

Fatigue damage occurs during normal power production but also other turbine states
contribute and should be taken into account. The parked idling rotor design situation is
one of these. The 1 Hz equivalent fatigue loads are extracted for the same idling rotor
speeds as the ultimate load analysis above (5% and 10% of rated) and compared with the

70

Design Load Cases

one of normal power production. This case corresponds to DLC 6.4 of the standard and
is considered relevant for the VAWTs too.
In figure 3.41 and figure 3.42 the equivalent 1 Hz fatigue loads for the turbine base bottom
and blade low root bending moments Mx and My are plotted. The loads arising from
the idling mode are considerably lower in comparison with the normal power production
equivalent loads for both the blade low root and the turbine bottom. A small difference
in loads can also be seen between the two different idling rotor speeds, while the loads
increasing as the wind speed increases.

2.2

10 5

1.8

1.6

Bending Moment M y [kNm]

Bending Moment M x [kNm]

DLC 6.4 (Idling 10% of rated )

1.8

10 5

2.2
DLC 1.1 (Normal power production)
DLC 6.4 (Idling 5% of rated )

1.4
1.2
1
0.8
0.6

1.6
1.4
1.2
1
0.8
0.6

0.4

0.4

0.2

0.2

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.41: VAWT equivalent 1 Hz fatigue base bottom Mx and My bending moments as a
function of wind speed for different idling rotor speeds and comparison with DLC 1.1.

10 4

10 4
DLC 1.1 (Normal power production)
DLC 6.4 (Idling 5% of rated )

DLC 6.4 (Idling 10% of rated )

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.5

1.5

0.5

0.5

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure 3.42: VAWT equivalent 1 Hz fatigue blade root Mx and My bending moments as a
function of wind speed for different idling rotor speeds and comparison with DLC 1.1.

3.8 Parked Rotor

3.8.3

71

Parked Standing Still Rotor - Ultimate Load Analysis

In this scenario the vertical axis wind turbine rotor is locked and the same turbulent
extreme wind conditions as in the previous case are applied. These conditions could
represent the DLC 6.2 of the standard which assumes loss of electrical network connection
and thus may not be able the rotor to be in idling mode but standing still.
The rotor is set to different azimuthal angles from 0 to 170 with a step of 10 degrees
and the extreme values are extracted from the time series of the results. Due to 2-bladed
rotor, symmetry exist from 180 to 350 orientation angles thus no simulations are realised
on that half. The simulations run for 900 sec where the last 600 sec (10 minutes) are used
in the analysis. The wind direction is always along the global y coordinate. The rotor with
the blades orientation at = 0 is depicted in figure 3.43 where increases following the
clockwise direction. The same 50 and 1 year return period turbulent EWM as described
in previous sub-section of parked idling rotor are applied.

Figure 3.43: Rotor cross-section at equator height, depicting the orientation of the blades at
= 0 .

First, the loads on wind turbine for the more severe 50 year recurrence period wind
conditions are examined. The turbine maximum and standard deviation tower bottom
bending moments as a function of rotor orientation are depicted in figure 3.44 on the
next page. Abrupt increment in loads exists when the rotor orientation is not close to 0
(equivalent 180 ) and 90 (equivalent 270 ) degrees orientation. Similar trend is followed
by the blade low root bending moments seen in figure 3.45 on page 73.
The blade upper roots bending moments have smaller magnitudes, thus the low roots were
selected to be plotted. In appendix D.4, figure D.10 on page 120 is given the maximum
torsional moment that the brake mechanism should withstand at different rotor azimuthal
angles. Excluding the abrupt peaks which will be investigated next all other maximum
values are below 5x104 kNm. In addition, are shown the results from the simulations with
the same wind conditions but with the dynamic stall model deactivated (appendix D.4
on page 120, figures D.11 and D.12). The loads on the turbine base bottom and the blade

72

Design Load Cases

roots are of similar magnitude but with higher peaks at the rotor orientations where very
large peaks occur. This design situation reveals that depending of the rotor orientation
the loads can be severe.
The blade-1 low root flatwise (Mx ) and edgewise (My ) absolute mean and standard deviation are plotted in figure 3.46 on the next page. It can be seen that except the regions with
very high loads ( = 10 to 50 and 130 to 150 ) in the remaining rotor orientations the
mean flatwise moment is much higher compared to the edgewise due to the contribution
of the blade weight but also due to high aerodynamic drag when the rotor is around 90 .
The peaks on blade low root bending moments originate from the edgewise component.
The standard deviation is relatively high compared to the mean values and reveals that
the blade vibrates. Next, it is tried to identify what causes this behaviour which due to
high change of loads between, for instance, from 10 to 20 rotor orientation probably an
instability occurs.

10 5
Maximum
STD

Bending Moment ||M 2x +M 2y || [kNm]

8
7
6
5
4
3
2
1
0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.44: Maximum and standard deviation of the turbine bottom bending moments
expressed at the global coordinate system. The wind turbine is parked in standing still mode at
different rotor orientations and exerted to 50 year recurrence period extreme winds.

As this happens at different orientations, where the blades change position relative to wind
direction and probably become unstable and in order to isolate the response and identify
the origin, the whole rotor is set as rigid body except the blades and the simulations are
repeated. Moreover, the same mean wind speed is applied but without turbulence. In
figure 3.47 on page 74 are plotted the blade 1 and 2 low root bending moments where it is
seen again the high change in load between the rotor azimuthal positions. The time series
of the blade-1 position at equator station in x, y and z direction expressed in blade-1
coordinate system are plotted in figure 3.48 on page 75 for 20 rotor azimuthal angle.
The results show an oscillation of the blade that is growing and for example, the out-ofplate motion (x, edgewise) of the blade reaches unrealistic high values of 15 m. Power
spectra density (PSD) analysis of the signals in the beginning of the growing oscillations

3.8 Parked Rotor

Bending Moment ||M2x +M 2y || [kNm]

12

73

10 4
Blade-1 low root (max)
Blade-2 low root (max)
Blade-1 low root (std)
Blade-2 low root (std)

10

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.45: Maximum and standard deviation of blade one and two low roots bending
moments expressed at local coordinate systems. The wind turbine is parked in standing still mode
at different rotor orientations and exerted to 50 year recurrence period extreme winds.

3.5

10 4
|M x | (mean)
|M x | (std)

Bending Moments [kNm]

|M y | (mean)
|M y | (std)

2.5

1.5

0.5

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.46: Mean and standard deviation of blade one low roots bending moments (Mx
flatwise, My edgewise) expressed at local coordinate system. The wind turbine is parked in
standing still mode at different rotor orientations and exerted to 50 year recurrence period
extreme winds.

gives a frequency which does not match to any of the natural frequencies of the blades.
Furthermore, initially the motion grows exponentially and not linearly, indicating that an

74

Design Load Cases

instability is the source of the phenomenon and not a resonance according also to a study
regarding turbine instabilities in reference [7, p. 3]. PSD analysis on the fully developed
oscillation shows that the first frequency which also has the highest energy corresponds
to the 1st blade edgewise natural frequency. In figure 3.49 on page 76 it is depicted
one period of the motion of blade-1 at equator station with the fully developed blade
instability (black line). The components of the same motion on x, y and z directions with
the angle of attack and the lift coefficient variations are plotted in figure 3.50 on page 76.

Bending Moment ||M2x +M 2y || [kNm]

12

10 4
Blade-1 low root (max)
Blade-2 low root (max)
Blade-1 low root (std)
Blade-2 low root (std)

10

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.47: Maximum and standard deviation of blade 1 and 2 low roots bending moments
expressed at local coordinate systems. The wind turbine structure is set as rigid body with only
the blades kept flexible. Simulation for parked-standing still mode at different rotor orientations
with 50 year recurrence period extreme winds without turbulence.

The analysis of the standing still wind turbine with extreme winds of 1 year recurrence
period (34 m/s mean wind speed at hub-height) is also performed. This could represent
the DLC 7.1 of the current standard where the turbine is parked with fault conditions.
The fault could be that the rotor is locked at a specific azimuthal angle without being
able to rotate slowly (idling mode). The turbine base and low root bending moments are
shown in figure 3.51 on page 77 and have similar abrupt changes at different orientations
as the results in figure 3.44 on page 72 and in figure 3.45 on the preceding page for the
50 year ETM wind conditions. The motion of the blade exceeds the 5 m from = 20 to
50 and = 140 to 150 . Again the simulations repeated without including the dynamic
stall model (appendix D.4, figure D.13 on page 121). The results reveal high loads at the
critical azimuthal angles and almost the same at the other orientations where the blades
are stable. This happens because when the turbine is in standing still the dynamic-stall
has very small influence in the lift coefficient.
Consequently, if there is a possibility during its time-life the turbine to be at standing
still mode under extreme winds, should be foreseen a way to rotate the rotor in azimuthal
angles around = 0 (equivalent = 180 ) or = 90 (equivalent = 270 ) where
no instabilities occur. Another way to avoid this situation could be to install air-brakes

75

20

Position y [m]

-20
150
70

Position z [m]

Position x [m]

3.8 Parked Rotor

55

200

250

300

350

400

200

250

300

350

400

200

250

300

350

400

60

50
150

50

45
150

Time [sec]

Figure 3.48: Time series of the blade-1 station position (x,y,z) at equator height expressed in
the blade-1 coordinate system. The wind turbine structure is set as rigid body with only the
blades kept flexible. Parked-standing still case at 20 rotor orientations with 50 year recurrence
period extreme winds without turbulence.

or active flap mechanisms. In addition, it is stressed that the above analysis was based
on the 2-bladed VAWT model while for a 3-bladed vertical wind turbine rotor where
always the rotor has a blade at critical angles may be more important. On the other
hand, potential instability of the one blade could be counteracted by the two others, so
further analysis is required for a 3-bladed VAWT. Lastly, it should be noticed that the
aerodynamic damping in a real case could be higher and prevent such an instability.
At table 3.8 on page 77 an overall comparison of maximum loads is made between the
horizontal and the vertical axis wind turbines in parked design situation. The HAWT
is in idling mode with 8 yaw misalignment, while the VAWT in idling (at 5% of rated
rotor speed) and standing still modes. It should be mentioned that the loads in standing
still mode for the VAWT are the highest and correspond to the rotor orientation where
instability occurs.

76

Design Load Cases

150

Position z [m]

100

50

0
-20

20

20

40

60

Position y [m]

Position x [m]

Figure 3.49: Motion of the blade-1 station position (x,y,z) at equator height for one period
(black line). The red line depicts the center-line of the tower and blade-1. The wind turbine
structure is set as rigid body with only the blades being flexible. Parked-standing still case at 20
rotor orientations with 50 year recurrence period extreme winds without turbulence.

70

60

-10

-0.5

-20

50

-1

30

-40

Cl [-]

x
y
z

[deg]

Position [m]

-30
40

-50

-1.5

-2

20
-60
-2.5

10

-70

0
-10
501

-3

-80

502

503

time [sec]

504

-90
501

502

503

time [sec]

504

-3.5
501

502

503

504

time [sec]

Figure 3.50: Position (x,y,z) of the blade-1 station at equator height together with the angle of
attack and the lift coefficient for one period of instability. The wind turbine structure is set as
rigid body with only the blades being flexible. Parked-standing still case at 20 rotor orientations
with 50 year recurrence period extreme winds without turbulence.

3.8 Parked Rotor

77

Turbine Base Bottom

10 5

Bending Moment ||M 2x +M 2y || [kNm]

Turbine bottom (maximum)


Turbine bottom (STD)

Blade Low Root

10 4

Blade-1 low root (max)


Blade-2 low root (max)
Blade-1 low root (std)
Blade-2 low root (std)

5
6
4
5
3

4
3

2
2
1
1
0

0
0

20

40

60

80

100

120

140

160

180

20

40

Rotor Orientation [deg]

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.51: Maximum and standard deviation of the turbine base bottom (left) and blade roots
(right) bending moments expressed at the global and blade-1,2 coordinate systems respectively.
The Wind turbine is parked in standing still mode at different rotor orientations and exerted to
1 year recurrence period extreme winds.

Table 3.8: Loads and deflections of wind turbines with idling rotor under 50 year recurrence
period wind conditions.

Maximum values

Tower bottom BM [kNm]


Blade low root BM [kNm]
Blade upper root BM [kNm]
Blade root BM [kNm]

VAWT

HAWT

Standing still

Idling

Idling

8.2x105
1.0x105
5.8x104
-

30.5x104
2.6x104
1.8x104
-

8.0x104
1.0x104

78

Design Load Cases

Influence of Blade Structural Properties


As the survival of the wind turbine is vital when is standing still and exposed to 50 year
return period extreme winds, further analysis is conducted by changing the structural
parameters of blades such as the stiffness and the damping to see the influence.
The flatwise, edgewise and torsional stiffness of the blades are increased 50% and 100%
separately each time. In figure 3.52 are shown the results from the simulations for the
blade-1 low root maximum bending moments for 100% increase of the blade stiffness.
Similar results where extracted when the stiffness increased 50% and are not plotted. For
rotor orientations from 10 to 50 the loads still increase abruptly. Looking on the time
series is found again the high oscillation on the blade loads and deflections.

16

10 4
baseline
100% increase of edgewise stif.
100% increase of torsional stif.
100% increase of edgewise stif.

Bending Moment ||M2x +M 2y || [kNm]

14

12

10

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.52: Sensitivity analysis of the influence of the blade stiffness on blade one low root
maximum bending moment. Moment expressed at blade-1 coordinate system. The wind turbine is
parked in standing still mode at different rotor orientations and exerted to 50 year recurrence
period extreme winds.

Finally, the structural damping of the blades is increased (through the Rayleigh damping
input parameter in HAWC2) in order the 1st blade modes to have a logarithmic damping
decrement of 5% ( = 0.05 [-]). This is equivalent to a damping ratio = 0.008 [-] as
converted using equation 3.20 on the facing page. The simulations are repeated keeping
all other conditions the same. It should be mentioned that for the traditional HAWT
simulations the logarithmic damping decrement for the 1st blade modes is set to around
3%. According to the DeepWind report [80, p. 16] for the case of VAWTs where the blade
has both ends connected and due to absence of further investigations, 5% is suggested as
the maximum upper value. In figure 3.53 on the next page the blade-1 low root bending
moment is plotted, where it can be seen that instabilities are still present due to the high
peaks in loads. The time series of the blade deflections were also checked and confirm

3.9 Guy wires

79

that the very high blade motion still exists at the critical rotor orientations.
1
=q
1+

12

(3.20)


2 2

10 4
baseline
Increased blade struct. damping

Bending Moment ||M2x +M 2y || [kNm]

10

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure 3.53: Sensitivity analysis of the influence of blade structural damping on blade one low
root maximum bending moment. Moment expressed at blade-1 coordinate system. Wind turbine
is parked in standing still mode at different rotor orientations and exerted to 50 year recurrence
period extreme winds.

The sensitivity analysis on blade stiffness and structural damping reveals that the instability is strong and is not possible to avoid it either with highly damped or stiffer blades at
specific rotor orientations due to aerodynamic loading from high winds of 42.5 m/s (storm
conditions). The long blades in combination with the absence of the stiffening effects from
centrifugal forces in standing-still mode could be considered the primary reasons that lead
to blade instability.

3.9

Guy wires

In the IEC 61400-3 does not foreseen any design load case with respect to wind turbines
with guy wire support. This is most probably due to the fact that in HAWTs the rotation
of the blades does not allow to install guy wires close to the tower top and the extra space
that would occupy at the ground does not make them an interesting option. So, the wind
turbine manufactures do not use at all tower cables and as a result, it is not included
any relevant design load case in the current standard. In this study a complementary
VAWT model was made with guy wires (see subsection 2.2.2 on page 28) since it is a
rather convenient option for Darrieus configurations and an analysis regarding the loads
is conducted below.
In figure 3.54 on the next page are compared the tower bottom and the blade low root
mean and standard deviation bending moments for the VAWT model with and without

80

Design Load Cases

the guy wires. The turbine operates normally at the same wind conditions as in section 3.2
on page 38. It can be seen that the turbine bottom bending moments are around half
when the guy wires are implemented. The peak on base bottom loads at 15.5 m/s wind
speed for the VAWT without guy wires as was mentioned many times it is because the
turbine operates close to the crossing of the 1st tower natural frequencies with the 4P .
The blade loads are almost the same and unaffected by the addition of the wires. The
new blade load peaks at 16.5 and 17.5 m/s wind speeds in the case of the turbine with guy
wires appear because the rotor rotates very close to 0.53 rad/s where the 5P crosses the
1st propeller mode. A comparison of the turbine base bottom bending moments between
the HAWT and the VAWT with guy wires is made in figure 3.55.
16

Blade Low Root

7000

Mean (Vawt without guy wires)


Mean (Vawt with guy wires)
STD (Vawt without guy wires)
STD (Vawt with guy wires)

14

Bending Moment ||M2x +M 2y || [kNm]

Turbine Base Bottom

10 4

6000

12
5000
10
4000
8
3000
6
2000
4
1000

0
0

10

15

20

25

Wind Speed [m/s]

10

15

20

25

Wind Speed [m/s]

Figure 3.54: Turbine base bottom bending moment (left) and blade low root bending moment
(right) as a function of wind speed for the VAWT with and without guy wires.

Base Bottom ||M 2x +M 2y || [kNm]

10 4
Mean (VAWT with guy wires)
Mean (HAWT)
STD (VAWT with guy wires)
STD (HAWT)

0
0

10

15

20

25

Wind Speed [m/s]

Figure 3.55: Comparison of mean and standard deviation turbine base bottom bending
moments as a function of wind speed between the VAWT with guy wires and the HAWT.

If a VAWT is designed with guy wires the tower and the bottom connection between the
rotor and the base could be designed according to the lower emerging loads in order to
minimize the cost. Thus, the wires will play a critical role to the structural integrity of the
turbine under extreme conditions and fatigue damage. This means that the possibility

3.9 Guy wires

81

of one wire to brake or lose considerable amount of its pretension could be catastrophic.
Due to this it is proposed that a revision of the standard should include a new predefined
design situation under normal power production plus occurrence of fault where one of
the wires will have much lower pretension than the nominal and the ultimate and fatigue
loads should be investigated. Furthermore, it is also recommended a new design load case
where the turbine is parked having one of the wires broken under 1-year recurrence period
wind conditions. Other aspects such as thermal effects, ice and possible collision of a loose
or broken wire with the blades should also be taken into consideration. Finally, safety
factors for the wires should be set. According to a paper regarding the guy cable design
for VAWTs by Thomas G. Carne of Sandia National Laboratories [12, p. 7] a minimum
safety factor of 3 is recommended for 30 years lifetime.

82

Design Load Cases

Chapter 4
Discussion of DLC Results and
Recommendations

The wind turbine components should withstand the ultimate loads and fatigue damage
under any situation during the turbines entire design lifetime. The IEC 61400-3 standard
defines a minimum set of DLCs that should be investigated and prove the structural
integrity of the turbine. Different DLCs were simulated in the previous chapter. Below
an overall critical review of the results is made, as well as, recommendations for future
amendments of the standard in order to include the specifics of VAWTs.
The VAWT base bottom and blade low root cross section bending moments were extracted, among other results, under different wind conditions and turbine states (table 3.3
on page 37). Sensitivity analyses were performed to gain insight of parameters of the turbine and the external conditions that affect the loads. In the figures below are compared
the absolute maximum values from each load case. Design situations that refer to power
production were simulated from cut-in up to cut-out speeds at the center of each wind
speed bin with width of 1 m/s, (4.5 up to 24.5 m/s). The emergency shut down simulations were performed at 8.5, 13.5, 18.5 and 24.5 m/s wind speeds and the case of extreme
operating gust at 8.5, 13.5, 19.5 and 24.5 m/s. The Parked standing still (locked rotor)
and idling situations for the selected baseline wind turbine class B correspond to 42.5 m/s
and 34 m/s extreme winds under 50 years and 1 year recurrence periods respectively. For
the cases where the simulations were conducted at various wind speeds the maximum
load is selected excluding the results at 15.5 m/s winds. This is done because the turbine
operates close to a resonance point and in practice this rotor speed will not be selected
as an operating point (see section 3.2 on page 38). In figure 4.1 on the following page are
plotted the VAWT turbine base bottom and blade low root bending moments. The loads
emerging from DLC 1.1 are the highest for both and follow the loads from the DLC 1.3
for the turbine base bottom and the EDC for the blade roots. The case of DLC 6.1
with the turbine standing still (locked rotor) is not included since instabilities occur at
some rotor orientations resulting to very high loads (see subsection 3.8.3) and in practice
these should be avoided. In addition, in the real case the aerodynamic damping may be

83

84

Discussion of DLC Results and Recommendations

higher than the one emerging by the simulations with the current dynamic stall model
and further analysis on this topic is proposed. If the instabilities are present one way is
to orient the rotor in azimuthal position where remains stable or to use devices such as
aero-brakes or flaps that will be activated in standing still mode.

Design Load Case

Turbine Base Bottom

Blade Low Root

DLC 6.3 1-yr EWM idl.

DLC 6.3 1-yr EWM idl.

DLC 6.1 50-yr EWM idl.

DLC 6.1 50-yr EWM idl.

DLC 5.1 emerg. shut down

DLC 5.1 emerg. shut down

Negative EDC

Negative EDC

Positive EDC

Positive EDC

DLC 2.3 EOG

DLC 2.3 EOG

DLC 1.5 EWV

DLC 1.5 EWV

DLC 1.5 EWS

DLC 1.5 EWS

DLC 1.3 ETM (c=3m/s)

DLC 1.3 ETM (c=3m/s)

DLC 1.3 ETM (c=2m/s)

DLC 1.3 ETM (c=2m/s)

DLC 1.1 NTM extrap. 50 years (Iref=0.16)

DLC 1.1 NTM extrap. 50 years (Iref=0.16)

DLC 1.1 NTM extrap. 50 years (Iref=0.14)

DLC 1.1 NTM extrap. 50 years (Iref=0.14)

DLC 1.1 NTM extrap. 50 years (Iref=0.12)

DLC 1.1 NTM extrap. 50 years (Iref=0.12)

Bending Moment ||M2x +M 2y || [kNm]

5
10 5

Bending Moment ||M2x +M 2y || [kNm]

6
10 4

Figure 4.1: Comparison of the absolute maximum turbine base bottom and blade low root
bending moments for each DLC.

Based on the results of the investigated DLCs and their critical review recommendations
are made below. It should be kept in mind that the outcomes are based on simulations of
one Darrieus VAWT and sensitivity analyses were made to cover the influence of various
factors. However, the recommendations regarding the standard should be demonstrated
and extended to other VAWT models before changes are made.
The large fluctuation of the aerodynamic loads on the blades with 1P dominant frequency
results in highly deterministic internal loads for the blades and the other components even
under turbulent wind. This
1. The current methods of extrapolation of loads in DLC 1.1 may be too conservative
since the loads in VAWTs are more deterministic and tend to be less affected from
the turbulence in contrast to the HAWTs.
2. The extreme turbulence model of design load case 1.3 has been created in order to
avoid the statistical extrapolation of DLC 1.1, this is why the c factor is proposed to
be increased if needed. This together with the more deterministic loads of VAWT
indicates that the ETM might need to be re-tuned in case of VAWTs.
3. Extreme operating gust wind condition simulations revealed that the rising in loads
is small but depends on the rotor orientation. As a result, the application of EOG
should be investigated in combination with different rotor orientations for the case
of VAWTs due to the 3D extension of the rotor in space.
4. In VAWTs a blade station during every revolution remains at the same height and
sees the same mean wind speed even if strong wind shear exists. This means that

85

any transient change in wind shear is not expected to highly increase the load level
in contrast to HAWTs. This was also confirmed in the analysis of EWS (DLC 1.5).
So, the EWS case is not critical and may become optional for the VAWT case.
5. Due to the ability of VAWTs to be omni-directional any abrupt change in wind
direction does not create large loads since there is not yaw misalignment conditions as in HAWTs. The case of EDC was studied and confirms that the change
in wind direction is not followed by large transient loads. Thus, the EDC wind
condition which in the standard is proposed as critical combination with the start
up of the wind turbine can be also optional for VAWTs if it could be demonstrated
in simulations including a controller.
6. Emergency shut down using a mechanical brake was investigated in section 3.7.
The loads arising from the emergency stop are lower compared with the statistical
extrapolated 50 year return period loads of DLC 1.1 when the brake torque is 3
and 4 times the generator torque. With these brake torques the turbine was able
to decelerate below 10% of rated rotor speed at less than 10 sec.
7. According to the results, the parked rotor situation under extreme wind conditions
seems to be important for VAWTs. Especially if a loss of electrical network or other
fault conditions occur (DLC 6.2, DLC 7.1) and the rotor is not able to idle in low
rotor speeds. This is due to the reason that the blades have one less degree of
freedom compared with HAWTs which can pitch to 90 degrees and minimize the
aerodynamic loads. For the VAWTs at high winds and specific rotor orientations
(locked rotor) the blade angle of attack is high and consequently the aerodynamic
loads, as well. This condition could result in blade instabilities as was found from
this study. If this can not be solved be altering the blade design, it is proposed a
system that can rotate the rotor to a safe orientation when the rotor is not able
to maintain a low rotor speed. It should be noted that in a 3-bladed configuration
where one blade is always orientated in an unfavourable position may not be able to
avoid instability phenomena in high wind speeds in case the rotor is standing still.
When the rotor is idling the ultimate loads are low compared to the other design
load cases results.
8. The current version of standard does not include any predefined DLC regarding the
possibility of guyed tower. In onshore VAWT applications in the past guy wires
for tower support were widely used due to some advantages. It is most probably
new large scale onshore VAWTs to adopt similar solutions. Guy wires introduce
some challenges in the design, for instance wire resonances and wire-blade clearance
that can appear due to temperature fluctuations or break of one wire. The case of
guy wires needs further analysis and might be necessary to include potential critical
design cases in the minimum requirements of the standard.

86

Discussion of DLC Results and Recommendations

Chapter 5
Conclusions and Recommendations
for Further Research

5.1

Conclusions

This work had as an ultimate target to study the design load cases of IEC 61400-3 ed.3
standard for the case of multi-megawatt onshore VAWTs to reveal potential critical cases
and make recommendations for future amendments. The whole analysis is based on
aeroelastic simulations of a 5 MW Darrieus VAWT model. In order to reach the objective
a broad range of aspects was studied.
Firstly, a numerical code able to simulate reliable the aerodynamics and the structural
response of VAWTs was sought. This part was accomplished by a special project report
which is complementary of the thesis work [20]. From the built large scale VAWTs
up until now, only the Sandia 34 m test bed performance details are widely reported
in the bibliography [2, 79]. Thus, this wind turbine was used for the code validation.
The HAWC2 numerical tool [39] was selected as the most complete to the knowledge of
the author among the few aeroelastic codes able to simulate VAWTs. Relatively high lift
coefficients close to the VAWT blades were found during the initial aeroelastic simulations.
These were considered unrealistic and a small modification to the code was made to bypass
the dynamic stall model at that regions. However, comparison of the rotor power and
loads revealed a very small change and only at high wind speeds. A good overall agreement
was found between the simulations and the published Sandia 34 m VAWT measurements.
Next for the needs of the study and since the utility scale HAWTs have rated power of the
order of MW it was decided to develop a 5 MW Darrieus VAWT aeroelastic model. This
also gave the opportunity to capture responses like instabilities or investigate influences,
for instance the wind shear that may only appear in large scale rotors. The final wind
turbine model was based on the 5 MW DeepWind rotor and the up-scaled Sandia 34 m
VAWT. The blade centreline, chord and airfoils of the DeepWind model were used and
the blade and tower structural properties were modified after consulting the up-scaled

87

88

Conclusions and Recommendations for Further Research

blade properties of the Sandia wind turbine. This led to a stiffer rotor compared with the
DeepWind one. An induction generator model with inverter able to operate at different
rotor speeds was added to the model. A secondary model with guy wires for tower support
was also developed and the loads under normal power production were investigated. Based
on the results the turbine base bottom bending moments can be reduced considerably if
guy wires used. The loads and the natural frequencies of the blades are not affected
while the tower natural frequencies are increased. Furthermore, the NREL HAWT model
with the same rated power of 5 MW was simulated and comparisons were made with
the VAWT results. The ultimate and 1 Hz equivalent fatigue loads of the blade root
and turbine base bottom were compared during normal operation and wind conditions in
order to give an insight of the load levels between the horizontal and vertical axis wind
turbines. According to the results the ultimate VAWT blade upper root bending moments
are smaller from the HAWT ones below rated wind speed and slightly higher above. On
the other hand the blade low roots on VAWT appera to be exerted to larger maximum
bending moments but the stresses could be kept at the desired limits by strengthening the
blade to tower connection areas. The maximum base bottom bending moments emerging
on VAWT during operation are higher compared to the HAWT. Similar trend is revealed
for the 1 Hz equivalent fatigue loads for both the base and the blade roots between the
horizontal and the vertical axis wind turbine comparisons. However, 2 points should be
kept in mind on the load comparison between the VAWT and the HAWT. The latter has
three blades unlike the developed VAWT which has two. Keeping the same solidity and
tip speed ratio a blade in a 3 bladed VAWT is exerted to lower aerodynamic forces and the
resulting overall rotor load variations are smaller, thus maximum loads will be smaller.
The second regards the developed VAWT model where it can be said that small room for
improvements exist, thus in a more optimum design (for instance using MDAO techniques)
as the NREL 5 MW reference turbine the maximum loads could be few percent less.
Then the design load cases of the IEC 61400-3 Ed.3 were studied and the effect of different
wind conditions and turbine properties were investigated. The conclusions regarding the
DLCs and the standard are summarized in the previous chapter on page 83 and are not
repeated here.

5.2

Recommendations for Further Research

Many different aspects around VAWTs were touched during this study regarding the
simulation tool, the aeroelastic model and the design load cases of the IEC 61400-3 ed.3
standard. Below are given the most important which are worth further research.
Regarding the aeroelastic code HAWC2, a modification of the dynamic stall model
is proposed in order to predict better the lift coefficient close to the VAWT blade
roots. Other dynamic stall models found in the bibliography for VAWTs [61, p. 199]
could also be implemented in HAWC2. In addition, the flow curvature effect which
arises in VAWTs when the blade rotates and act as a virtual camber is not taken
into account in the current code. This effect becomes more important as the wind
turbine solidity increases and should be investigated.

5.2 Recommendations for Further Research

89

Load analysis and the corresponding DLCs of the standard, normal start up and
shut down, which are highly dependent on the wind turbine controller, were not
considered. Implementation of a controller and analysis of the aforementioned DLCs
including possible different control strategies is needed.
The study was focused on a two bladed Darrieus wind turbine configuration. The 3
bladed configuration was used in the past due to some advantages and is likely to be
adopted in new designs. Some of the outcomes regarding the design load cases may
be different and an investigation of a 3 bladed VAWT is proposed. Furthermore,
rotors with helical blades or H-type could be tested.

90

Conclusions and Recommendations for Further Research

References

[1] P.D. Andersen. Introduction to Wind Turbines Technology and Aerodynamics,


Course 46300. Lectures material, DTU, 2013.
[2] T.D. Ashwill. Measured Data for the Sandia 34-Meter Vertical Axis Wind Turbine.
Sandia National Laboratories, SAND91-2228, July 1992.
[3] T.D. Ashwill. Structural Design and Fabrication of the Sandia 34-Meter Vertical
Axis Wind Turbine. Proceedings of the 1987 ASME-JSME Solar Energy Conference,
ASME, March 1987.
[4] T.D. Ashwill. The Status of the US VAWT Program. Proceedings of the Seventh
Annual National Conference of the Canadian Wind Energy Association, pages 375
394, March 1992.
[5] D. E. Berg. Structural Design of the Sandia 34-Meter Vertical-Axis Wind Turbine.
Sandia National Laboratories, SAND84-1287, April 1985.
[6] A. Betz. Wind-energie und ihre ausnutzung durch windm
uhlen. Vandenhoeck, 1926.
[7] G. Bir and J. Jonkman. Aeroelastic Instabilities of Large Offshore and Onshore Wind
Turbines. Journal of Physics: Conference Series, 75:012069, 2007.
[8] J. P. Blasques. BECAS software - a cross section analysis tool for anisotropic and
inhomogeneous beam sections of arbitrary geometry. DTU Ris - National Laboratory
for Sustainable Energy, January 2012.
[9] E.A. Bossanyi. GH Bladed. Theory Manual, 282/BR/009, September 2010.
[10] E.A. Bossanyi. GH Bladed. User Manual, 282/BR/010, September 2010.
[11] P. Brndsted, H. Lilholt, and A. Lystrup. Composite Materials for Wind Power
Turbine Blades. Annual Review of Materials Research, 35:505538, 2005.
[12] T.G. Carne. Guy Cable Design and Damping for Vertical Axis Wind Turbines.
Sandia National Laboratories, SAND80-2669, 1980.

91

92

References

[13] S. Eriksson, H. Bernhoff, and M. Leijon. Evaluation of different turbine concepts


for wind power. Renewable and Sustainable Energy Reviews, 12(5):1419 1434,
2008. ISSN 1364-0321. URL http://www.sciencedirect.com/science/article/
pii/S1364032107000111.
[14] C.S. Ferreira. The near wake of the VAWT: 2D and 3D views of the VAWT aerodynamics. Ph.D. Thesis, Delft University of Technology, 2009.
[15] C.S. Ferreira, G. van Bussel, and G. van Kuik. Wind tunnel hotwire measurements,
flow visualization and thrust measurement of a VAWT in skew. Journal of Solar
Energy Engineering, 128(4):487497, 2006.
[16] C.S. Ferreira, K. Dixon, C. Hofemann, G. van Kuik, and G. van Bussel. The VAWT
in skew: stereo-PIV and vortex modeling. In 47th AIAA Aerospace Sciences Meeting,
pages 58, 2009.
[17] C.S. Ferreira, H.A. Madsen, M. Barone, B. Roscher, P. Deglaire, and I. Arduin.
Comparison of aerodynamic models for Vertical Axis Wind Turbines. Journal of
Physics: Conference Series, 524(1):012125, 2014. ISSN 17426588, 17426596. doi:
10.1088/1742-6596/524/1/012125.
[18] P.P. Friedmann. Renaissance of Aeroelasticity and Its Future. Journal of Aircraft,
36(1):105121, 1999. ISSN 00218669.
[19] Richtlinie f
ur Windkraftanlagen. Einwirkungen und Standsicherheitsnachweise f
ur
Turm und Gr
undung. Schriften des Deutschen Instituts f
ur Bautechnik, DIBt, Hrsg.,
Germany, 1993.
[20] C. Galinos. Aeroelastic study of Vertical Axis Wind Turbine based on Sandia 500 kW
test-bed. Special Project, Technical University of Denmark, January 2015.
[21] Germanischer Lloyd. Guideline for the Certification of Wind Turbines. July 2010.
[22] P. Gipe. Photos of vertical axis wind turbines. [Website cited 2015 Feb 26]. URL
http://www.wind-works.org/cms/.
[23] H. Glauert. The elements of aerofoil and airscrew theory. Cambridge University
Press, 1948.
[24] E. Hau. Wind Turbines, Fundamentals, Technologies, Application, Economics, 3rd
edition. Springer, 2012.
[25] G. Hu and E. Hau. Lastfalldefinition f
ur WKA-60. MAN-Report, 1986.
[26] International Energy Agency (IEA). International Recommended Practices for Wind
Energy Conversion Systems Testing 3. Fatigue Characteristics, 1st edition. IEA
Expert Group Study, 1984.
[27] International Electrotechnical Commission (IEC). Wind turbine generator systems
Part 1: Safety requirements, IEC 61400-1, 1st edition. IEC, Switzerland, 1994.
[28] International Electrotechnical Commission (IEC). Wind turbines Part 1: Design
requirements, IEC 61400-1, 3rd edition. IEC, Switzerland, 2005.

References

93

[29] Adecon Energy Systems Inc. Performance Testing of an Adecon A 19 m 150 KW


Vertical Axis Wind Turbine at the Atlantic Wind Test Site. Energy Mines and
Resources, Enerdemo Canada Program, September 1990.
[30] Dansk Ingenirforening. Load and Safety for Wind Turbine Structures, DS 472. Danish
Standards, Denmark, 1992.
[31] International Electrotechnical Commission (IEC). Wind turbines Part 3: Design
requirements for offshore wind turbines, IEC 61400-3. IEC, Switzerland, 2009.
[32] P. Jamieson. Innovation in wind turbine design. John Wiley and Sons, 2011.
[33] Z. Jiang, M. Karimirad, and T. Moan. Dynamic response analysis of wind turbines
under blade pitch system fault, grid loss, and shutdown events. Wind Energy, 17(9):
13851409, 2014.
[34] J.M. Jonkman and L. Marshall. FAST Users Guide, Technical Report, NREL/EL500-38230. August 2005.
[35] J.M. Jonkman, S. Butterfield, W. Musial, and G. Scott. Definition of a 5 MW Reference Wind Turbine for Offshore System Development. National Renewable Energy
Laboratory, US, 2009.
[36] B.S. Kallese, A.M. Hansen, et al. Dynamic mooring line modeling in hydro-aeroelastic wind turbine simulations. In Twenty-first International Offshore and Polar
Engineering Conference, Maui, Hawaii, USA, 2011.
[37] L. Bergami and M. Gaunaa. ATEFlap Aerodynamic Model, a dynamic stall model
including the effects of trailing edge flap deflection. Technical Report, Ris-R1792(EN), Ris DTU National Laboratory for Sustainable Energy, 2012.
[38] T.J. Larsen. Slip generator model implemented in HAWC2 as an external DLL. Ris
National Laboratory, Technical University of Denmark, 2005.
[39] T.J. Larsen and A.M. Hansen. How 2 HAWC2, the users manual, Ris-R-1597(ver.
4-5)(EN). Ris National Laboratory, Technical University of Denmark, July 2014.
[40] T.J. Larsen and H.A. Madsen. On the way to reliable aeroelastic load simulation on
VAWTs. In Proceedings of EWEA 2013, 2013.
[41] T.J. Larsen, H.A. Madsen, A.M. Hansen, and K. Thomsen. Investigations of stability
effects of an offshore wind turbine using the new aeroelastic code HAWC2. Proceedings of Copenhagen Offshore Wind 2005, Copenhagen, Denmark, pages 2528, 2005.
[42] J.G. Leishman and T.S. Beddoes. A generalised model for airfoil unsteady aerodynamic behaviour and dynamic stall using the indicial method. In Proceedings of the
42nd Annual forum of the American Helicopter Society, pages 243265. Washington
DC, 1986.
[43] C. Lindenburg. PHATAS and TURBU Focus tools for aeroelastic analysis and load
calculation of wind turbines. Energy research Centre of the Netherlands, Dutch Wind
Workshops, October 2006. doi: 10.2172/1035336.

94

References

[44] H.A. Madsen. The Actuator Cylinder, a flow model for vertical axis wind turbines.
Part 1,2 and 3. Ph.D. Thesis, Aalborg university centre, 1982.
[45] H.A. Madsen, T.J. Larsen, L. Vita, and U.S. Paulsen. Implementation of the Actuator Cylinder flow model in the HAWC2 code for aeroelastic simulations on Vertical
Axis Wind Turbines. In 51st AIAA Aerospace Sciences Meeting including the New
Horizons Forum and Aerospace Exposition, 2013.
[46] J. Mann. The spatial structure of neutral atmospheric surface-layer turbulence.
Journal of Fluid Mechanics, 273:141168, 1994.
[47] J.F. Manwell, J.G. McGowan, and A.L. Rogers. Wind Energy Explained: Theory,
Design and Application,2nd Edition. Wiley, December 2009.
[48] M.H. Hansen, and M. Gaunaa, and H.A. Madsen. A Beddoes-Leishman type dynamic stall model in state-space and indicial formulations. Technical Report, RisR-1354(EN), Ris National Laboratory, 2004.
[49] P.G. Migliore, W.P. Wolfe, and J.B. Fanucci. Flow curvature effects on Darrieus
turbine blade aerodynamics. Journal of Energy, 4(2):4955, 1980.
[50] P.J. Moriarty, W.E. Holley, and S.P. Butterfield. Effect of turbulence variation on
extreme loads prediction for wind turbines. Journal of solar energy engineering, 124
(4):387395, 2002.
[51] P.J. Moriarty, W.E. Holley, and S.P. Butterfield. Extrapolation of Extreme and
Fatigue Loads Using Probabilistic Methods. November 2004.
[52] P. Musgrove. Wind Power. Cambridge University Press, February 2010.
[53] C. Nath, C. Eriksson, F. van Hulle, C. Skamris, W. Stam, and P. Vionis. Harmonisation of wind turbine certification in Europe, JOULE project EWTC. Wind Energy
for the Next Millennium. Proceedings, pages 563567, 1999.
[54] Nenuphar-wind. Vertiwind 2 MW VAWT. [Website cited 2015 Feb 22]. URL http:
//www.nenuphar-wind.com/en.
[55] Nederlands Normalisatie-institut.
Safety requirements for wind generators,
NEN6096. Netherlands Standards, The Netherlands, 1991.
[56] Sandia National Laboratories-US Department of Energy and the American Wind
Energy Association. Vertical Axis Wind Turbines, The History of DOE Program,
April 1985.
[57] B.C. Owens, J.E. Hurtado, J.A. Paquette, D.T. Griffith, and M. Barone. Aeroelastic Modeling of large offshore Vertical-axis Wind Turbines: Development of
the Offshore Wind Energy Simulation Toolkit. Collection of Technical Papers
- AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials
Conference, pages AIAA 20131552, 2013. ISSN 02734508.

References

95

[58] B.C. Owens, D.T. Griffith, and J.E. Hurtado. Modal Dynamics and Stability of Large
Multi-megawatt Deepwater Offshore Vertical-axis Wind Turbines: Initial Support
Structure and Rotor Design Impact Studies. 32nd Asme Wind Energy Symposium,
2014.
[59] S. ye. Dynamic stall-simulated as time lag of separation. In Proceedings of the 4th
IEA Symposium on the aerodynamics of wind turbines, 1991.
[60] I. Paraschivoiu. Double-multiple streamtube model for studying vertical-axis wind
turbines. Journal of propulsion and power, 4(4):370377, 1988.
[61] I. Paraschivoiu. Wind Turbine Design with Emphasis on Darrieus Concept. Polytechnic International Press, 2002.
[62] P. Passon, M. Kuhn, S. Butterfield, J. Jonkman, T. Camp, and T.J. Larsen. OC3benchmark exercise of aero-elastic offshore wind turbine codes. Journal of Physics:
Conference Series, 75:012071, 2007. ISSN 17426596, 17426588.
[63] U.S. Paulsen, L. Vita, H.A. Madsen, J. Hattel, E. Ritchie, K.M. Leban, P.A.
Berthelsen, and S. Carstensen. 1st DeepWind 5 MW baseline design. Energy Procedia, 24:2735, 2012. ISSN 18766102.
[64] U.S. Paulsen, H.A. Madsen, J.H. Hattel, I. Baran, and P. Nielsen. Design optimization of a 5 MW floating offshore vertical-axis wind turbine. Energy Procedia, 35:
2232, 2013. ISSN 18766102. URL http://dx.doi.org/10.1016/j.egypro.2013.
07.155.
[65] DeepWind EU Project. The DeepWind 5 MW VAWT. [Website cited 2014 Dec 15].
URL http://www.deepwind.eu.
[66] P. Ragan and L. Manuel. Statistical Extrapolation Methods for Estimating Wind
Turbine Extreme Loads. Journal of Solar Energy Engineering, 130(3):031011, 2008.
[67] B. Richards. Initial Operation of Project Eole 4MW Vertical Axis Wind Turibne
Generator. Windpower 87, San Francisco, October 1987.
[68] Germanischer Lloyd Rules and Regulations. Regulation for the Certification of Wind
Energy Conversion Systems Vol. IV Non-Marine Technology, Part 1 Wind Energy.
Germanischer Lloyd, Hamburg, Germany, 1993.
[69] A.A. Shabana. Dynamics of multibody systems. Cambridge university press, 2013.
[70] G. Sieros, P. Chaviaropoulos, J.D. Srensen, B.H. Bulder, and P. Jamieson. Upscaling
wind turbines: theoretical and practical aspects and their impact on the cost of
energy. Wind Energy, 15(1):317, 2012.
[71] N.N. Srensen and H.A. Madsen. Modelling of transient wind turbine loads during
pitch motion. In Proceedings of the European Wind Energy Conference, 2006.
[72] W.A. Stephenson. Test Plan for the 34-Meter Vertical Axis Wind Turbine Test Bed
Located at Bushland, Texas. Sandia National Laboratories, SAND86-1623, December 1986.

96

References

[73] J.H. Strickland. Darrieus turbine: a performance prediction model using multiple
streamtubes. Sandia National Laboratories, Technical Report SAND75-0431, 1975.
[74] H.J. Sutherland, D.E. Berg, and T.D. Ashwill. A Retrospective of VAWT Technology.
Sandia National Laboratories, SAND2012-0304, January 2012.
[75] K. Taeseong. Lecture notes of Loads, Aerodynamics and Control of Wind Turbines
course 46320, DTU, September 2014.
[76] R. Templin. Aerodynamic performance theory for the NRC vertical-axis wind turbine. National Research Council of Canada, Technical Report LTR-LA-160, 1975.
[77] Wikipedia: the free encyclopedia. Brake lining. [Website cited 2015 June 5]. URL
http://en.wikipedia.org/wiki/Brake_lining.
[78] P.S. Veers. Three-dimensional wind simulation. Sandia National Laboratories, Technical Report SAND88-0152, 1988.
[79] P.S. Veers. Selected Papers on Wind Energy Technology. Sandia National Laboratories, SAND90-1615, February 1992.
[80] D.R. Verelst, H.A. Madsen, K.A. Kragh, and F. Belloni. Detailed Load Analysis of
the baseline 5 MW DeepWind Concept, September 2014.
[81] Det Norske Veritas and Ris National Laboratory. Guidelines for Design of Wind
Turbines. Denmark, 2002.
[82] L. Vita. Offshore Vertical Axis Wind Turbines with Rotating Platform. Ph.D. Thesis,
Technical University of Denmark, August 2011.
[83] A. Yde, T.J. Larsen, D.R. Verelst, and L. Bergami. HAWC2 online course as part of
Loads, Aerodynamics and Control of Wind Turbines course 46320, DTU, September
2014.

Appendix A
Modules of the Aeroelastic Code
HAWC2

This appendix gives an overview of the modelling approach of wind turbines in HAWC2
aeroelastic code and the theory behind it for both VAWT and HAWT, as in the thesis
comparison in some cases with a HAWT is conducted using HAWC2 results, as well. The
code is developed by the Aeroelastic Design Section, Department of Wind Energy at DTU
which is located at Ris Campus [39]. HAWC2 performs time domain simulations of the
system, where the dynamics are described by differential equations and solved using the
Newmark numerical integration scheme. It is divided into modules which can be combined
depending what the user wants to include in the analysis. These are the structural,
the aerodynamics, the wind, the hydrodynamics and the soil modules. In addition, the
controller, the generator or other special futures can be implemented creating DLL files
and link them with the code.

A.1

Aerodynamics Module

Different models have been proposed in the literature through the years for the aerodynamics of wind turbine rotors. These include:
1. the Momentum models for horizontal and vertical axis wind turbines
Blade Element Momentum model (BEM) for HAWTs by Glauert and Betz,
1930s and 1940s [6, 23]
Single Stream Tube (SST) model for VAWTs by Templin, 1974 [76]
Multiple Stream Tube (MST) model for VAWTs by Strickland, 1975 [73]
Double Multiple Stream tube (DMST) model for VAWTs by Paraschivoiu,
1988 [60]

97

98

Modules of the Aeroelastic Code HAWC2

2. the Actuator Cylinder (AC) model for VAWTs by Madsen, 1982 [44]
3. the Vortex models for HAWTs and VAWTs
Fixed wake models
Free wake models by Ferreira, 2009 [14], et al.
4. the Computational Fluid Dynamic (CFD) models for HAWTs and VAWTs
2D CFD models
3D CFD models
In HAWC2 are implemented the BEM model for the case of HAWTs and the AC model
for the VAWTs [45]. These models combine low computational time with an acceptable
accuracy which is the key for an engineering tool where many combinations of turbine
configurations and external conditions are simulated.
For HAWTs the classical BEM model accounts for a rotor with infinite number of blades
under constant load and a uniform inflow. In order to enhance the accuracy of the
results approaching the real conditions, known correction models are applied. These
include the Prandtl tip-loss factor to account for a finite number of blades, the Glauert
correction for heavily loaded rotors, the Glauert and Colemann correction to account for
induction variations due to non uniform inflow and the dynamic inflow model to capture
the dynamics of the wake which in turn change the loads on the rotor. Two first order
low pass filters in parallel are applied in time domain in the induced velocities to capture
the dynamic inflow phenomenon as proposed in [71]. One for the near wake and the other
for the far wake effects.
The rotor in HAWC2 is divided into a polar grid and the induced velocities are computed
at every section of the grid which makes it possible to have radial and azimuthal variations
of the induction factor (figure A.1).

Figure A.1: Discretisation of the rotor area into a polar grid as implemented in HAWC2 [83].

The AC model, which is used in HAWC2 for VAWT computations, extends the actuator
disc model to the actuator surface (swept area) of a vertical axis wind turbine. In comparison with the MST models which are used widely in VAWT calculations it predicts better

A.1 Aerodynamics Module

99

the physics as it captures not only the parallel, but also the lateral in-plane component of
the induced velocity that causes the flow to expand as it is decelerated passing through
the rotor [17]. The AC model as implemented in HAWC2 splits the 3D rotor surface to
a number of 2D cylinders even though the aerodynamics of VAWTs are complex and 3dimensional effects exist. Then the solution of the 2D problem is sought for every cylinder
in order to have low computational time requirements which is desirable in aeroelastic
simulations. This is based on a modified linear solution using the Euler and continuity
equations. The simplified solution is corrected based on the thrust coefficient and the
induction factor as in the case of HAWT [40], since a deviation exists in the results of
the downwind part compared with the non-linear solution. It should be mentioned that
the interaction effect of the generated and transported vortices from the upwind part of
the rotor to the downstream blades is neglected compared to CFD and vortex methods
where is taken into account. As it is suggested by H.A. Madsen and T.J. Larsen [40] this
impact is likely very small but it should be investigated further.
The rotor (actuator surface) is divided into sections perpendicular to the rotational axis
where each one has cylindrical swept area. In figure A.2 on the following page left, a
section is represented with the tangential Qt and normal Qn volume forces acting on the
surface (circle). These volume forces are the reactions from the normal Fn and tangential
Ft aerodynamic blade forces (figure A.2 on the next page right) acting to the flow and
can be computed be equations A.1 and A.2. The power is given by eqrefpow.
Fn () cos () Ft () sin ()
2R

(A.1)

Ft () cos () + Fn () sin ()
2R

(A.2)

N (Ft () cos () + Fn () sin ()) Rd

(A.3)

Qn () = N

Qt () = N
1
P =
2

At each cylindrical section the loads and the induction factors are computed at 36 azimuth
positions, accounting for the local free stream conditions. Dynamic inflow is also taken
into account by a low pass filter applied to the induced velocities as in the case of HAWTs
[40], however evaluation of the outcome should be performed.

100

Modules of the Aeroelastic Code HAWC2

Figure A.2: The AC flow model and the corresponding normal Qn and tangential Qt forces
acting to a cylindrical cross section (left), and the aerodynamic blade section forces lift L and
drag D with the projected normal Fn and tangential Ft components on the airfoil (right) [45].

A.1.1

2D Airfoil Aerodynamics

Both the BEM and the AC models compute the aerodynamic forces act on the airfoil
locally (blade element) for finding the induced velocities. These are dependent from the
local flow conditions, as well as, the geometry of the blade. The blade is composed by one
or more airfoil profiles each one having a lift, a drag and a moment coefficient as a function
of angle of attack and Reynolds number (Re). Usually one set of lift, drag and moment
coefficients are used assuming constant Re number. These are usually obtained by wind
tunnel experiments or computational methods (CFD, Panel codes) and refer to steady
state conditions. If the coefficients are available for small angles of attack semi-empirical
models can be used to extend the range; and 3D rotation corrections may be applied for
the 3D rotational effects. It is stressed that the Re number (equation A.4) changes a lot
in the case of VAWTs not only due to rotor geometry (for instance, in Darrieus rotor
configuration the rotor radius varies, figure 1.4 on page 3) but also because the variation
of the relative velocity at every revolution (figure A.3 on the facing page). Thus, it is
suggested the investigation of simulations having airfoil polars for different Re numbers.
Finally, aeroelastic codes like HAWC2 does not give the distribution of velocity and
pressure field around the airfoil as this would require high computational times and the
application of vortex methods or CFD modelling.
Re =

uca

(A.4)

Where u is the local relative velocity, ca the local blade chord and the kinematic viscosity
of air.
In the case of VAWT as the blade element rotates the local velocity (figure A.2) changes
causing a variation in angle of attack at every revolution (figure A.4 on page 102). This
dynamic effect is dominant, especially in low tip speed ratios, even if the wind speed is
constant and implies different aerodynamic forces than the ones predicted by using the
quasi-steady airfoil polars. Many different models have been proposed in the literature
known as dynamic stall models, which are based on the physics that the boundary layer
adapts slowly to the new flow regime after an instantaneous change of the local flow

A.1 Aerodynamics Module

101

65

60

Vrel [m/s]

55

50

45

7.25 m/s Wind speed


10.25 m/s Wind speed
13.25 m/s Wind speed

40

35

50

100

150

200

azimuth [deg]

250

300

350

Figure A.3: Sandias 500 kW VAWT blade relative velocity fluctuation at equator over one
revolution for three different wind speeds at 28 rpm rotor speed [20].

conditions. In the current version of HAWC2 three dynamic stall models are implemented,
mainly developed for HAWTs, the Stig ye dynamic stall model [59], the Morten H.
Hansen modified Beddoes-Leishmann model for attached flow and stall dynamics [42,
48] and the Active Trailing Edge Flap (ATEFlap) model that accounts for steady and
dynamic effects of trailing edge flap deflections and can be used in the case of trailing
edge flap implementations [37]. The Stig ye model computes only the dynamics on the
lift coefficient, while the Beddoes-Leishmann approach reproduces the effects in drag and
moment coefficients, as well. Both models assume that the flow separation starts at the
trailing edge and extends towards the leading edge and do not model deep stall conditions.
This phenomenon however, might be present in a Darrieus VAWT rotor close to the blade
roots even under normal operating conditions and needs further investigation. The Stig
ye and the modified Beddoes-Leishmann dynamic stall models are used in this study
in VAWT and HAWT aerodynamic simulations, respectively. The lift, drag and moment
coefficients as a function of angle of attack for each profile and the chord and relative
thickness distribution along the blade span are defined in external files.
The difference in power output with and without dynamic stall model can bee seen in
figure A.5 on the next page, where the Sandia power curve is compared with the results
obtained using HAWC2 code with and without the Stig ye dynamic stall model. The
effect of dynamic stall on the lift coefficient can be seen in figure A.6 on page 103. The
results reveal that the dynamic stall model is very important for correct aerodynamic
simulations in VAWTs.
Another effect that arises in VAWTs is the flow curvature effect. As the blade rotates,
sees the flow curved and in turn the angle of attack for a given section at different chordwise locations varies acting as a virtual camber (figure A.7 on page 103). This effect
becomes important as the wind turbine solidity increases. As it is suggested in [40] and
implemented in HAWC2 the corresponding angle of attack at three quarter chord from
the leading edge is used 0.75 .
Finally, the effect of the centrifugal forces on the boundary layer [49], and as a result
to the airfoil polars that are used are not modelled in HAWC2 and could be examined

102

Modules of the Aeroelastic Code HAWC2

15

Angle of attack [deg]

10

10
7.25 m/s Wind speed
10.25 m/s Wind speed
13.25 m/s Wind speed

15

20

50

100

150

200

azimuth [deg]

250

300

350

Figure A.4: Sandias 500 kW VAWT blade angle of attack variation at equator over one
revolution for three different wind speeds at 28 rpm rotor speed [20].

500

Rotor Power [kW]

400

300

200

100

Sandia measurements
HAWC2 model ( with Dyn. stall)
HAWC2 model ( without Dyn. stall)

-100
4

10

12

14

16

18

20

22

Wind speed [m/s]

Figure A.5: Sandias 500 kW VAWT rotor power as function of wind speed at 34 rpm rotor
speed with and without the Stig ye dynamic stall model implementation [20].

further in the case of VAWTs. In HAWTs usually the airfoil polars are modified taking
into account this effect outside HAWC2 and then the corrected coefficients are used in
HAWC2 simulations.

A.2 Wind Module

103

Figure A.6: Sandias 500 kW VAWT blade lift coefficient at equator over one revolution for
different wind speeds at 28 rpm rotor speed. Stig ye dynamic stall model is included in the
HAWC2 simulations [20].

Figure A.7: Flow curvature effects on an airfoil in curved flow [40].

A.1.2

Aerodrag module

Except the blades which interact with the wind, it is possible in HAWC2 to capture the
aerodynamic drag forces of other components of the wind turbine, such as the tower and
the nacelle both in the case of horizontal and vertical axis wind turbine configurations.
This is done through a simple drag force calculation based on the components drag
coefficient and the projected area normal to the mean wind velocity direction.

A.2

Wind Module

In HAWC2 can be reproduced deterministic and stochastic wind fields. The wind conditions are defined in a general way and can be used in horizontal and vertical axis wind
turbine simulations.
Different wind shear profiles can be set such as constant, logarithmic, power law or
user defined. Furthermore, extreme coherent and extreme operating gusts according to
IEC 61400-1 ed.3 international standard can be generated and used in the simulations.
The Veers and Mann models can be implemented for stochastic wind simulations. The
Veers approach [78] is based on the creation of one-point turbulence field in the center

104

Modules of the Aeroelastic Code HAWC2

of the rotor and a Szinozouka method for the generation of the turbulence wind in the
other points of a polar grid using coherence functions (multivariate Fourier simulation)
at it can be seen in figure A.8 left. This formulation has the disadvantage that the
created fluctuations of the wind speed components in three dimensions u0 , v 0 and w0 are
uncorrelated.

Figure A.8: Turbulent wind field representations in HAWC2. Left the Veers approach
(multivariate Fourier simulation) represented by a polar grid. Right the Mann model
implementation creating a turbulence box [83].

The Mann model [46] creates a spatial vector field in Cartesian coordinates, based on the
solution of Navier-Stokes equations for the fluctuations u0 , v 0 and w0 of the flow. With
this formulation u0 , v 0 and w0 are correlated and a fully coherent 3D-turbulent wind field
is reproduced. This represents quite well the normal atmospheric turbulence with neutral
stability where the energy spectrum follows closely the Kaimal spectrum. Practically,
in HAWC2 the temporal and spatial turbulent wind vector field is generated using the
Taylors f rozen-turbulence approximation. As a result, a turbulence box is created
before the simulation, containing the mean and the fluctuating wind field components
u + u0 , v + v 0 and w + w0 (figure A.8 right). During the turbine time domain simulation
the box is moved through the rotor representing the real wind field. It can be also rotated
to represent skewed inflow and inclination of the flow. The Mann turbulence model is
recommended to be used for wind turbine simulations by the IEC 61400-1 ed.3 standard
and is the one used in this study.

A.2.1

Tower shadow module

The tower influence in the wind field is also modelled (Tower shadow model) and can be
included in the simulations. The effect upstream of the rotor is captured by the potential
flow solution around a cylinder, while downstream is based on the boundary layer solution
of an axisymmetric jet stream (jet model).
Finally, it is possible to simulate the wake of a wind turbine in a wind farm configuration
through the dynamic wake meandering model.

A.3

Structural Module

A brief description of the formulation of the structure of a wind turbine is given in this
section. All the components such as the blades, the tower and the hub are defined as

A.3 Structural Module

105

main bodies. Depending of the part and the required accuracy can be composed by one
or more sub-bodies, called multibody formulation. More details on the theory behind
this representation can be found in reference [69]. For example, due to high flexibility of
the blades which causes large deformations, using more bodies will lead to more accurate
results. The sub-bodies consist from beam elements based on Timoshenko beam theory
where the deformations are assumed small. Each element has its local coordinate system
and up to 6 degrees of freedom, 3 rotations and 3 translations in 3D space (figure A.9).

Figure A.9: Main body formulation in HAWC2, consisting by bodies and elements [83].

The structural properties of each main body are defined by 19 parameters written to an
input file. These are given along the curved length of the body and for the various centre
positions the origin of the local coordinate system is assumed to be at the geometric
centre of the section. The properties are:
Curved length
Center of gravity position
Shear center position
Elastic center position
Mass per length
Area moments of inertia with respect to principal axes
Radii of gyration with respect to principal axes
Cross-section area
Youngs and shear modulus
Shear factors
Structural pitch
The structural damping of every main body is set through the Rayleigh damping parameters which are proportional to mass and stiffness matrices. With these constants the
desired logarithmic decrement can be achieved.
Constrains are used to set the connection between bodies and body or bodies to the
ground. It should be mentioned that anisotropic materials can be used and converted
to the above 19 equivalent parameters through BECAS [8] or Abaqus FEA software,
but coupling effects in the results due to anisotropy can not be predicted in the current
version. Two examples of implementation of a HAWT model and a blade are depicted in
figures A.10 and A.11, respectively. The aerodynamic properties of the blade along the
span are modelled as was explained in subsection A.1.1 on page 100.

106

Modules of the Aeroelastic Code HAWC2

Figure A.10: Main bodies in HAWC2, consisting by sub-bodies with Timoshenko beam
elements, [83].

Figure A.11: The blade (main body) formulation, consisting by sub-bodies with Timoshenko
beam elements [83].

A.4

Generator

The electrical generator of a wind turbine can be implemented in HAWC2 as an external


DLL (Dynamic Link Library) file. Thus, there is the freedom of using different size and
kind of generators and simulate their dynamic response coupled with the other parts of
the turbine.

A.5 Controller

A.5

107

Controller

The controller of a wind turbine like the generator can be modelled and used through
a DLL file. In this way different control strategies for normal operation, start-up, shut
down or abnormal operation can be evaluated.

A.6

Hydrodynamics-Sea and Soil module

In the case of offshore wind turbine simulations HAWC2 can simulate the sea waves and
currents. The hydrodynamic loads are taken into account in the wind turbine foundation
using the Morison equation for calculations of the relative kinematics, drag and inertia
and added mass effects. Airy wave theory is used with modification for free surface effects
(e.g. Wheeler-stretching for shallow waters) both for deterministic and stochastic waves.
JONSWAP and Pierson-Moskowitz spectrum models can also be simulated for limited
fetch and fully developed sea irregular waves, respectively.
Soil module can be added when fixed foundations for offshore wind turbines, for instance
monopiles, jackets or tripods are used to account for the soil-foundation interaction. This
influences the damping of the system and boundary conditions of the foundation. Simple empirical engineering models have been implemented, such as the apparent fixity
length, the uncoupled springs and the distributed non-linear spring model using forcedisplacement curves (p-y curves) which have low computational time requirements. More
details about the hydrodynamics and soil module implementation can be found in reference [41].

A.7

Dynamic Wires module

Finally, a dynamic wires future has been developed as an external DLL and can be used
in the case of wind turbines with guy wires or mooring lines for offshore wind turbines.
The wires are implemented as non-linear beam elements with longitudinal flexibility and
no bending stiffness [36]. Drag and buoyancy forces can be simulated and concentrated
masses can be added on specific points on them. Two examples that used this module for
the representation of the mooring lines and the guy wires for the tower-top support are;
the DeepWind offshore VAWT [80] and the Sandia 34 m test-bed VAWT [20] aeroelastic
models, respectively.

108

Modules of the Aeroelastic Code HAWC2

Appendix B
VAWT Model-Complementary Data

More details regarding the properties of the DeepWind and the Sandia scaled VAWTs
are given in this appendix. In addition, details about the modified DeepWind rotor are
presented.

B.1

Blade Structural Properties


1400

5.5

Mass per Length [kg/m]

DeepWind blade
Sandia scaled blade

Chord [m]

4.5

3.5

1200
1000

800
600
400
200

20

40

60

80

100

Distance along blade from bottom to top [%]

20

40

60

80

100

Distance along blade from bottom to top [%]

Figure B.1: DeepWind and Sandia scaled blade chord (left) and mass per length (right)
distributions.

109

110

VAWT Model-Complementary Data

10 9

5
DeepWind blade
Sandia scaled blade

10 10

2.5

Edgewise Stiffness (EI ) [Nm 2]

Flatwise Stiffness (EI x) [Nm2]

2
1.5
1
0.5

0
0

20

40

60

80

100

20

40

60

80

100

Distance along blade from bottom to top [%]

Distance along blade from bottom to top [%]

Figure B.2: DeepWind and Sandia scaled blade flatwise (left) and edgewise (right) stiffness
distributions.

10 9

3.5
DeepWind blade
Sandia scaled blade

2.5

Longitudinal Stiffness (EA) [N]

Torsional Stiffness (GJ) [Nm2]

2
1.5
1
0.5
0

10 10

3
2.5
2
1.5
1
0.5
0

20

40

60

80

100

Distance along blade from bottom to top [%]

20

40

60

80

100

Distance along blade from bottom to top [%]

Figure B.3: DeepWind and Sandia scaled blade torsional (left) and longitudinal (right)
stiffness distributions.

B.2 Loads on the Modified DeepWind VAWT

B.2

111

Loads on the Modified DeepWind VAWT

0.62
16886

0.6

11193

13091

14989

0.58
5500

3602

7398

Rotor speed [rad/s]

0.56

9295

0.54

0.52

0.5

0.48

0.46

0.44
11193

14989

13091

16886

0.42
4

10

12

14

16

18

20

22

24

Wind speed [m/s]

Figure B.4: Contours of blade-1 low root maximum flatwise bending moment in kNm for the
modified DeepWind turbine. Wind speed from cut-in to cut-out and rotor speeds from 0.4 to
0.62 rad/s.

0.62
10746

24268
20887

0.6

14127

17507

3985
0.58

Rotor speed [rad/s]

0.56

0.54

0.52

0.5

0.48

0.46
14127

0.44

10746

17507
20887
24268

7366
0.42
4

10

12

14

16

18

20

22

24

Wind speed [m/s]

Figure B.5: Contours of blade-1 low root maximum edgewise bending moment in kNm for the
modified DeepWind turbine. Wind speed from cut-in to cut-out and rotor speeds from 0.4 to
0.62 rad/s.

112

VAWT Model-Complementary Data

0.62
10746

24268
20887

0.6

14127

17507

3985
0.58

Rotor speed [rad/s]

0.56

0.54

0.52

0.5

0.48

0.46
14127

0.44

10746

17507
20887
24268

7366
0.42
4

10

12

14

16

18

20

22

24

Wind speed [m/s]

Figure B.6: Contours of turbine base bottom maximum Mx bending moment in kNm for the
modified DeepWind turbine. Wind speed from cut-in to cut-out and rotor speeds from 0.4 to
0.62 rad/s.

0.62
257278
216611
0.6

297944
338610

94613
135279

175945

0.58

Rotor speed [rad/s]

0.56

0.54

0.52

0.5

0.48

0.46

0.44

135279

94613

175945
216611

53947
0.42
4

10

12

14

16

18

20

22

24

Wind speed [m/s]

Figure B.7: Contours of turbine base bottom maximum My bending moments in kNm for the
modified DeepWind turbine. Wind speed from cut-in to cut-out and rotor speeds from 0.4 to
0.62 rad/s.

Appendix C
Airfoil Polars

In this appendix are plotted the airfoil polars for the symmetric NACA 0018 and NACA 0025
profiles at 1.2x107 [-] Reynolds number as calculated at DeepWind report [80, p. 11].
1.8
C l NACA 0018

1.6

C l NACA 0025
C d NACA 0018

1.4

C d NACA 0025

C l and C d [-]

1.2
1
0.8
0.6
0.4
0.2
0
0

20

40

60

80

100

120

140

160

180

[deg]

Figure C.1: Airfoil lift and drag coefficients of NACA 0018 and NACA 0025 as a function of
angle of attack at Re=1.2x107 [80, p. 14].

113

114

Airfoil Polars

1.8
C l NACA 0018
C l NACA 0025
1.6

C d NACA 0018

1.5

C d NACA 0025
1.4
1

1.2
NACA 0018
NACA 0025
C d NACA 0018

C l [-]

C l and C d [-]

0.5

C d NACA 0025

0.8
-0.5
0.6

-1
0.4

-1.5

0.2

0
0

10

20

30

-2
0.005

40

0.01

0.015

[deg]

0.02

0.025

0.03

0.035

0.04

C d [-]

Figure C.2: Airfoil lift and drag coefficients of NACA 0018 and NACA 0025 from = 0 to
40 (left) and lift versus drag coefficients (right) at Re=1.2x107 [80, p. 14].

0.05
NACA 0018
NACA 0025

0
-0.05

C m [-]

-0.1
-0.15
-0.2
-0.25
-0.3
-0.35
-0.4
0

20

40

60

80

100

120

140

160

180

[deg]

Figure C.3: Airfoil moment coefficient of NACA 0018 and NACA 0025 as a function of angle
of attack at Re=1.2x107 [80, p. 14].

Appendix D
Complementary Results from the
DLCs Analysis

In this appendix are given some complementary results regarding the analysis of the
different design load cases.

D.1

Normal Power Production


900
VAWT

800

HAWT

Rotor Thrust [kN]

700
600
500
400
300
200
100
0

10

15

20

25

Wind Speed [m/s]

Figure D.1: Mean rotor thrust as a function of wind speed for vertical and horizontal axis wind
turbines.

115

116

10

Complementary Results from the DLCs Analysis

10 4

Mx [kNm]

With turbulence

0
-5
200

My [kNm]

Wind Speed [m/s]

Without turbulence

300

400

500

600

700

800

300

400

500

600

700

800

300

400

500

600

700

800

10 5

-1
200
15
10
5
0
200

Time [sec]

Figure D.2: Time series of VAWT base bottom bending moments (Mx and My ) for 7.5 m/s
mean wind speed with and without turbulence.

Mx [kNm]

0
-2000
-4000
-6000
200

With turbulence

Without turbulence

300

400

500

600

700

800

300

400

500

600

700

800

300

400

500

600

700

800

My [kNm]

5000
2500
0
-2500

Wind Speed [m/s]

200
15
10
5
0
200

Time [sec]

Figure D.3: Time series of VAWT blade low root flatwise Mx and edgewise My moments for
7.5 m/s mean wind speed with and without turbulence.

D.1 Normal Power Production

117

10 4

10 4
VAWT low-root
VAWT upper-root

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.5

1.5

0.5

0.5

0
0

10

15

20

25

10

Wind Speed [m/s]

15

20

25

Wind Speed [m/s]

Figure D.4: Equivalent 1 Hz fatigue blade upper and low root bending moments Mx and My as
a function of wind speed for the VAWT.
2

10 5

2
VAWT Iref = 0.14 [-]

1.8

HAWT I ref = 0.14 [-]

1.6

HAWT I ref = 0.00 [-]

Bending Moment M y [kNm]

Bending Moment M x [kNm]

1.8

VAWT Iref = 0.00 [-]

1.6

10 5

1.4
1.2
1
0.8
0.6

1.4
1.2
1
0.8
0.6

0.4

0.4

0.2

0.2

0
4

10

12

14

16

18

20

22

24

10

Wind Speed [m/s]

12

14

16

18

20

22

24

Wind Speed [m/s]

Figure D.5: Equivalent 1 Hz fatigue base bottom Mx and My bending moments as a function
of wind speed for the vertical and horizontal wind turbines with and without turbulence.
10 4

10 4
VAWT Iref = 0.14 [-]

VAWT Iref = 0.00 [-]


HAWT I ref = 0.00 [-]

Bending Moment M y [kNm]

Bending Moment M x [kNm]

HAWT I ref = 0.14 [-]

1.5

0.5

1.5

0.5

0
4

10

12

14

16

Wind Speed [m/s]

18

20

22

24

10

12

14

16

18

20

22

24

Wind Speed [m/s]

Figure D.6: Equivalent 1 Hz fatigue blade root Mx and My bending moments as a function of
wind speed for the vertical and horizontal wind turbines with and without turbulence. The blade
low root is given for the VAWT.

118

Extreme Operating Gust Case

Mx

[kNm]

10 5

25

1
20
0

V hub

20

-1

-1
220
2

25
My

V hub

[m/s]

10 5

[m/s]

[kNm]

D.2

Complementary Results from the DLCs Analysis

230

240

250

260

10 5

15
270

-2
220

230

240

250

260

10 5

25

15
270
25

20

-1

-1
220
2

[m/s]

20

[m/s]

[kNm]

[kNm]

10

230

240

250

260

15
270

-2
220

10

25

230

240

250

260

15
270

25

Rotor Azimuth Angle [deg]

-1
220

20

[m/s]

0
-1

230

240

250

260

15
270

300

200

100

0
220

[m/s]

20
0

[kNm]

-2
220

Rotor Azimuth Angle [deg]

[kNm]

230

240

250

Time [sec]

260

270

230

240

250

260

15
270

230

240

250

260

270

300

200

100

0
220

Time [sec]

Figure D.7: Time series of VAWT base bottom bending moments Mx (left) and My (right)
during the EOG at 19.5 m/s wind speed. The gust frontal passage passes the rotor plane at 3
different time stamps (corresponding to different rotor orientations), each one represented at the
1st , 2nd and 3rd row of sub-plots.

D.3 Extreme Wind Direction Change Case

5000

119

25
Mx

25

V hub

My

15000

V hub

10000

-15000
220

20

5000

-10000

[m/s]

20

[m/s]

-5000

[kNm]

[kNm]

0
230

240

250

260

5000

15
270

-5000
220

230

240

250

260

25

15
270
25

15000
10000

-15000
220

20

5000

-10000

[m/s]

20

[m/s]

-5000

[kNm]

[kNm]

0
230

240

250

260

5000

15
270

-5000
220

230

240

250

260

25

15
270
25

15000
10000

0
230

240

250

260

15
270

300

200

100

0
220

-5000
220

Rotor Azimuth Angle [deg]

Rotor Azimuth Angle [deg]

-15000
220

20

5000

-10000

[m/s]

20

[m/s]

-5000

[kNm]

[kNm]

230

240

250

260

230

240

250

260

15
270

230

240

250

260

270

300

200

100

0
220

270

Time [sec]

Time [sec]

Figure D.8: Time series of VAWT blade-1 low root bending moments Mx (left) and My (right)
during the EOG at 19.5 m/s wind speed.The gust frontal passage passes the rotor plane at 3
different time stamps (corresponding to different rotor orientations), each one represented at the
1st , 2nd and 3rd row of sub-plots.

D.3

Extreme Wind Direction Change Case


10 4
4

3.5

Bending Moment ||M2x +M y ||2 [kNm]

Bending Moment ||M2x +M y ||2 [kNm]

10 5
3.5

2.5

1.5

0.5

Normal power production


Positive wind direction change
Negative wind direction change

2.5

1.5

0.5

0
0

10

15

Wind Speed [m/s]

20

25

10

15

20

25

Wind Speed [m/s]

Figure D.9: Comparison of ultimate loads emerging from positive and negative EDC with the
maximum values during normal wind conditions. Turbine base bottom and blade low root bending
moment magnitudes left and right respectively.

120

D.4

Complementary Results from the DLCs Analysis

Parked Standing Still Rotor

2.5

10 5

Torsional Moment |Mz| [kNm]

Maximum
STD
2

1.5

0.5

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure D.10: Maximum and standard deviation of the braking torque expressed at global
coordinate system. The wind turbine is parked in standing still mode at different rotor
orientations and exerted to 50 year recurrence period extreme winds.

Bending Moment ||M 2x +M 2y || [kNm]

14

10 5
Maximum
STD

12

10

0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure D.11: Maximum and standard deviation of the turbine bottom bending moments
expressed at the global coordinate system. The wind turbine is parked in standing still mode at
different rotor orientations and exerted to 50 year recurrence period extreme winds (Dynamic
stall model is deactivated).

D.4 Parked Standing Still Rotor

16

121

10 4
Blade-1 low root (max)
Blade-2 low root (max)
Blade-1 low root (std)
Blade-2 low root (std)

Bending Moment ||M2x +M 2y || [kNm]

14
12
10
8
6
4
2
0
0

20

40

60

80

100

120

140

160

180

Rotor Orientation [deg]

Figure D.12: Maximum and standard deviation of blade one and two low roots bending
moments expressed at local coordinate systems. The wind turbine is parked in standing still mode
at different rotor orientations and exerted to 50 year recurrence period extreme winds (Dynamic
stall model is deactivated).

10 5

10 4
Turbine bottom (maximum)
Turbine bottom (STD)

Bending Moment ||M 2x +M 2y || [kNm]

12

10

10

Blade-1 low root (max)


Blade-2 low root (max)
Blade-1 low root (std)
Blade-2 low root (std)

12

0
0

50

100

Rotor Orientation [deg]

150

50

100

150

Rotor Orientation [deg]

Figure D.13: Maximum and standard deviation of the turbine bottom (left) and blade roots
(right) bending moments expressed at the global and local coordinate systems respectively. Wind
turbine is parked in standing still mode at different rotor orientations and exerted to 1 year
recurrence period extreme winds (Dynamic stall model is deactivated).

122

Complementary Results from the DLCs Analysis

Appendix E
NREL 5 MW HAWT Properties

The basic properties of the NREL 5 MW, 3-bladed reference wind turbine are summarized
in this section (table E.1) for completeness as reported in [35].
Table E.1: Operational and geometric data of the NREL 5 MW horizontal axis wind turbine.

Property

Value

Rated electrical power


Rated rotor speed
Rated wind speed
cut-in wind speed
cut-out wind speed
Rated tip speed
Hub height
Number of blades
Rotor diameter
Hub diameter
Swept area
Hub inertia on LSS
Rotor inertia (blades and hub) on LSS
Generator inertia on HSS
Gearbox ratio
Rotor mass
Blade mass
Hub mass
Nacelle mass
Tower mass

5
12.1
1.27
11.4
3
25
80
90
3
126
3
12445
115.93x103
35.444x106
534
97:1
110000
17740
56780
240000
347460

[MW]
[rpm]
[rad/s]
[m/s]
[m/s]
[m/s]
[m/s]
[m]
[-]
[m]
[m]
[m2 ]
[kgm2 ]
[kgm2 ]
[kgm2 ]
[-]
[kg]
[kg]
[kg]
[kg]
[kg]

In figures E.1 and E.2 on the following page are illustrated the schematic of the wind
turbine and the coordinate systems where the loads are expressed. Turbine base (or

123

124

NREL 5 MW HAWT Properties

equivalent tower bottom) loads are always given in the fixed global coordinate system
XY ZG while blade root loads are expressed at local blade coordinate systems which
rotate with the blades.

Figure E.1: NREL 5 MW, 3-bladed horizontal axis wind turbine schematic.

Figure E.2: Definition of coordinate systems. The global coordinate system is fixed to the
ground (tower bottom), while blade-1, 2 and 3 coordinate systems are located at blade-1, 2 and 3
roots respectively and rotate with the blades. Positive wind direction is towards yglobal , [83].

You might also like