0% found this document useful (0 votes)
86 views

Stevenvaneck Thesis Final PDF

Uploaded by

ersin
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
86 views

Stevenvaneck Thesis Final PDF

Uploaded by

ersin
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 208

Solar updraft tower

Structural optimisation under dynamic


wind action

by

S.R.G. van Eck

in partial fulfilment of the requirements for the degree of

Master of Science
in Civil Engineering

at the Delft University of Technology,

An electronic version of this thesis is available at


http://repository.tudelft.nl/.
Student: Steven Roel Gerardus van Eck
Menno ter Braaklaan 1, 2624 TA Delft
Student ID: 1360477
E: [email protected]
T: (+31) (0) 6 14 72 68 15

Thesis Committee: Chairman:


Prof. dr. ir. J.G. Rots
Delft University of Technology
E: [email protected]
T: (+31) (0) 15 278 3799

Daily supervisor:
dr. ir. J.L. Coenders
Delft University of Technology
BEMNext Laboratory
E: [email protected]

ir. C. Kasbergen
Delft University of Technology
E: [email protected]
T: (+31) (0) 15 278 2729

ir. R.M.J. Doomen


Pieters Bouwtechniek
E: [email protected]
T: (+31) (0) 15 219 0300
Each problem that I solved became a rule
which served afterwards to solve other
problems.
– René Descartes
Preface
This Master’s thesis was written as part of the Master curriculum of Building
Engineering with a specialisation in Structural Design. Building Engineering is
a Master track of Civil Engineering at the Delft University of Technology. The
research was carried out in collaboration with the Faculty of Civil Engineering and
Geosciences and the Dutch engineering company ’Pieters Bouwtechniek’.
I would like to acknowledge everyone who played a vital role in the completion of
my thesis work.
First of all, I would like to express my gratitude towards dr. ir. Jeroen Coenders, my
daily supervisor at the Delft University of Technology, for inspiring me to pursue
a graduation topic in this field and for his continuous support and advice.
The other person who was vital in the completion of my Master’s thesis is ir. Rob
Doomen, my supervisor at Pieters Bouwtechniek. His genuine interest in the topic
and his engineering experience always provided me with new insights during our
weekly meetings and ensured I never got too far off track.
I would also like to thank my other two committee members. Prof. dr. ir. Jan
Rots, for chairing my committee and keeping an eye on the general organisation,
and ir. Cor Kasbergen, for helping me with all of my issues pertaining to finite
element modelling.
My gratitude goes out to Pieters Bouwtechniek, for providing me with a pleasant
working environment and the resources needed to complete my Master’s thesis.
I would like to thank ir. Hein Brakel in particular, for paving the way for this
graduation topic within Pieters Bouwtechiek.
Internationally my thanks goes out to Prof. Gideon van Zijl from the University
of Stellenbosch, Dr.-Ing. Arndt Goldack from the Technische Universität Berlin,
and Dipl.-Ing. Markus Tschersich from the Bergische Universität Wuppertal, all of
whom provided me with valuable information which benefited me greatly during
my research.
Lastly, on a more personal note, I would like to thank my parents, my sister, my
friends and my colleagues at Pieters Bouwtechiek and BEMNext for their motiva-
tion and support.

S.R.G. van Eck


Delft, December 2014

iii
Abstract
As fossil fuel reserves are rapidly being depleted, sustainable alternatives have
to be found to fulfil the world’s energy demand. Numerous concepts have been
proposed to generate electricity by harnessing renewable energy sources such as
solar or wind power. One of these concepts is the the so-called solar updraft tower
(SUT). The SUT consists of three elements: a solar air collector, wind turbines and
a chimney. The taller the chimney, the larger the stack effect and thus the more
energy which can be generated by the turbines. The proposed concepts for this
chimney schematise it as a reinforced concrete cylindrical shell, with the bottom
half shaped like a hyperboloid and the top half as a flared cylinder, outfitted with
ten stiffening rings evenly distributed over the height. Chimneys as tall as 1500m
have been proposed, and, previous research shows that these tall structures have
very low eigenfrequencies which come very close to the peak of the wind power
spectrum. This makes them extremely vulnerable to resonance induced by storm
actions.
Two types of resonance can be distinguished in these structures; along-wind reso-
nance, and across-wind resonance. Along-wind resonance is caused by turbulence
in along-wind gusts. The second type, across-wind resonance, is caused by the
alternating shedding of vortices. This leads to pulsating excitation forces in the
across-wind direction, and, if the frequency of the vortex shedding is the same as
one of the eigenfrequencies of the chimney, resonance will occur.

Figure 1: Pre-existing design for a 1000m tall SUT (courtesy of Krätzig & Partner, Bochum)
(Niemann et al., 2009)

In this thesis, a finite element model is created based the pre-existing design shown
in Figure 1. This so-called base model is then analysed to determine which key
problem areas could benefit from improvement. The analyses show that especially
the first two eigenfrequencies are critical for along-wind resonance as well as across-
wind resonance. These eigenfrequencies are seen as two individual problem areas
as improvements to one eigenfrequency not necessarily guarantee improvements to
the second eigenfrequency. Furthermore, tension on the windward side leads to
cracks in the stiffening rings which negatively influence the eigenfrequencies and
thus the dynamic response. The last area which could benefit from optimisation
is the cost of the chimney; an optimal solution does not use more material than

iv
Abstract

necessary.
A design tool called SUMAT (Solar Updraft Modal Analysis Tool) is created which
enables the user to analyse multiple chimney configurations at once, subsequently
being able to compare their results. Various sensitivity analyses are carried out
to determine the influence of geometric and material parameters on the four key
problem areas of the chimney. A multi-objective optimisation process is followed
to optimise each of the key problem areas, ie. objective functions, by hand. The
first step in optimising the structure is to subdivide the parameters which were
researched into four categories, depending on their usefulness. The second step
consists of gradually introducing these parameter changes into the base model.
The optimisation process revealed that the objective functions can be maximised as
follows: increasing the moment of inertia of the rings by changing their aspect ratio
ensures that the chimney is fully loaded in compression. An increase in the throat
height further improves the reduction of tension on the windward side and the first
eigenfrequency. A reduction in wall thickness at the top of the chimney improves
the first eigenfrequency while also reducing material use. Lastly, it appears that the
stiffening rings at the bottom serve little to no purpose. Removing them leads to a
reduction in material use while some of the material gained can be used to increase
the dimensions of the top rings, consequently improving the second eigenfrequency
and reducing tension.
More thorough analyses revealed that the optimisation process has indeed led to
an overall improved structure when compared to the original base model. While
along-wind resonance does not pose as great a threat as was initially assumed, due
to the influence of aerodynamic admittance, the results do show that the improved
eigenfrequencies led to a smaller increase in deflection as a result of dynamic wind
action. Vortex shedding also no longer poses a threat as the improved second
eigenfrequency resulted in critical wind speeds which are much larger than could
ever occur at the chosen reference location. Future optimisations should therefore
focus more heavily on the second eigenfrequency than on the first, assuming that
the accompanying mode shapes stay the same.

v
Contents
Abstract iv
Acronyms ix
1 Introduction 1
1.1 Problem definition . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Domain and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Material choice . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.3 Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Research objective . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Research questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4.1 Main research question . . . . . . . . . . . . . . . . . . . . . 5
1.4.2 Sub research questions . . . . . . . . . . . . . . . . . . . . . 5
1.5 Hypothesis and expectations . . . . . . . . . . . . . . . . . . . . . . 5
1.6 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 Geometry of the chimney 9
2.1 Reference chimney . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Hyperboloid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4 Matching conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.5 Stiffening rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3 Actions working on the chimney 17
3.1 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3 Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3.1 Mean wind profile. . . . . . . . . . . . . . . . . . . . . . . . 19
3.3.2 Reference hourly mean wind speed . . . . . . . . . . . . . . 19
3.3.3 Wind pressure . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3.4 Pressure coefficients . . . . . . . . . . . . . . . . . . . . . . 22
3.3.5 Density of air . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4 Reference location . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.5 Load case for the chimney . . . . . . . . . . . . . . . . . . . . . . . 25
3.6 Dynamic wind loading . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.6.1 Along-wind oscillations . . . . . . . . . . . . . . . . . . . . . 26
3.6.2 Across-wind oscillations . . . . . . . . . . . . . . . . . . . . 31
3.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4 Finite element model 35
4.1 Discretisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.1.1 Software choice . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.1.2 Element choice . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1.3 Element sizing . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.1.4 Mesh refinement . . . . . . . . . . . . . . . . . . . . . . . . 38
4.1.5 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Material properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2.1 Concrete. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.2.2 Reinforcement steel . . . . . . . . . . . . . . . . . . . . . . . 40

vi
Contents

4.2.3 Stiffening rings . . . . . . . . . . . . . . . . . . . . . . . . . 41


4.3 Convergence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.4 Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.5 Linear vs nonlinear analysis . . . . . . . . . . . . . . . . . . . . . . 44
4.5.1 Quasi nonlinear analysis . . . . . . . . . . . . . . . . . . . . 45
4.5.2 Fictional Young’s modulus . . . . . . . . . . . . . . . . . . . 46
4.6 Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.7 Foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.8 Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.9 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5 Analysing the base model 53
5.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.2 Summary of the model set-up . . . . . . . . . . . . . . . . . . . . . 55
5.2.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2.2 Mesh. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2.3 Material properties and actions . . . . . . . . . . . . . . . . 55
5.3 Static analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.3.1 In-plane forces . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.3.2 Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.3.3 Quasi nonlinear static analysis . . . . . . . . . . . . . . . . . 58
5.3.4 Reinforcement. . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.3.5 Cracking. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.3.6 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.3.7 Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.4 Dynamic analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.4.1 Modal analysis . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4.2 Quasi nonlinear modal analysis . . . . . . . . . . . . . . . . 62
5.4.3 Harmonic analysis . . . . . . . . . . . . . . . . . . . . . . . 63
5.4.4 Response spectrum . . . . . . . . . . . . . . . . . . . . . . . 64
5.4.5 Vortex shedding. . . . . . . . . . . . . . . . . . . . . . . . . 64
5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6 Creating a design tool: SUMAT 67
6.1 The need for a design tool . . . . . . . . . . . . . . . . . . . . . . . 67
6.2 Input parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.3 APDL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.4 Batch mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.5 Output of results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.6 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.6.1 Geometrical shortcomings . . . . . . . . . . . . . . . . . . . 70
6.6.2 Material shortcomings . . . . . . . . . . . . . . . . . . . . . 71
6.6.3 Loading shortcomings . . . . . . . . . . . . . . . . . . . . . 72
6.6.4 Accuracy of results . . . . . . . . . . . . . . . . . . . . . . . 72
6.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
7 Sensitivity analysis 75
7.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2 Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.3 Geometric parameters . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.4 Material parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 81
7.5 Loading parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . 83
7.6 Damping parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 84
7.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
8 Optimisation of the base model 89
8.1 Multi-objective optimisation . . . . . . . . . . . . . . . . . . . . . . 89
8.2 Solution path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

vii
Contents

8.3 Categorising the parameters . . . . . . . . . . . . . . . . . . . . . . 91


8.3.1 Primary parameters . . . . . . . . . . . . . . . . . . . . . . 91
8.3.2 Secondary parameters . . . . . . . . . . . . . . . . . . . . . 92
8.3.3 Tertiary parameters . . . . . . . . . . . . . . . . . . . . . . 92
8.3.4 Residual parameters . . . . . . . . . . . . . . . . . . . . . . 92
8.4 Optimisation step 1. . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.5 Optimisation step 2. . . . . . . . . . . . . . . . . . . . . . . . . . . 94
8.6 Optimisation step 3. . . . . . . . . . . . . . . . . . . . . . . . . . . 96
8.7 Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
9 Analysing the optimised model 101
9.1 Comparison method . . . . . . . . . . . . . . . . . . . . . . . . . . 101
9.2 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
9.3 Static analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
9.3.1 In-plane forces . . . . . . . . . . . . . . . . . . . . . . . . . 103
9.3.2 Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . 104
9.3.3 Quasi nonlinear static analysis . . . . . . . . . . . . . . . . . 105
9.3.4 Buckling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
9.4 Dynamic analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
9.4.1 Modal analysis . . . . . . . . . . . . . . . . . . . . . . . . . 106
9.4.2 Quasi nonlinear modal analysis . . . . . . . . . . . . . . . . 106
9.4.3 Harmonic analysis . . . . . . . . . . . . . . . . . . . . . . . 107
9.4.4 Response spectrum . . . . . . . . . . . . . . . . . . . . . . . 108
9.4.5 Vortex shedding. . . . . . . . . . . . . . . . . . . . . . . . . 109
9.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
10 Discussion 111
10.1 Geometry of the chimney. . . . . . . . . . . . . . . . . . . . . . . . 111
10.2 Actions working on the chimney . . . . . . . . . . . . . . . . . . . . 111
10.3 Finite element model . . . . . . . . . . . . . . . . . . . . . . . . . . 112
10.4 Results of the base model . . . . . . . . . . . . . . . . . . . . . . . 112
10.5 The need for a design tool . . . . . . . . . . . . . . . . . . . . . . . 112
10.6 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
10.7 Optimisation process and verification of the results . . . . . . . . . 113
11 Conclusions and recommendations 115
11.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
11.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
Bibliography 119
A Literature review 123
A.1 Solar updraft tower . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
A.2 Structural principles . . . . . . . . . . . . . . . . . . . . . . . . . . 130
A.3 Construction process . . . . . . . . . . . . . . . . . . . . . . . . . . 133
A.4 Comparison of vertical wind-profiles. . . . . . . . . . . . . . . . . . 137
A.5 Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
B Sensitivity analysis 145
B.1 Geometric parameters . . . . . . . . . . . . . . . . . . . . . . . . . 146
B.2 Material parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 159
B.3 Loading parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . 166
B.4 Damping parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 170
C SUMAT user manual 173
D List of Figures 186
E List of Tables 191

viii
Acronyms
ASCE American Society of Civil Engineers
AMSL Above Mean Sea Level
APDL ANSYS Parametric Design Language
CFPP Coal-Fired Power Plant
EC Eurocode
FE(A)/(M) Finite Element (Analysis)/(Method)
(G)UI (Graphical) User Interface
(H)DM (Human) Decision Maker
ISO International Organisation for Standardisation
LEC Levelised Electricity Cost
MTOE Mega Tonnes of Oil Equivalent
NDCT Natural Draft Cooling Tower
PSD Power Spectral Density
RC Reinforced Concrete
SUMAT Solar Updraft Modal Analysis Tool
SUT Solar Updraft Tower
UHPFRC Ultra High Performance Fiber Reinforced Concrete

ix
Introduction
1
With the decline of fossil fuel reserves, there has been an ongoing search for more
sustainable alternatives. Numerous concepts have been proposed to generate elec-
tricity by harnessing renewable energy sources such as solar or wind power. Desert
areas are, thanks to their high annual solar irradiation, the perfect candidate for
housing large scale affordable solar power plants. The problem is that most types
of solar power plants require large amounts of cooling water; a scarce commodity
in the desert. A concept which does not require cooling water is the so-called solar
updraft tower (SUT) (Schlaich et al., 2013). The solar updraft tower consists of
three elements (Figure 1.1), described below:

Figure 1.1: Working principle (Krätzig et al., 2009)

Solar air collector A large circular transparent roof functions as the collector
area. The collector area is open at the perimeter so that cold air can enter.
The transparent roof allows solar radiation to heat up the air underneath the
roof and trap it there, functioning like a greenhouse.

1
1. Introduction

Chimney/Tower The heated air exits the solar air collector through the chimney
as a result of air buoyancy. The taller the chimney, the greater the thermal
difference and thus the greater the buoyancy force.

Turbines The airflow caused by the air collector and chimney drives pressure-
staged turbines at the base of the chimney. The mechanical energy in the
turbines is then converted into electrical energy by conventional generators.

The taller the chimney, the larger the stack effect and thus the more energy can be
generated by the turbines. Chimneys as tall as 1500 meters have been proposed,
which, naturally, come with a whole slew of structural issues. The proposed con-
cepts for this chimney schematise it as a reinforced concrete cylindrical shell, with
the bottom half shaped like a hyperboloid and the top half as a flared cylinder
(Figure 1.2). Wind is the dominant action for such a tall and relatively slender
structure and since a structure of such height has never been executed before there
are no simplified load assumptions available which would lead to a safe design
(Niemann et al., 2009). Furthermore, most wind design profiles are only valid for
altitudes up to 200m since there is a lack of meteorological data for wind speeds
above 300m (Harte and Van Zijl, 2007).

Figure 1.2: Pre-existing design for a 1000m tall SUT (courtesy of Krätzig & Partner, Bochum)
(Niemann et al., 2009)

Perhaps the biggest obstacle standing in the way of the solar updraft tower is the
high cost of it. A small scale power plant will not be economically viable while a
large one will require huge amounts of capital which scares off investors, since the
technology has not been proven on a large scale. A 200 MW coal-fired power plant
costs around e300 million to construct while a solar updraft tower with a similar
energy production will cost around e750 million. Solar updraft towers, however,
have a very long lifespan and do not require costly fuels to operate, making them
economically attractive in the long run. Note that more background information
on the solar updraft tower can be found in Appendix A.1.

2
1.1. Problem definition

1.1. Problem definition

Previous studies have shown that the eigenfrequencies of such a tall chimney are
very low and come very close to the peak of the wind power spectrum (Harte
and Krätzig, 2011). This makes these chimneys extremely vulnerable to resonance
induced by storm actions. Not only that, dynamic loading causes fatigue which
can become troublesome in a structure with a design working life of 100 years. It
has to be noted that the eigenfrequencies are too low to fall into the range of the
earthquake power spectrum (Figure 1.3).

Economically speaking, an extremely thin concrete shell would be the best solu-
tion, but the thinner the shell gets, the harder it becomes to counteract the wind
forces and evenly distribute the forces around the circumference. Several stiffening
measures have been proposed to improve the structural behaviour of such a tall
chimney and make it act more like a cantilever beam instead of a shell structure. A
stiffening measure can be the application of several rings beams over the height of
the tower, the introduction of spoked wheels inside the cross-section of the chimney,
or a combination of both.

Research into the dynamic behaviour of a stiffened chimney has already been per-
formed (Alberti, 2006; Rousseau, 2005), but this was for a fixed geometry of a
1500m tall chimney. The current state is that there is only limited knowledge on
how material and geometric parameters influence this dynamic behaviour and sug-
gests that there is still a lot of room left for the optimisation of these parameters.

Figure 1.3: Dynamic excitation frequencies of structures by wind and earthquake (Holmes, 2007)

1.2. Domain and scope

The domain and scope of this research is limited by the following assumptions and
boundary conditions.

3
1. Introduction

1.2.1. Material choice


The material for the chimney will be reinforced concrete as studies have shown it to
be the most durable and economical choice for almost all of the proposed locations
(Lohaus and Abebe, 2010). The chimney will have to be realised using conventional
construction methods for concrete, with in-situ concrete and formwork, because
prefab concrete is not a feasible option due to the immense dimensions of the
structure (Helmus et al., 2010).

1.2.2. Dimensions
The height of the chimney will be fixed at 1000m; literature shows that this height
has the lowest cost per kW installed and is therefore the most financially attrac-
tive solution. Furthermore, a lot of research has already been conducted on the
structure of a 1000m tall chimney, which allows for a more precise validation of the
found results. Some research has been conducted on a 1500m tall chimney, but for
now a chimney of this height is considered a utopia (Krätzig et al., 2013).

1.2.3. Assumptions
The chimney will be modelled as one continuous structure without the large holes
required for turbines at the base. The effects of these holes on the structural
stability and possible solutions to mitigate these effects are considered to be beyond
the scope of this research. Furthermore, concerning the base fixity of the chimney,
the foundation will be modelled as completely fixed. The validity of this approach
is discussed in Chapter 4.

1.3. Research objective


The objective of this thesis is to optimise the proposed design for the chimney
structure of a solar updraft tower under dynamic wind action. The underlying
process of this optimisation is to tune the eigenfrequencies of the structure to
minimise resonance effects in the structure, without negatively influencing the static
response of the structure (Bachmann et al., 1995). As discussed, some research has
been conducted on the dynamic behaviour of a tall chimney, but none of these
studies have extensively examined the influence of various geometric and material
parameters.
The optimisation procedure can be defined as follows (Figure 1.4): first a base
model will be created in a finite element software package, the results of which will
be validated using results from earlier studies. Then, once the modelling approach
is deemed correct, the static and dynamic response of the base model is thoroughly
analysed to identify key problem areas. Using this knowledge, a design tool will
be built which will allow for several configurations of the chimney to be analysed
simultaneously. This design tool can then be used to conduct sensitivity analyses
for these key problem areas. Once the sensitivity of all these parameters is known,
an optimisation step can take place.
The first step of the optimisation process is to place all of the parameters which
were researched into four categories: primary parameters which bring about pos-
itive change in all four key problem areas, secondary parameters which positively
influence three out of four key problem areas, tertiary parameters which only pos-
itively influence one or two out of four key problem areas, and lastly residual
parameters which should not be tampered with.

4
1.4. Research questions

The second step consists of gradually introducing the primary, secondary and ter-
tiary parameter changes into the base model. The tertiary parameter changes will
only be introduced if one of the key problem areas has not improved compared to
the base model, or worse, has deteriorated due to the introduction of secondary
parameter changes. The results of the final, optimised structure will then be com-
pared to the original base model to verify that all key problem areas have been
optimised.

Create Verify Analyse Create Conduct Optimise


base base base design sensitivity base
model model model tool analysis model

Figure 1.4: Flowchart depicting the research methodology

1.4. Research questions


1.4.1. Main research question
The main research question, as derived from the previously defined objective, is as
follows:
How can the proposed structure for the chimney of a solar updraft tower be optimised
for static and dynamic wind action, without increasing the cost?

1.4.2. Sub research questions


The sub research questions focus on various aspects of the main research question.
The chapters in which these questions are answered are denoted in parentheses.
1. How can the geometry of the chimney be defined? (Chapter 2)
2. What characterises wind actions in tall, slender, cylindrical structures? (Chap-
ter 3)
3. How can the design be translated to a finite element model? (Chapter 4)
4. What are the key problem areas in the current design? (Chapter 5)
5. What are the requirements for a design tool to analyse chimney structures?
(Chapter 6)
6. What is the influence of geometric and material parameters on the static
response, the dynamic response, and the cost of the chimney? (Chapter 7)
7. Which parameter configurations improve the chimney structure, in terms of
static and dynamic response, and cost? (Chapter 8)

1.5. Hypothesis and expectations


Based on earlier research, the stiffening rings will be the most critical aspect in
the design of the chimney for a solar updraft tower. The structural integrity of
the chimney hinges on the stiffness, placement, and amount of these rings. It is
expected that most of the gains in the optimisation process can be achieved by
tweaking the parameters linked to these rings. The shape and material parameters
of the shell structure itself will most likely take a back seat in the optimisation of
the chimney.

5
1. Introduction

Tension on the windward side of the shell is a problem which needs to be dealt
with. The reduction in stiffness of the structure as a result of these tension forces
can significantly increase the dynamic response of the chimney. Increasing the
dead loads of the structure to mitigate the tension forces of the wind, applying
pretensioning in the rings, or increasing the stiffness of the rings to improve the
circumferential load distribution could be possible solution paths. Lastly, based on
the calculation methods used in many design codes, the first eigenfrequency will
be the most critical in terms of dynamic wind loading.

1.6. Thesis outline


The main report is divided into eleven chapters. The topics which are addressed
in those chapters are described in this section.
Chapter 2: Geometry of the chimney
Chapter 2 describes how the geometry for the chimney can be defined. The for-
mulae for the hyperboloid and the flared cylinder will be discussed here, as well
as matching conditions for both shapes. Furthermore, the wall thickness profile of
the shell will be defined and the placement of the stiffening elements is chosen.
Chapter 3: Actions working on the chimney
Dead loads and quasi-static wind loads can be determined from the codes. A
suitable vertical wind profile has to be found which can be applied to extremely
tall structures. The pressure distribution around the circumference can be derived
from the codes for natural draft cooling towers. Lastly, a summary is given of
dynamic wind actions which could affect tall, slender structures.
Chapter 4: Finite element model
Once the geometry and actions are determined, a finite element model can be
built. Firstly, the base model will be defined and tested for which mesh density the
solution converges. Subsequently, a modified base model, similar to the chimney
which is described in other studies, will be verified by comparing the results to those
found in the aforementioned literature. Analysis procedures for quasi nonlinear
material behaviour and buckling wil be described. Furthermore, the influence of
soil conditions on the base fixity is researched. Lastly, a constant damping ratio
for the chimney will be specified.
Chapter 5: Analysing the base model
In Chapter 5, all of the outcomes regarding the definition of the base model will be
summarised. The static and dynamic response of the base model will be analysed
to define key problem areas which could benefit from improvement.
Chapter 6: Creating a design tool: SUMAT
SUMAT (Solar Updraft Modal Analysis Tool) is a design tool to quickly study
several different chimney configurations. The reasons for designing such a tool and
its inner workings will be discussed. Additionally, attention will be given to the
limitations of the design tool.
Chapter 7: Sensitivity analysis
Chapter 7 depicts the process which was followed to determine the influence of the
geometric and material parameters on the static and dynamic response, and the
cost of the chimney. Each parameter set consists of five chimney configurations.
One configuration represents the original base model and the other four represent
similar configurations where only the studied parameter is varied. These parameter
variations are chosen in such a way that they still exist in the domain of feasibility,
otherwise there is little use in researching them.

6
1.6. Thesis outline

Chapter 8: Optimisation of the base model


The results of the sensitivity analysis are used to optimise the base model manually
using the aforementioned design tool. The first step of the optimisation process is
to determine the usefulness of all of the parameters which were researched. The
second step consists of gradually introducing these parameter changes into the base
model.
Chapter 9: Analysing the optimised model
In this chapter, the optimised model from Chapter 8 will be thoroughly analysed.
The results will be compared to those of the base model from Chapter 5. The cost
increase of each of the optimisation steps will be discussed and weighed against the
improvement in static and dynamic behaviour. Attention will also be given to the
validity and shortcomings of the chosen optimisation method.
Chapter 10: Discussion
The conclusions from each individual chapter are discussed here in light of the
results found in the optimisation step.
Chapter 11: Conclusions and recommendations
Conclusions are drawn about the conducted research and suggestions for future
research directions are given.

7
Geometry of the chimney
2
In this chapter the geometry of the chimney will be defined. Formulae will be given
for the definition of the hyperboloid and the flared cylinder, as well as matching
conditions. The default values for the geometric parameters of the base model
will also be given. Additionally, the wall thickness profile and the placement of
stiffening elements for the base model is described.

2.1. Reference chimney


The geometry of the base model which is used in this research is based on pre-
existing designs for a SUT. Figure 2.1.a, originally created by Krätzig & Partners,
shows one of the designs for a 1000m tall SUT.
The design schematises the chimney as a concrete shell, that starts off as a hy-
perboloid at the bottom and then gradually transforms into a flared cylinder (Fig-
ure 2.1.b). The large base of the hyperboloid shape will help with resisting the
base overturning moment caused by the wind load over the height of the chimney.
The reason for implementing a flared cylinder has to do with the flow characteris-
tics of the chimney. Over the height of the chimney, the density of the air flowing
through it decreases, which causes the air to expand. If the diameter of the chim-
ney stays constant, the airflow has to accelerate which subsequently leads to a
drop in pressure and therefore losses in the energy conversion process. By enlarg-
ing the cross-sectional area of the top of the chimney by around 14% (Backström
and Gannon, 2000), this effect can be mitigated.
Several variations of the above design exist, with throat heights ranging from 350m
to 450m. The base model used in this research will use the average of these two
values for the throat height, namely 400m. Above that height the hyperboloid
transitions into the flared cylinder. The radii at the bottom and the throat also
vary from design to design. For these two parameters, 140m and 75m are chosen
respectively. The wall thickness profile for the base model is also as described in
Figure 2.1, ranging from 0.6m at the bottom to 0.25m at the top. The height,

9
2. Geometry of the chimney

as stated before, will be fixed at 1000m, because of economical reasons and the
feasibility of the design (see Appendix A.1).

(a) Pre-existing design for a 1000m tall (b) Basic shape of the chimney
SUT (courtesy of Krätzig & Partner,
Bochum) (Niemann et al., 2009)

Figure 2.1: Deriving the geometric parameters of the chimney

Directional design could benefit the economical feasibility of a solar updraft tower.
An example of directional design is shown in Figure 2.2. The cross-section on the
left shows the structure being loaded in the dominant wind direction, requiring a
large amount of reinforcement on the windward side and a large amount of concrete
material on the leeward side. The cross-section in the centre depicts the same
structure being loaded in the opposite direction by a smaller wind load, requiring
less material and reinforcement. The sum of these cross-sections is optimised for
both wind directions, leading to a substantial decrease in material costs. If one
were to optimise a single solar updraft tower for a given location then directional
design would be a valid design choice, but since the goal of this research is to
optimise a generic design of the chimney which should be suitable for any given
location around the globe, the design of the chimney will be axisymmetric.

10
2.2. Hyperboloid

Figure 2.2: Example of directional design

2.2. Hyperboloid
A hyperboloid is an axisymmetric, doubly ruled surface with negative Gaussian
curvature. One way to create such a surface is to rotate a hyperbola around its
central axis. The general formula for a hyperbola is
x2 y2
− =1 (2.1)
a2 b2

By redefining the x-axis as the radius (r) of the hyperboloid and the y-axis as the
height (z) of the hyperboloid the equation can be rewritten as

r2 z2
− =1
a2 b2
r2 z2
=1+ 2
a2 r b
r z2
= 1+
a b2
r
r 1 p 2
= · b + z2
a b2
a p
r = · b2 + z 2 (2.2)
b

The hyperbola can then be offset from the central axis by ∆r:
a p
r(z) = ∆r + · b2 + z 2 (2.3)
b

In this formula, a and b are the classical parameters for defining a hyperbola while
∆r allows for the curve to be offset from its central axis. By rewriting these
equations the hyperbola can be defined for the previously stated parameters: Ht ,
rt , rbot and φb . The diameter at the throat (rt ) and the bottom (rbot ) can be
defined as
rt = r(z = 0) (2.4)
rbot = r(z = −Ht ) (2.5)

The following relationship exists between the inclination angle φ and the derivative
of the radius r(z)
dr
r0 (z) = = tan(φ) (2.6)
dz
so that
tan(φb ) = r0 (z = −Ht ) = rbot
0
(2.7)

11
2. Geometry of the chimney

Figure 2.3: Hyperbola

The offset from the central axis (∆r) can now be written as

∆r = rt − a (2.8)

a and b can then be defined as (Schindelin, 2002)


r − rt
a= 1
p bot (2.9)
b b2 + Ht2 − 1

s
0
 
Ht rbot rbot − rt
b=− 0 0 − Ht (2.10)
2(rbot − rt ) − Ht rbot rbot

With these equations, every imaginable combination of Ht , rt , rbot and φ can be


defined. Yet there are some limitations. First of all, the radius at the bottom (rbot )
always has to be larger than the radius at the throat (rt ). If rt = rbot then the
hyperboloid turns into a cylinder. The problem is that this cannot be defined using
the formula for a hyperbola anymore since the curvature will become zero. If rt >
rbot , one can no longer speak of a hyperboloid, but rather an ellipsoid which has
positive Gaussian curvature. As stated before, for the base model, rbot will be set
at 140m and rt will be set at 75m.
Another limitation is that the inclination angle has to be within a certain range.
The angle has to be large enough to accommodate the decrease in radius from
rbot to rt over the height Ht . On the other hand, if the angle is too steep, the
curvature becomes so strong that for a given rbot and Ht , rt cannot be reached.
The minimum and maximum values for the inclination angle are
   
rbot − rt rbot − rt
arctan < φb < arctan 2 (2.11)
Ht Ht
For the base model this means that the minimum angle is 9.2◦ and the maximum
angle is 18.0◦ . The default angle will be set as the average of these two values:
13.6◦ .

12
2.3. Cylinder

2.3. Cylinder
The top part of the chimney is shaped like a cylinder with a slight flare, essentially
turning it into an upside down cone. The flare should cause an increase in area of
14%. Based on this, the relationship between the radius at the top of the cylinder
and the radius at the bottom is

Atop = 1.14Ab
2
πrtop = 1.14πrb2

rtop = 1.14rb (2.12)

However, the cylinder does not go all the way from the top to the bottom of the
chimney; the bottom is shaped as a hyperboloid and only from the throat height
onwards does it become a flared cylinder. The radius at the bottom is therefore
unknown. What is known is the radius at the throat. By using an imaginary part
of the cylinder which continues over the entire height of the chimney (Figure 2.1.b),
the radius at the bottom can be calculated with
H
rb = rtop + (rt − rtop )
H − Ht
√ √ H
rb = 1.14rb + (rt − 1.14rb )
H − Ht
rt
rb = H−Ht
√ H−Ht
 (2.13)
H + 1.14 1 − H

The slope of the cylinder part of the chimney can now be defined as

0 dr rtop − rb
rcyl = = (2.14)
dz H

2.4. Matching conditions


If the hyperboloid is cut off at the throat height and then continues as a cylinder
with a constant slope, a kink will occur. Therefore, matching conditions have to be
found to allow for a smooth transition of one shape into the other. For a smooth
transition the slopes of the hyperboloid and the cylinder have to match, so that
0 0
rhyp (Hh ) = rcyl (2.15)

The derivative of the cylinder is constant as can be seen in Equation 2.14. The
derivative of the hyperboloid is
0 a z
rhyp (z) = √ (2.16)
b b + z2
2

The distance z from the throat height at which the slopes of the two shapes are
equal is (Schindelin, 2002)
a z 0
√ = rcyl
b b + z2
2
p b 0
z = b2 + z 2 rcyl
a
0
b2 rcyl
z=q (2.17)
02
a2 − b2 rcyl

13
2. Geometry of the chimney

The height of the hyperboloid can now be defined as

0
b2 rcyl
Hh = Ht + q (2.18)
02
a2 − b2 rcyl

At this height, the slope of the two shapes match, resulting in a smooth transition.

2.5. Stiffening rings

Natural draught cooling towers, the closest relative of a solar updraft tower, are
usually constructed without intermediate stiffening rings. The chimney of a solar
updraft tower, with its extreme height, would require a significantly increased wall
thickness if this were the case. For this reason, stiffening rings are introduced to
keep the wall thickness to a minimum. They serve the following purposes (Harte
et al., 2013):

• reduce shell buckling lengths to improve the chimney’s buckling resistance

• reduce the vibration wave length to improve the natural frequencies

• constrain typical shell-like behaviour under wind actions towards a more


beam-like behaviour

The base model for the chimney will have nine intermediate stiffening rings spaced
evenly over the height of the chimney, and an upper edge ring also found in most
natural draught cooling towers. A single ring has a width of 4.5m and a height of
1.5m. These dimensions are based on the chimney described in Borri et al. (2011).
Figure 2.4 shows technical drawings for the intermediary stiffening rings and the
top ring. Note that the wall thickness increases temporarily near each ring stiffener.

(a) Intermediary ring (b) Top ring

Figure 2.4: Technical drawings for the ring stiffeners (Harte and Krätzig, 2011; Niemann et al.,
2009)

14
2.6. Conclusions

2.6. Conclusions
The geometry of the chimney is based on pre-existing designs by Krätzig & Part-
ners. It is characterised by a hyperboloid shape at the bottom and a flared cylinder
at the top. The hyperboloid will end at a height of 400m, allowing the chimney to
transition into the cylinder. Matching conditions have been derived which describe
this transition. The radius at this throat height where the transition takes place
is set at 75m for the base model, the radius at the bottom of the chimney is set
at 140m. The angle of inclination with which the hyperboloid converges from the
radius at the bottom to the radius at the throat height is set at 13.6◦ ; the average
of the minimum and maximum allowable angle. The flare of the chimney is set at
14%, based on previous research.
Additionally, the chimney has a relatively thin wall thickness profile, causing it
to require stiffening rings which stabilise the structure. For the base model, nine
intermediate stiffening rings and a top ring are applied, spaced 100 meters apart
from one another. Each stiffening ring has a width of 4.5m and a height of 1.5m.

15
Actions working on the
3
chimney
In this chapter, loading conditions for the chimney will be discussed, ranging from
dead loads to the vertical wind profile and circumferential wind distribution for
quasi-static wind actions. Furthermore, a summary is given of dynamic wind ac-
tions which could affect tall, slender structures.

3.1. General remarks


The chimney of a solar updraft tower is exposed to various actions. Typical load
actions working on the chimney are (Borri et al., 2011):
• Dead loads D
• Wind loads W (consisting of external pressure and suction We , and internal
suction Wi )
• Shrinkage effects S
• Temperature effects T
• Soil settlements B
• Seismic loads E
• Construction loads M
By avoiding seismic regions, earthquake loads will not affect the chimney’s design.
Furthermore, the chimney is not very susceptible to seismic actions since the first
natural frequency of the chimney is very low (Krätzig et al., 2008). Temperature
effects on the chimney are relatively small compared to dead loads and wind loads
and can be therefore be neglected (Dyk, 2008). Unlike natural draught cooling
towers, solar updraft towers have no neighbouring buildings which could cause
differences in soil settlements. Construction loads are not included in the model as

17
3. Actions working on the chimney

there is still very little known about the construction process and therefore would
warrant a research of its own. Therefore, the load model for the chimney will be
simplified as to only include dead loads D and wind loads W .

3.2. Gravity

For the gravity loading of the structure, the self-weight of the concrete is set at
γc = 25kN/m3 . For the wind ribs, described later on in this chapter, an additional
load of ∆γ = 0.2kN/m3 is taken into account. This gives a total self-weight of γ =
25.2kN/m3 . In the sensitivity analysis in Chapter 7, several types of lightweight
concrete mixtures will be compared to regular density concrete to determine the
effect of concrete density on the dynamic behaviour.

3.3. Wind

Wind loads are often dominant in the design of structures and the importance of
wind design increases exponentially when the structure increases in height. Often-
times if structures fail it will be because not enough attention was given to this
aspect in the design phase. Wind loads are determined by the wind climate at
the site’s location, the shape of the structure, and its structural properties. The
most important parameter in wind design is the mean wind speed, which is defined
as the extreme wind speed with a certain return period, averaged over a certain
period of time, e.g. one hour.

In the atmospheric boundary layer, the velocity of wind increases with height. The
increase in the mean velocity can be described with a wind profile (Figure 3.1).
However, this wind velocity is not a constant, due to a phenomenon known as
turbulence. Wind is turbulent because of the friction with the terrain it passes
over. The wind profile can thus be described as having a mean component and a
turbulence component, both of which vary with height (Dyrbye and Hansen, 1997).

Figure 3.1: Mean wind profile with turbulence component (Bachmann et al., 1995)

18
3.3. Wind

3.3.1. Mean wind profile


The mean wind profile can be described with several types of functions. In this
research, a logarithmic function is chosen. (Note that other types of wind profiles
are compared and discussed in Appendix A.4). The standard logarithmic profile
used in many codes is, unfortunately, not valid at high altitudes, because it does
not take the Coriolis effect into account. The Coriolis effect is caused by the Earth’s
rotation. A corrected logarithmic profile, as described below, has been developed
by Harris and Deaves (Deaves and Harris, 1978) which does take this effect into
account. Note that the last three terms in the equation below only come into play
for altitudes of 300m and above.
(      2  3  4 )
u∗ z z z 4 z 1 z
Vz = ln + 5.75 − 1.88 − + (3.1)
k z0 zG zG 3 zG 4 zG

In this formula, the actual height is normalised by a gradient height zg which


incorporates the Coriolis effect. The gradient height can be calculated using the
formulae below, with f as the Coriolis parameter.

u∗
zG = (3.2)
6f
f = 2Ω sin φ (3.3)

In the logarithmic profile, u∗ is the frictional velocity. For strong winds, u∗ is


usually in the range of 1 − 2m/s and can be described with
q
u∗ = Vref 2 ·κ (3.4)

 2
k 
κ=   (3.5)
10
ln z0

where

zg gradient height
z0 roughness length
f Coriolis parameter
φ latitude
Ω angular velocity of the Earth’s rotation (72.9 · 10−6 radians/s)
u∗ frictional velocity
k Von Kármán constant (0.4)
Vref reference hourly mean wind speed

In the logarithmic profile, the terrain roughness is represented by a parameter


known as the roughness length z0 . The roughness length is the size of a charac-
teristic vortex, formed by the friction between the ground and the wind, as can be
seen in Figure 3.2. The roughness length for several terrain types can be seen in
Figure 3.3.

3.3.2. Reference hourly mean wind speed


The method for determining the reference wind speed to be used in design calcula-
tions varies all over the world. The basic wind speed in the ASCE code (ASCE7-02,
2002) is based on a 3-second gust speed at 10m above the ground in a flat open

19
3. Actions working on the chimney

Figure 3.2: Simplified representation of the roughness length (Dyrbye and Hansen, 1997)

country, with a return period of 50 years. The fundamental basic wind velocity in
Eurocode (EC) (EN1991-1-4, 2010), on the other hand, is based on a 10-minute
mean wind velocity in the same conditions. The Harris and Deaves logarithmic
wind profile is based on hourly mean wind speeds. The international ISO stan-
dard (ISO-4354, 2009) provides a simplified method for converting these different
averaging intervals, defined as
3sec 1hr
Vref = 1.53Vref (3.6)
10min 1hr
Vref = 1.05Vref
3sec 10min
Vref = 1.46Vref

The average lifespan of a solar updraft tower should exceed 100 years, otherwise
it will not be cost-effective in comparison to other types of power plants. The
reference hourly mean wind speed should also reflect this design life. Table 3.1
provides correction factors for converting the mean wind speed to a different return
period.

Return period
Correction factor
(years)
500 1.23
200 1.14
100 1.07
50 1.00
25 0.93
10 0.84
5 0.78

Table 3.1: Correction factor for other return periods (ASCE7-02, 2002)

3.3.3. Wind pressure


When the wind flow approaches a structure, it is forced over and around it, creating
areas of pressure and suction. The amount of suction generated is dependent on
the roughness of the structure (Figure 3.4).
First of all, the free stream velocity pressure perpendicular to the structure has to
be defined, using
1 2
qz = ρz Vz,p (3.7)
2

20
3.3. Wind

Figure 3.3: Roughness length z0 for several terrain types (ISO-4354, 2009)

In this formula, ρz is the density of air for a given height z and Vz,p is the peak
wind speed for that same height. The peak wind speed is the summation of the
mean wind speed and a turbulence component (gv Iv )

Vz,p = Vz · (1 + gv Iv ) (3.8)

where

gv peak factor (3.0)


1
Iv turbulence intensity z

ln z0

Once the peak wind pressure is known, the pressure distribution on the structure
can be determined. For a cylindrical shape the pressure is dependent on two
coefficients: the external wind pressure cpe,θ which varies with the angle θ, and the
internal wind pressure cpi , which is constant over the cross-section. For height z
and angle θ, the pressure acting on the surface is (Dyk, 2008)

qz,θ = [cpe,θ − cpi ] · qz (3.9)

where

cpe,θ external pressure coefficient


cpi internal pressure coefficient

21
3. Actions working on the chimney

Figure 3.4: Pressure distribution for chimney with ribs (left) and with a smooth surface (right)
(Niemann et al., 2009)

3.3.4. Pressure coefficients


The pressure coefficients used in this research are derived from the codes for ND-
CTs. The internal pressure is considered axisymmetric which means that a constant
value of 0.5 can be assumed for the internal pressure coefficient cpi which is valid
for the entire cross-section (Gould, 2005).
The definition of the external pressure coefficient is a little bit more complex. The
code specifies several curves for different roughness values of the outer surface
(Figure 3.5) (VGB-R-610E, 2010). The curves are divided in three sectors. Sector
I, which causes pressure at the front of the structure, sector II, which causes suction
at the sides of the structure, and sector III which causes suction at the back of the
structure.

Figure 3.5: Pressure distributions for the design of NDCTs (Gould, 2005)

22
3.3. Wind

The formulae for describing the sectors for the different pressure distributions are
shown in Table 3.2. The suction coefficient in sector III is constant for all of the
curves.

Curve Min Sector I Sector II III


90 2.267 90 2.395
K 1,0 -1.0 1 − 2.0(sin 71 θ) −1.0 + 0.5(sin( 21 (θ − 70))) -0.5
90 2.239 90
K 1,1 -1.1 1 − 2.1(sin 71 θ) −1.1 + 0.5(sin( 22 (θ − 71)))2.395 -0.5
90 2.206 90
K 1,2 -1.2 1 − 2.2(sin 71 θ) −1.2 + 0.5(sin( 23 (θ − 72)))2.395 -0.5
90 2.166 90
K 1,3 -1.3 1 − 2.3(sin 73 θ) −1.3 + 0.5(sin( 24 (θ − 73)))2.395 -0.5
90 2.104 90
K 1,5 -1.5 1 − 2.5(sin 75 θ) −1.5 + 0.5(sin( 27 (θ − 70)))2.395 -0.5
90 2.065 90
K 1,6 -1.6 1 − 2.6(sin 76 θ) −1.6 + 0.5(sin( 28 (θ − 76)))2.395 -0.5

Table 3.2: Equations for the pressure distribution curves H(θ) (Gould, 2005)

To determine which curve should be used, the roughness parameter of the surface
needs to be known. The roughness parameter is defined as the ratio between
the thickness of the wind ribs k and the distance between two adjacent ribs aR ,
measured at around one-third of the height of the tower. The values are shown
in Figure 3.6 and Table 3.3. As can be seen, especially the suction at the side is
significantly affected by this roughness parameter. Meanwhile, the pressure at the
front corresponds with the peak pressure which was previously calculated.

Figure 3.6: Surface roughness and type of pressure distribution (Gould, 2005)

3.3.5. Density of air


With a structural height of 1000m, the density of air for calculating wind pressure
can no longer be taken as a constant value. The density decreases as the altitude
increases, leading to a lower wind pressure for the higher regions of the chimney.
The air pressure at a given altitude h (AMSL) can be calculated using
  gM
Lh RL
p = p0 1− (3.10)
T0

23
3. Actions working on the chimney

Execution Roughness parameters k/ar Pressure min cpa Pressure distribution


With ribs 0.025-0.100 -1.0 K 1,0
0.016-0.025 -1.1 K 1,1
0.010-0.016 -1.2 K 1,2
0.006-0.010 -1.3 K 1,3
Without ribs Smooth -1.5 K 1,5
-1.6 K 1,6

Table 3.3: Surface roughness and type of pressure distribution (Gould, 2005)

The temperature at altitude h can be approximated by

T = T0 − Lh (3.11)

The density is plotted for an altitude range of 0 − 2000m in Figure 3.7. It was
calculated using the molar form of the ideal gas law

pM
ρ= (3.12)
RT

where

p0 sea level standard atmospheric pressure (101325P a)


T0 sea level standard temperature (288.15K)
g earth-surface gravitational acceleration (9.81m/s2 )
L temperature lapse rate (0.0065K/m)
R ideal gas constant (8.31447J/(mol · K))
M molar mass of dry air (0.0289644kg/mol)

2,000

1,500
Altitude h AMSL (m)

1,000

500

0
1 1.05 1.1 1.15 1.2 1.25
Density of air ρ (kg/m3 )

Figure 3.7: Density of air as a function of altitude

24
3.4. Reference location

3.4. Reference location


The chosen location for the solar updraft tower will be somewhere in the arid
belt of Africa. These countries have an exceptionally high annual solar irradiation
and very low wind speeds due to the fact that most countries are landlocked.
The Central African Standards Institution specifies a 3-second gust wind speed of
35m/s, with a return period of 50 years (Holmes, 2007). These countries have an
approximate latitude coordinate of 15◦ North, which is needed to determine the
Coriolis parameter. The altitude for the reference location is 300m AMSL.

3.5. Load case for the chimney


The loads described in Section 3.2, 3.3 and 3.4 will be summarised here to give a
comprehensive overview of all the considered actions. Firstly, the dead load caused
by self-weight will be set at γ = γc + ∆γ = 25 + 0.2 = 25.2kN/m3 . This is with
the inclusion of wind ribs so that pressure distribution curve K1,0 can be used for
wind calculations.
The variable loads only include wind loads, with the wind pressure being described
as
qz,θ = [cpe,θ − cpi ] · qz (3.13)

The external pressure coefficient cpe,θ in the formula above can be computed using
the formulae in Table 3.2 for curve K1,0

θ ≤ 71◦

 1 − 2.0 · (sin 90
71 θ)
2.267

cpe,θ = 90
−1.0 + 0.5 · (sin( 21 (θ − 70))) 2.395
71 ≤ θ ≤ 91◦

−0.5 θ ≥ 91◦

While the internal pressure coefficient cpi is set at 0.5.


The peak pressure qz is defined as

1 2
qz = ρz Vz,p (3.14)
2
The density of air ρz will not be a constant value but will vary in altitude as
depicted in Figure 3.7. The altitude for the chosen location is h = 300m. This
means that the density varies from 1.19kg/m3 at the base of the chimney (+300m
AMSL) to 1.07kg/m3 at the top of the chimney (+1300m AMSL).
The peak wind speed for given height z can be put together using the expression
for the mean wind speed with the addition of the turbulence intensity
(      2  3  4 )
u∗ z z z 4 z 1 z
Vz = ln + 5.75 − 1.88 − +
k z0 zG zG 3 zG 4 zG

Vz,p = Vz · (1 + gv Iv ) (3.15)
Terrain roughness category 2 is chosen for the solar updraft tower, with a roughness
length z0 of 0.03m. The reference latitude coordinate for the calculation of the
Coriolis parameter is 15◦ North. The 3-second gust wind speed for the chosen
location is 35m/s with a return period of 50 years. To compute the hourly mean
wind speed with a return period of 100 years, the correction factors 1/1.53 and 1.07
need to be applied. This gives a reference hourly mean wind speed Vref of 24.5m/s.
The free stream velocity pressure for these parameters is shown in Figure 3.8

25
3. Actions working on the chimney

1,000

800

Height (m)
600

400

200

0
0 500 1,000 1,500 2,000 2,500
Wind pressure (N/m2 )

Figure 3.8: Free stream velocity pressure

3.6. Dynamic wind loading


Several types of dynamic wind loading exist, a classification of which is shown
in Figure 3.9. Along-wind oscillations originate due to pressure fluctuations in
gusts or because of increased turbulence due to buffeting (vortex-shedding caused
by obstacles). Buffeting also falls in the category of aerodynamic interference
caused by surrounding structures. Across-wind oscillations are the result of either
forced vibrations caused by vortex shedding or because of aerolastic effects such as
galloping and fluttering. For a solar updraft tower, it can be assumed that there
are no large surrounding structures, because of the large diameter of the collector
area. This research will focus on the two aspects which pose the greatest threat
towards solar updraft towers, namely gust response and vortex shedding.

Figure 3.9: Classification of dynamic effects from wind (Bachmann et al., 1995)

3.6.1. Along-wind oscillations


The fluctuation of the wind speed, also known as wind gusts, can cause a structure
to resonate if an eigenfrequency of the structure matches the gust frequency. The
process for finding the spectral response of a structural system to wind gusts can
be seen in Figure 3.10. The elements shown in Figure 3.10 are described below
(Bachmann et al., 1995):

26
3.6. Dynamic wind loading

Figure 3.10: Spectral densities and magnification functions (Bachmann et al., 1995)

Gust spectrum Measured wind turbulence contains a variety of frequencies and


accompanying amplitudes. By using a Fourier Analysis, a gust spectrum
can be calculated from this measurement data. Several models for the gust
spectrum exist, one of which will be described later on in this section. Gust
spectra show a peak for a frequency of around 0.1Hz. However, even struc-
tures with eigenfrequencies of 1Hz and higher can still have quite a significant
dynamic response; most notably structures with very little damping.
Aerodynamic admittance function This function portrays the relationship be-
tween the gust frequency and the area of influence. Gusts of a low frequency
have a large area of influence and vice versa. This function can also be seen
as a surface correction factor.
Spectral density of wind If the square of the aerodynamic admittance function
is multiplied with the gust spectrum, the spectral density of wind is found
for a particular structure. In other words, the dynamic force acting on a
structure.

27
3. Actions working on the chimney

Mechanical amplification function This represents the response of a structure


to a harmonic excitation, expressed as a dynamic amplification factor for the
static response. If multiple eigenfrequencies are considered, the mechanical
amplification function will also show several peaks.
Spectral density of system response If the spectral density of wind is multi-
plied with the square of the mechanical amplification function, the spectral
density of the system response is found. This spectrum shows the frequen-
cies for which a structure is most susceptible to resonate as a result of wind
turbulence.
The equation for the spectral density of the system response is (Ruscheweyh, 1982)

f · Sy (f ) f · Sz (f ) 2 2
=4 2 χ χ (3.16)
Iz2 ȳ 2 Iz ū(z)2 a m

where
f ·Sy (f )
Iz2 ȳ 2 spectral density of system response
f ·Sz (f )
Iz2 ū(z)2 gust spectrum
χa aerodynamic admittance function
χm mechanical amplification function

Gust spectrum
Various researchers have developed spectra which describe longitudinal fluctuations
in the mean wind speed. Kolmogorov law states that the spectra have to approach
5
an asymptotic limit which is proportional to f − 3 when it reaches high frequencies.
The reason for this decay of turbulence is due to ever-smaller vortices occurring at
higher frequencies (Burton et al., 2011). These spectra are usually expressed as a
reduced spectrum FD (Vrouwenvelder, 2014)

f · Sz (f )
FD = (3.17)
Iz2 ū(z)2

One of the most commonly used spectra is the Von Kármán spectrum, which is
also the preferred spectrum in the international ISO standard. The Von Kármán
PSD can be defined as
 
f Lv
f · Sz (f ) 4 Vm
= (3.18)
Iz2 ū(z)2 
f Lv
2 5/6
1 + 70.8 Vm

where

f frequency
Sz longitudinal spectrum of turbulence in terms of wind speed
Lv longitudinal integral length scale (280m)
Vm mean wind speed
σv standard deviation of wind speed

28
3.6. Dynamic wind loading

0.3

0.2
f ·Sz (f )
Iz2 ū(z)2
0.1

0
10−3 10−2 10−1 100 101 102
Frequency (Hz)

Figure 3.11: Von Kármán PSD

Aerodynamic admittance function

The aerodynamic admittance function describes the inverse relationship between


the gust frequency and the area of influence, it can be defined as (Ruscheweyh,
1982)
1
χa =  √ 4/3 (3.19)
1 + 2fū(z)A

where

f frequency
A vertical cross-sectional area of the chimney
ū(z) wind speed at 2 /3 of the structural height

For a 1000m tall chimney with a diameter of 150m (neglecting the widening of
the base) the vertical cross-sectional area of the chimney is 1.5 · 105 m2 . The wind
speed at two-thirds of the height is estimated at 60m/s. A plot of the aerodynamic
admittance function used in this research is shown in Figure 3.12.

0.8

0.6
χa
0.4

0.2

0
10−3 10−2 10−1 100 101 102
Frequency (Hz)

Figure 3.12: Aerodynamic admittance function

29
3. Actions working on the chimney

Spectral density of wind


By multiplying the square of the aerodynamic admittance function with the gust
spectrum, the spectral density of wind for this particular structure is found (Fig-
ure 3.13). As can be seen, due to the enormous size of the structure, the pressure
fluctuations for high frequencies are evened out. As a result, only the first few
eigenfrequencies of the chimney will be critical in terms of along-wind oscillations.
The odds that eigenfrequencies above 0.3Hz will cause resonance in the chimney
are very slim.
For a solar updraft tower of 1000m, this means that the first two eigenfrequen-
cies will be governing for the design. The first eigenfrequency is usually in the
range of 0.17Hz and is associated with a beam bending mode, while the second
eigenfrequency often occurs around 0.2Hz and is typified by a shell bending mode.

0.3

0.2
f ·Sz (f ) 2
Iz2 ū(z)2 χa
0.1

0
10−3 10−2 10−1 100 101 102
Frequency (Hz)

Figure 3.13: Wind force spectrum

Spectral density of system response


The area under the curve of the spectral density of system response is the variance
of the dynamic response and can be calculated using (Schindelin, 2002)
Z ∞
σy2 = Sy (f ) df (3.20)
0
Z ∞ 2 2
Iv ȳ f Sy (f )
σy2 = df
0 f Iv2 ȳ 2
Z ∞
σy2 2 1 f Sy (f )
2
= Iv df
ȳ 0 f Iv2 ȳ 2

Figure 3.14: Composite trapezoidal rule (en.wikipedia.org)

The variance can be computed using the composite trapezoidal rule (Figure 3.14).
The square root of the found variance is the standard deviation σȳv . The dynamic

30
3.6. Dynamic wind loading

response for the chimney can then be found by combining the static response and
the standard deviation for the dynamic wind response, using

ydyn = ȳ + nσv (3.21)


σv
ydyn = ȳ + n ȳ

σv
ydyn = ȳ(1 + n )

where n is dependent on the exceedance probability P . Using the normal distri-


bution, if P is set at 5% then the Z-score and thus n is 1.645. Now that all of the
required parameters are known, the equation for the gust factor G can be described
as
σv
G=1+n (3.22)

While the wind model previously described already accounts for turbulence inten-
sity, now individual chimney structures can be compared based on their dynamic
response by comparing gust factors.

3.6.2. Across-wind oscillations


Chimney structures, characterised by their cylindrical and relatively slender shape,
cause vortices to be shed along the direction of the wind, alternating between the
left and right side of the cross-section (Figure 3.15). The result of this is that the
chimney will be subjected to pulsating excitation forces in the across-wind direction.
If the frequency of the vortex shedding is the same as one of the eigenfrequencies
of the structure then the structure will start to resonate (Bachmann et al., 1995).

Figure 3.15: Vortex shedding in transcritical regime (Sawka, 2004)

This happens when


fe · d
vcrit,ovaling = (3.23)
St · N
where

fe eigenfrequency [Hz]
d diameter of the chimney [m]
St Strouhal number (0.2 for cylindrical cross-sections)
N number of waves around the cross-section

Vortex shedding occurs for a structure not in motion, so it falls in the category of
forced vibration. However, once the structure starts to move due to these forced
vibrations, the vortex shedding can synchronise with one of the eigenfrequencies of
the structure in a certain range below and above the critical wind speed, leading to
a so-called lock-in effect (Figure 3.16). For a structure which is subjected to a pure

31
3. Actions working on the chimney

forced vibration, the traditional mechanical amplification would suffice, but for a
critical vortex excitation where a lock-in effect occurs, the mechanical amplification
function is greatly broadened.

Figure 3.16: Across-wind vibration amplitude for reduced wind speed ur = u/dfe (Bachmann
et al., 1995)

The critical wind speed vcrit,ovaling is dependent on the number of waves around
the cross-section (Figure 3.17). Because of this, higher eigenfrequencies which have
more waves around the cross-section could lead to critical wind speeds lower than
those calculated for lower eigenfrequencies. Shell bending modes have a minimum
of two waves around the cross-section while a beam bending mode only has one
wave around the cross-section. Consider two eigenfrequencies which are similar in
value; one has a beam bending mode shape and the other a shell bending mode
shape with two waves. The critical wind speed for the shell bending mode shape
will be twice as low as the critical wind speed for the beam bending mode shape.
This aspect has to be considered when tuning the eigenfrequencies of the chimney.

The Eurocode (EN1991-1-4, 2010) specifies that the effect of vortex shedding does
not need to be investigated if

vcrit,ovaling > 1.25 · vm (3.24)

where vm is the characteristic mean wind velocity (averaged over 10 minutes) at


the cross section where vortex shedding occurs.

Figure 3.17: Ovalising shapes for cylindrical shell structures (Ruscheweyh, 1982)

32
3.7. Conclusions

3.7. Conclusions
The load case for the chimney consists of two types of loading, namely the dead
load and the wind load. The dead load is based on the self-weight of concrete and
the additional inclusion of wind ribs. The wind load can be calculated as follows:
first of all, a reference hourly mean wind speed is defined, based on the location of
where the chimney will be built and the working design life of the chimney. This
reference hourly mean wind speed can then be used to calculate the wind speeds
over the entire height of the chimney, using the corrected logarithmic profile, as
developed by Harris and Deaves. These wind speeds can then be converted to
peak wind pressures, which, using the circumferential pressure distribution of the
guidelines for NDCTs, can subsequently be converted to pressure loads which act
on the chimney.
Additionally, two types of resonance can be distinguished in these structures; along-
wind resonance, and across-wind resonance. Along-wind resonance is caused by
turbulence in along-wind gusts. The second type, across-wind resonance, is caused
by the alternating shedding of vortices. This leads to pulsating excitation forces in
the across-wind direction, and, if the frequency of the vortex shedding is the same
as one of the eigenfrequencies of the chimney, resonance will occur.

33
Finite element model
4
In this chapter, the modelling procedure and validation of the finite element (FE)
model are described. Firstly, the base model will be defined and tested for which
mesh density the solution converges. Subsequently, a modified base model, similar
to the chimney which is described in other studies, will be verified by comparing
the results to those found in the aforementioned literature. Analysis procedures for
quasi nonlinear material behaviour and buckling wil be described. Furthermore,
the influence of soil conditions on the base fixity is researched. Lastly, a constant
damping ratio for the chimney will be specified, based on research described in
Appendix A.5.

4.1. Discretisation
The first step in creating a finite element model is the meshing of the geometry.
Meshing entails that the geometry is subdivided in numerous small elements which
contain several integration points. Stresses and strains can be computed in these
integration points after loading, which leads to the general solution (Bathe, 1982).
The subdivision of the geometry in these mesh elements is a key aspect of the finite
element modelling process. Choose too few elements and the solution will deviate
too much from the actual analytical solution, choose too many elements and the
computation time will increase drastically. The steps required for building a FE
model are displayed in Figure 4.1.
A convergence study for the mesh density will determine a suitable amount of
elements for the chimney. Convergence is reached when an increase in mesh density
has little to no influence on the outcome.

4.1.1. Software choice


The modelling and analysis will take place in ANSYS Mechanical APDL, a software
suite for finite element analysis. The reason for this choice is twofold. First of all

35
4. Finite element model

Figure 4.1: The eight steps of creating a FE model: defining keypoints, drawing splines, creating
areas, setting mesh divisions, meshing, setting boundary conditions, adding stiffening rings,
applying loads

due license availability, secondly due to the enormous versatility of the software
package; nothing is predefined which gives the user more freedom than most FE
packages suited for structural analysis.

4.1.2. Element choice


A great many different element types are available in ANSYS. To model a continu-
ous concrete shell structure, two types of elements can be used, solid elements and
shell elements. Solid elements are the most common in finite element modelling,
but they are rarely used in the analysis of shell structures. The reason for this is
that the aspect ratio of solid elements should be close to a cubic shape, requiring
a tremendous amount of elements in relatively thin shell structures, thus greatly
increasing the computation time. Shell elements only have nodes in a 2D space
and an extra parameter which defines its thickness. This makes them suitable for
structures which have a relatively small thickness compared to the dimensions in
the other directions, such as the chimney of a solar updraft tower.
The next choice to be made is whether to use linear or quadratic elements. Quadratic
elements have midside nodes which make them more suitable for analysis problems
which deal with bending. Linear elements have fewer nodes which in general de-
creases the computation time. However, a great amount of linear elements are
required to properly deal with bending. The chimney of a solar updraft tower is

36
4.1. Discretisation

mainly loaded in bending, so the choice is made to use quadratic elements. The
most commonly used quadratic shell element in ANSYS is the SHELL281 element
(Figure 4.2).

Figure 4.2: SHELL281 element (ANSYS, 2013)

For the modelling of the stiffening rings, beam elements are used. The choice is
made to use quadratic beam elements due to compatibility with the quadratic shell
elements, allowing them to share corner and midside nodes (Figure 4.3). ANSYS
offers quadratic beam elements in the form of the BEAM189 element.

To summise, the following two element types will be used in the modelling process:

SHELL281 is suitable for analysing thin to moderately-thick shell structures. It


has 8 nodes with six degrees of freedom per node.

BEAM189 is suitable for analysing slender to moderately-thick beam structures.


It is a quadratic beam element with 3 nodes, having six degrees of freedom
per node.

Figure 4.3: SHELL281 elements share their nodes with BEAM189 elements

37
4. Finite element model

4.1.3. Element sizing

Special attention has to be given, not to the actual dimensions of each element,
but to the aspect ratio of it. While more elements might be required in the vertical
direction than in the circumferential direction, the ratio between the width and the
height of an element cannot become too large, otherwise a phenomenon known as
shear locking might occur.

Shear locking occurs because linear, and sometimes even quadratic elements cannot
accurately portray the curvature caused by bending. The result of this is that
a shear stress is introduced in the element. This additional shear stress, which
in reality does not exist, causes the element to reach equilibrium with smaller
displacements, making it stiffer than it actually is. A good aspect ratio to strive for
to avoid shear locking is 1:2. Especially for this type of chimney, shear locking can
greatly influence the accuracy of the results, because the thinner a shell structure
is, the larger the influence of shear locking will become (Chapelle and Bathe, 2003).

4.1.4. Mesh refinement

Some areas of the chimney will require more elements than others. Areas where
large rotations occur for instance. Since it would be unnecessary to increase the
global mesh density of the chimney just to accommodate for these few areas where
large rotations occur, it would be beneficial to vary the mesh density over the
height of the chimney. Places where large rotations are most likely to occur are
near the placement of stiffening elements. The elements in the row just above and
below each stiffening ring are subdivided an x number of times over their height
(Figure 4.4). Careful attention has to be given to the aspect ratio of the resulting
elements to avoid shear locking.

Figure 4.4: Mesh without refinement (left) and with refinement (right)

38
4.2. Material properties

4.1.5. Symmetry

The mesh density can be decreased further without sacrificing accuracy. By only
modelling half of the chimney and applying symmetry boundary conditions where
the section is made, the number of elements needed can be cut in half (Figure 4.5).
This simplification is valid because the wind loading is symmetrical as well. The
only difference is that when the chimney is modelled in full, there will be two
eigenmodes for each eigenfrequency, oriented perpendicular to each other. If only
the half of the chimney is modelled, there will be one eigenmode for each eigen-
frequency. Furthermore, torsional eigenmodes do not exist when only modelling
half of the chimney, but these are not governing, since early studies have shown
that the first torsional eigenfrequency for the base model occurs at 0.7374Hz, well
outside of the range of the wind power spectrum. On top of that, the chimney will
be loaded symmetrically by wind, thus making it impossible for torsional modes to
be activated.

Figure 4.5: Full chimney (left) and half chimney with symmetry boundary conditions (right)

4.2. Material properties

Once the geometry of the chimney is meshed, material properties have to be as-
signed to these newly created mesh elements. As stated before, the chimney of
a solar updraft tower will be constructed using reinforced concrete. To model
reinforced concrete, the characteristic values of the used concrete aggregate and
reinforcement steel need to be known.

39
4. Finite element model

4.2.1. Concrete
The chosen concrete quality for the base model is C50/60 (Niemann et al., 2009).
The design value for the compressive strength can be determined with
αcc fck fck
fcd = = (4.1)
γc 1.5

The design value of the tensile strength can be found with


αct fctk,0,05 fctk,0,05
fctd = = (4.2)
γc 1.5

where

fctk,0,05 = 0.7fctm (4.3)


(2/3)
fctm = 0.30fck for concrete classes ≤ C50/60 (4.4)

The Young’s modulus for a specific concrete strength class is


 0.3
fcm
Ecm = 22000 · (4.5)
10

where
fcm = fck + 8N/mm2 (4.6)

Figure 4.6: Bi-linear stress-strain diagram for concrete (EN1992-1-1, 2011)

Figure 4.6 shows the bi-linear relationship between stresses and strains in concrete.
For concrete quality C50/60 the design value fcd = 33.3N/mm2 , the design value
of the tensile strength is 1.90N/mm2 , and the Young’s modulus is 37000N/mm2
(Braam and Lagendijk, 2010). The Poisson’s ratio for concrete is ν = 0.2 and its
density ρ is 2500kg/m3 .

4.2.2. Reinforcement steel


If the concrete is loaded by tension forces larger than it can withstand, small cracks
will form, causing the reinforcement steel to be activated. Reinforcement steel is
different from regular structural steel in that it has somewhat different strength
and stiffness properties. The steel quality used in this design is B500A.

40
4.3. Convergence

Figure 4.7: Idealised (A) and design (B) stress-strain diagrams for reinforcing steel (EN1992-1-1,
2011)

The design value for the reinforcement steel can be calculated using (Figure 4.7)
fyk
fyd = (4.7)
γs

With a material factor γs = 1.15 the design value becomes: 435N/mm2 . The
Poisson’s ratio for steel is ν = 0.3. Its Young’s modulus is 200000N/mm2 . Lastly,
its density is 7850kg/m3 .
The aim of the chimney’s design is to use as little reinforcement steel as possible,
in order to be cost-effective. VGB guidelines for natural draft cooling towers state
that the minimum reinforcement percentage is 0.3% in the meridional direction
and 0.3% and 0.4% in the circumferential direction, in the bottom and top part
respectively (Niemann et al., 2009; VGB-R-610E, 2010).

4.2.3. Stiffening rings


For the base model, the concrete used in the stiffening rings is the same as that
which is used in the overall chimney shell. The optimisation study might lead to
an increase in either the cross-sectional area of the stiffening rings, its stiffness
properties, or perhaps the inclusion of some form of pretensioning.

4.3. Convergence
The more elements the finite element model consists of, the longer the computation
time will be. Therefore, efforts have to be made in reducing the number of elements
without sacrificing the accuracy of the solution. To find a mesh density which
meets both the demands of accuracy and fast computation time, a short study
comparing different mesh densities has been carried out. In this study, a very
dense verification model is built, consisting of 30000 elements. Several models with
different mesh densities are then compared to this verification model. The chosen
mesh densities are tested are as follows: very coarse (240 elements), coarse (720
elements), intermediate (1800 elements), fine (3840 elements), and very fine (6000
elements) (Figure 4.8). For all of these models, the first 25 eigenfrequencies are
extracted and compared to those of the verification model (Figure 4.9).
It can be clearly seen that there is no discernible difference for the first seven
eigenfrequencies for all of the tested mesh densities. However, from this point on,

41
4. Finite element model

Figure 4.8: Different degrees of mesh densities, ranging from very coarse (left) to very fine (right)

1
Frequency (Hz)

Very coarse
Coarse
0.5
Intermediate
Fine
Very fine
Verification
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
Mode
Figure 4.9: Comparing the first 25 eigenfrequencies

the very coarse mesh becomes less and less accurate, eventually becoming unable
to depict the complex shapes associated with the higher modes. The coarse mesh
also eventually deviates from the verification model around mode 17. The same
goes for the intermediate mesh around mode 19. Only the fine and very fine mesh
are able to closely follow the verification model.
As was mentioned before, extra elements can be added near the stiffening rings to
improve the accuracy of the model without increasing the overall mesh density. In
all of the five cases the row of elements above and below the stiffening rings have
been subdivided five times. These new five cases have been dubbed the ’+’-variants
(Figure 4.10). Once again, the first 25 eigenfrequencies are extracted and compared
to those of the verification model (Figure 4.11).

Figure 4.10: Different degrees of mesh densities, ranging from very coarse (left) to very fine
(right)

The refined versions of the very coarse and the coarse mesh (very coarse+ and
coarse+) do not fare well either for the higher modes. The intermediate+ mesh,
however, now performs almost as well as the fine and very fine mesh, being able to

42
4.3. Convergence

closely mimic the results of the verification model over the entire spectrum. The
error margins for the original intermediate mesh and the intermediate+ mesh are
given in Table 4.1. Based on these results, the intermediate+ model is chosen as
the basis for all future mesh divisions in this research.

1
Frequency (Hz)

Very coarse+
Coarse+
0.5
Intermediate+
Fine+
Very fine+
Verification
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
Mode
Figure 4.11: Comparing the first 25 eigenfrequencies for the ’+’-variants

Frequency
Mode Validation Intermediate Error % Intermediate+ Error %
1 0.161 0.161 0.0% 0.161 0.0%
2 0.1941 0.1941 0.0% 0.1941 0.0%
3 0.3353 0.3353 0.0% 0.3353 0.0%
4 0.4851 0.4852 0.0% 0.4852 0.0%
5 0.5017 0.5017 0.0% 0.5017 0.0%
6 0.5571 0.5572 0.0% 0.5572 0.0%
7 0.5873 0.5873 0.0% 0.5873 0.0%
8 0,6661 0.6665 0.1% 0.6664 0.0%
9 0.6784 0.6785 0.0% 0.6785 0.0%
10 0.6862 0.6863 0.0% 0.6862 0.0%
11 0.7795 0.7809 0.2% 0.7807 0.2%
12 0.7934 0.794 0.1% 0.7939 0.1%
13 0.8507 0.8609 1.2% 0.859 1.0%
14 0.856 0.8657 1.1% 0.8595 0.4%
15 0.8654 0.8661 0.1% 0.8657 0.0%
16 0.8694 0.8703 0.1% 0.8702 0.1%
17 0.8697 0.8882 2.1% 0.8833 1.6%
18 0.8739 0.8949 2.4% 0.8882 1.6%
19 0.8742 0.9083 3.9% 0.8886 1.6%
20 0.8876 0.928 4.6% 0.8978 1.1%
21 0.923 0.9541 3.4% 0.9429 2.2%
22 0.9275 0.9613 3.6% 0.9477 2.2%
23 0.935 0.9625 2.9% 0.9536 2.0%
24 0.9396 0.9713 3.4% 0.9578 1.9%
25 0.9516 0.9764 2.6% 0.9624 1.1%

Table 4.1: Error margins for the intermediate and intermediate+ mesh

43
4. Finite element model

4.4. Verification
Now that a suitable mesh density is found for all future finite element models, the
eigenfrequencies for a modified base model can be determined and compared with
values found in literature. In Harte and Krätzig (2011) a chimney comparable
to the base model is described. While the text itself is rather vague about the
parameters used in the modelling process, educated guesses come a long way in
trying to copy the chimney’s geometry and material parameters, allowing for a
comparison to be made.
Its first six natural frequencies are: 0.17Hz, 0.21Hz, 0.30Hz, 0.46Hz, 0.56Hz and
0.63Hz. As can be seen in Figure 4.12 and Table 4.2, the two chimneys do not
behave exactly alike. Yet the similarity in values and the fact that they have
identical mode shapes suggest that the modelling process used in this research is
correct.

0.8
Frequency (Hz)

0.6

0.4

0.2 Model
Literature
0
1 2 3 4 5 6
Mode
Figure 4.12: Comparing the calculated eigenfrequencies with those found in literature

Frequency
Mode Literature Model Error %
1 0.17 0.1715 0.9%
2 0.21 0.2068 -1.5%
3 0.30 0.3537 17.9%
4 0.46 0.5170 12.4%
5 0.56 0.5345 -4.6%
6 0.63 0.5936 -5.8%

Table 4.2: Error margins for calculated model compared to values found in literature

4.5. Linear vs nonlinear analysis


Reinforced concrete is a composite material (Figure 4.13) in which the low ten-
sile strength and ductility of concrete is compensated for by the inclusion of steel
reinforcement. To accurately model such a material, effects such as tension crack-
ing, tension stiffening, the nonlinear stress-strain relationship in compression, the
yielding of reinforcement and the nonlinear concrete-steel bond have to be taken
into account (Figure 4.14).
The goal of this thesis is to optimise the structure of a chimney, which requires
analysing many different chimney configurations and comparing their results. The
problem is that nonlinear analyses are very computation-heavy tasks which take a

44
4.5. Linear vs nonlinear analysis

Figure 4.13: Layered model for a reinforced concrete shell (Gould, 2005)

Figure 4.14: Nonlinear material properties of reinforced concrete (Gould, 2005)

long time to set-up and calibrate, therefore making them unsuited for this type of
research. Linear elastic analysis, on the other hand, is a much faster type of analysis
which is relatively easy to set-up in comparison. It is based on the classical bending
theory of thin shells, assuming linear elastic material behaviour and the initial
geometry, and it is based on linear kinematic law (Gould, 2005). Although it is
not completely accurate in portraying the complex material behaviour of reinforced
concrete, it does provide insight in the static and dynamic response of the chimney,
and the effects certain parameter changes bring about.

Furthermore, early studies of the base model show that mainly the stiffening rings
are subjected to large tension forces. The concrete shell itself is loaded mostly in
compression, aside from a very small region on the windward side. If the chimney
is loaded mainly in compression, and it is assumed that the stiffening rings do not
crack due to pretensioning, then the behaviour of the chimney can be approximated
with linear elastic behaviour.

4.5.1. Quasi nonlinear analysis


Although this research focuses mostly on linear elastic analyses, some form of
nonlinear analysis is used to determine what would happen if the stiffening rings
were to crack. Nonlinear material behaviour is approximated as follows: first a
linear elastic analysis is performed, then the elements which are found to be under
tension have their Young’s modulus of those elements is reduced, and subsequently

45
4. Finite element model

the analysis is run again to study the effect of a locally reduced Young’s modulus.
This process can be repeated for multiple iterations, however, as can be seen in
Figure 4.15, the solution is already completely converged after a single iteration.

1,000
Linear elastic
Quasi nonlinear (1 iteration)
800 Quasi nonlinear (2 iterations)
Quasi nonlinear (3 iterations)

Height [m] 600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 4.15: Linear elastic analysis compared to quasi nonlinear analysis (1, 2 and 3 iterations)

4.5.2. Fictional Young’s modulus


In order to perform a quasi nonlinear analysis, the Young’s modulus of the elements
under tension has to be reduced. If one were to base the Young’s modulus of the
cracked concrete purely on the Young’s modulus of the reinforcement steel then the
effect of tension stiffening would not be included. Tension stiffening occurs when
concrete starts to crack and the residual concrete zones between the cracks increase
the stiffness of the element. The method used to reduce the Young’s modulus in this
research is the fictional Young’s modulus (Ef ) method from the Dutch National
Annex of Eurocode 2 (NEN-EN1992-1-1, 2011). This method defines the following
formulae for calculating Ef for a cross-section subjected to a normal force and a
bending moment (concrete quality C50/60)

[2.65 + 455ρ + (31.5 − 265ρ)αn ] · 103 ≥ 6050



if αn ≤ 0.5
Ef =
(27.6 + 484ρ)(1 − 32 αn ) · 103 if 0.5 < αn ≤ 0.9

where

ρ (Ast + Asc )/Ac


αn NED /(fcd A + (Ast + Asc )fyd )
Ast Cross-sectional reinforcement area on the tension side
Asc Cross-sectional reinforcement area on the compression side
Ac Area of the concrete cross-section
NED Normal force in the cross-section

Figure 4.16 shows the fictional Young’s modulus for various values of αn for rein-
forcement percentages of 1% and 2%. If the shell is dominated mostly by normal
forces, which is the case in the hyperboloid part of the chimney where tension
forces are the largest, and the wall thickness of the shell is optimised, then alphan
will approach 1. The fictional Young’s modulus is in this case 10813N/mm2 and
12427N/mm2 , for 1% and 2% reinforcement respectively. The average of these two
values suggest a reduction in the Young’s modulus of 1 − 11620

37000 · 100% = 68.6% ≈

46
4.6. Buckling

70%. Therefore, for all quasi nonlinear analyses, a stiffness reduction of 70% will
be applied to the tension zones.

30,000

Ef (N/mm2 )
20,000

10,000
1%
2%
0
0 0.2 0.4 0.6 0.8 1
αn

Figure 4.16: Fictional Young’s modulus for different reinforcement percentages

4.6. Buckling

While this research focuses mostly on the dynamic response of the chimney, buck-
ling is also an important design criterion to check when optimising the structure.
The design guidelines for NDCTs state that a buckling safety factor of 5 is deemed
sufficient for the stability of a chimney. Since there are no design guidelines for
SUTs as of now, the guidelines for NDCTs will be used for this research.

ANSYS supports two types of buckling, eigenvalue buckling and nonlinear buckling.
Nonlinear buckling is the most accurate of these two since it gradually increases the
load during a static analysis until a small increase in load will lead to a large deflec-
tion in the structure. Unfortunately, nonlinear analyses are very time consuming
to set up and compute and therefore do not fit into the scope of this research.

Eigenvalue buckling is also known as classic Euler buckling, the results of which
are very unconversative, making it unsuited for actual design calculations; in real-
life, imperfections and nonlinearities in the structure prevent it from reaching the
eigenvalue predicted buckling strength. The method is, however, useful to find out
if one chimney configuration performs better than another in terms of buckling
stability.

The eigenvalues calculated by ANSYS are the buckling safety factors for the struc-
ture for the given loading situation. The eigenvalues represent scaling factors for
the applied loads. The problem is that all loads are scaled up in the eigenvalue
buckling analysis. Normally only the variable loads should be scaled up while the
dead loads should remain constant. This problem can be solved by using an itera-
tive approach (Figure 4.17). By increasing the partial safety factor for the variable
loads until an eigenvalue of 1 is found, the buckling safety factor can be derived
from the partial safety factor which was found iteratively and the original partial
safety factor used in the static analysis. For example, if the partial safety factor
used for wind is set at 1.5, and an eigenvalue of 1 is found by increasing the partial
safety factor up to 15, then the buckling safety factor is 10.

Note that for the buckling analysis, a higher mesh density needs to be used due
to the complex geometrical shapes of most buckling modes. For this research the
fine+ mesh density is used for buckling analyses.

47
4. Finite element model

Figure 4.17: Adjusting variable loads to find an eigenvalue of 1.0 (ANSYS, 2013)

4.7. Foundation
The foundation of the chimney will most likely be constructed as a closed ring with a
width of approximately 20m, depending on the conditions of the soil (Harte, 2013).
To determine the effect of soil properties on the base fixity of the chimney, a minor
study was conducted. The most basic method of modelling the soil is by using
springs which represent the vertical, horizontal, rocking and torsional stiffness of
the soil. The foundation is modelled as a circular disc, of which the total vertical
spring stiffness is (Dowrick, 2007)

4GR
kv = (4.8)
1−ν

in which R is the radius at the bottom of the chimney. G is the shear modulus of
the soil and can be derived from the Young’s modulus E and the Poisson’s ratio ν
E
G= (4.9)
2(1 + ν)

When the chimney is modelled as a full cylinder, it has 32 elements in the circum-
ferential direction. Since quadratic elements also have midside nodes, the chimney
will have 64 nodes at the base level. Due to the nature of quadratic elements,
the stiffness of midside nodes should be higher than the stiffness of corner nodes
(Figure 4.18). Each spring connected to a corner node will therefore have 13 · 32
1
of
2 1
the total stiffness and each spring connected to a midside node will have 3 · 32 of
the total stiffness, because of the 2-D nature of shell elements.
To restrain the chimney, six spring elements are introduced per node. Three springs
which restrain translation in the x-, y- and z-direction and three springs which
restrain rotation in the x-, y- and z-direction (Figure 4.19). The formulae for the
stiffness for these six degrees of freedom are displayed in Table 4.3.

Motion Spring stiffness k


4GR
Vertical 1−ν
8GR
Horizontal 2−ν
8GR3
Rocking 3(1−ν)
16GR3
Torsion 3

Table 4.3: Spring stiffnesses of the foundation (Dowrick, 2007)

Several different soil stiffnesses were tested on the base model to compare the results
of the static and modal analysis. The chosen stiffnesses are: 10MPa, 100MPa,
1000MPa (1GPa), and 10000MPa (10GPa). The results of the static analysis can
be seen in Figure 4.20.

48
4.7. Foundation

Figure 4.18: Equivalent nodal allocations for a) 2-D elements, b) 3-D elements, c) triangular 3-D
elements (ANSYS, 2013)

Figure 4.19: FE model with the foundation springs

The chimney where the soil stiffness is 10GPa behaves almost exactly like the fixed
chimney; their results overlap almost consistently. The case where the soil stiffness
is 1000MPa shows some deviations compared to the fixed chimney, especially close
to the foundation level. The other two stiffnesses which were tested (100MPa and
10MPa) can no longer fully restrain the chimney, allowing for large tension and
compression stresses to develop in the shell.
The modal analysis shows similar results: the case where the soil stiffness is 10GPa
behaves similarly to the fixed chimney, with eigenfrequencies having an error mar-
gin of under 0.3% (Figure 4.21 and Table 4.4). The other cases show similar
deviations as in the static analysis.
In Table 4.5, the Young’s modulus for various rock types is shown. All of these
types have a stiffness higher than 10GPa, so if the chimney were to be founded on
a soil composed of one of these types of rock substrate, then the chimney can be
considered fully fixed. In conclusion, it appears that if the soil is stiff enough, the
springs which represent the soil properties have a negligible effect on the results of

49
4. Finite element model

1,000
Fixed
10000MPa
800 1000MPa
100MPa
10Mpa

Height [m]
600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 4.20: Comparing the in-plane meridional forces for several soil stiffnesses with the fixed
foundation

0.6
Frequency (Hz)

0.4

Fixed
0.2 10000MPa
1000MPa
100MPa
10Mpa
0
1 2 3 4 5 6
Mode
Figure 4.21: Comparing the eigenfrequencies for several soil stiffnesses with the fixed foundation

the static as well as the modal analysis. Therefore, the base of the chimney is fully
restrained against translation and rotation.

4.8. Damping

For the base model, a constant critical damping ratio of 1.5% is used. In the
sensitivity analysis, other damping values will be used in the harmonic analysis
and compared against the results of the base model. The reasoning behind the
chosen damping method and the damping ratio which will be used in combination
with that method can be found in Appendix A.5.

50
4.9. Conclusions

Frequency
Mode Fixed 10000MPa 1000Mpa 100Mpa 10Mpa
1 0.1647 0.1643 0.1609 0.1361 0.0681
2 0.1920 0.1920 0.1920 0.1920 0.1545
3 0.3390 0.3377 0.3265 0.2473 0.1921
4 0.4756 0.4734 0.4498 0.3638 0.3247
5 0.4864 0.4851 0.4825 0.4418 0.3512
6 0.5416 0.5407 0.5229 0.4646 0.4198

Table 4.4: Comparing the eigenfrequencies for several soil stiffnesses with the fixed foundation

Rock type Stiffness [GPa]


Sandstone 10.6
Shale 12.6
Limestone 16.6
Mudstone 19.2

Table 4.5: Young’s modulus of various rock types (Pariseau, 2011)

4.9. Conclusions
The geometry which was defined in Chapter 2 can be translated into a finite element
model by a combination of shell and beam elements. Quadratic elements have to
be used to accurately portray the curvature induced by some of the more the
complex mode shapes. Calculation time can be decreased, however, by decreasing
the mesh density. By only modelling half of the chimney and by using symmetry
boundary conditions, the mesh density can be cut in half. Furthermore, by using
more elements in regions where curvature is high (near the stiffening rings), and
using fewer elements in the other regions, the mesh density can be decreased even
further without sacrificing accuracy.
The finite element model is validated by comparing results of the modal analysis
with previous studies. Considering the fact that not all of the parameters of the
chimneys described in literature are known, the results found are similar enough
to those from earlier research. Next to the eigenfrequencies, the base fixity of the
chimney is also investigated to validate that the foundation of the chimney can be
modelled as completely rigid. Lastly, a constant damping ratio is defined which is
to be used in the harmonic analyses.

51
Analysing the base model
5
In this chapter, all of the outcomes regarding the definition of the base model
will be summarised. The static and dynamic response of the base model will be
analysed to define key problem areas which could benefit from improvement.

5.1. Introduction
Before optimisation can take place, the weaknesses of the chimney have to be de-
fined. To find these weaknesses, various types of analyses are conducted. The goal
of this research is to optimise the structure under dynamic wind action. Dynamic
wind, however, also has a static component. Therefore, conducting only a dynamic
analysis would not be sufficient in uncovering all of the key problem areas of the
chimney. Furthermore, any optimisation that might take place on the dynamic
front cannot negatively influence the static or buckling response of the chimney,
nor should it increase the costs. The static analysis consists of the following com-
ponents:
In-plane forces The linear static analysis shows where tension forces occur in the
chimney. The more the circumferential distribution of forces approaches the
cosine-optimum, the stiffer the structure can be regarded as, which is also
beneficial in terms of dynamic response. Reducing these tension forces will
also decrease the cost of the chimney as less reinforcement will be required
in the tension zones.
Deformations While the structure will not be inhabited by anyone, the deforma-
tions cannot become too large. Furthermore, large deformations cause fatigue
which could increase the dynamic response in the later years of operation.
Quasi nonlinear static analysis Nonlinear material behaviour is approximated
to determine which areas are prone to crack formations.
Reinforcement The volume of reinforcement steel can be determined based on
the quasi nonlinear behaviour. Optimisation of the structure should not lead

53
5. Analysing the base model

to an increase in reinforcement as this would increase the cost.


Cracking Crack formation has to be mitigated as a reduced stiffness of the rings
would lead to a stronger dynamic response. This analysis visualises which
areas are prone to cracking and could benefit from improvement.
Materials The material quantities for constructing the base model are calculated.
Optimisation should not lead to an increase in overall volume, because this
would also increase the cost.
Buckling Oftentimes, optimisation of the eigenfrequencies will also lead to an
improvement in buckling stability. However, some parameters which might
influence the eigenfrequencies in a positive way, could negatively influence the
buckling stability. Checks have to be done at the end to see if the optimisation
process did not negatively influence the buckling stability.
The dynamic analysis consist of these components:
Modal analysis Before any complex dynamic analyses are conducted, first the
eigenfrequencies of the structure should be determined. Chapter 3 showed
that eigenfrequencies under 0.3Hz can be lead to resonance, so the eigenfre-
quencies alone, accompanied by their mode shapes, already speak volumes
about the dynamic response.
Quasi nonlinear modal analysis By considering quasi nonlinear material be-
haviour, the influence of ring cracking on the eigenfrequencies can be deter-
mined. Lower eigenfrequencies lead to a greater dynamic response, so the
magnitude of this effect has to be analysed.
Harmonic analysis The harmonic analysis produces the mechanical admittance
function as was described in Chapter 3. It depicts which eigenfrequencies
lead to the greatest resonance response and deserve the most attention in the
optimisation process.
Response spectrum The response spectrum shows the dynamic response of the
chimney for along-wind gustiness. From this spectrum, a gust factor can be
derived which represents an amplification factor for the turbulence component
affecting this particular structure.
Vortex shedding Across-wind oscillations can also induce resonance in the chim-
ney if the wind speed reaches the critical wind speed for vortex shedding.
The first few eigenfrequencies have to be checked if they could lead to vor-
tex shedding. If so, these eigenfrequencies have to be increased during the
optimisation process to mitigate this effect.

54
5.2. Summary of the model set-up

5.2. Summary of the model set-up


The set-up for the model summarises all of the design choices which have been
made up until now; geometry parameters which were defined in Chapter 2, actions
which were defined in Chapter 3 and the FE model which was defined in Chapter
4.

5.2.1. Geometry
The geometry for the base model reflects the shape described in Figure 2.1.b. with
the following parameters:

H Height 1000m
Ht Throat height 400m
Rt Throat radius 75m
Rbot Bottom radius 140m
Rtop Top radius 78m (due to a flare of 14%)
φb Angle of inclination at the bottom 13.6◦ (average of the min and max angle)

The wall thickness profile can be seen in Figure 2.1.a. The chimney will have 9
intermediate stiffening rings and one ring at the top, also shown in Figure 2.1.a.
The cross-sectional dimensions of a single ring are 4.5x1.5m (WxH).

5.2.2. Mesh
The chosen mesh division setting is ’intermediate+’. This setting has symmetry
boundary conditions at the edges, with 18 elements in the circumferential direction
and 100 in the vertical direction. It has a mesh refinement factor of 5, meaning
that each element above and below a stiffening ring is divided into 5 elements,
heightwise.

5.2.3. Material properties and actions


The chosen concrete quality for the base model is C50/60 with a Young’s modulus
of 37000N/mm2 . The density of the material itself is 2500kg/m3 . The self-weight,
however, will be set at 25.2kN/m3 due to the inclusion of wind ribs. The stiffening
rings have the same stiffness and density as the concrete used in the rest of the
shell.
The wind loads will be based upon the corrected logarithmic profile as described
in Chapter 3 with a reference hourly mean wind speed of 24.5m/s.
The partial safety factors for the ULS are γG = 1.0 for self-weight and γQ = 1.5
for wind actions (VGB-R-610E, 2010).

5.3. Static analysis


The goal of the research is to optimise the structure of the chimney under dynamic
wind action. An important boundary condition of this optimisation process is that
the static response of the structure is also within certain limits. Tension in the shell
has to be avoided, especially near the foundation, to keep the costs of the structure
down. Furthermore, if compressive stresses become too high, a higher concrete
quality has to be used, bringing about all sorts of new issues such as increased cost

55
5. Analysing the base model

and a lower pumpability (see Appendix A.3). The stiffer the shell, the better it is
able to evenly distribute the forces around the circumference, thus lowering tensile
and compressive stresses throughout. The results of the static analysis are given as
the resultant in-plane forces in the shell elements. Two types of in-plane forces are
characterised, N22, and N11. N22 is the resultant in-plane force in the meridional
direction while N11 is the resultant in-plane force in the circumferential direction
(Figure 5.1). Especially N22 gives a lot of insight in the static response of the
chimney and its ability to evenly distribute forces around the circumference. The
stiffer the shell gets, the more cosine-like the distribution of the meridional forces
around the circumference will be.

(a) N22 (b) N11

Figure 5.1: In-plane forces coordinate system

5.3.1. In-plane forces


The static analysis in this section uses the geometry of the base model and assumes
linear elastic behaviour of the RC shell. Effects such as tension cracking, tension
stiffening, the nonlinear stress-strain relationship in compression, the yielding of
reinforcement and the nonlinear concrete-steel bond are not taken into account
here. This analysis is predominantly being carried out to determine problem areas
in the shell which could use improvement in the optimisation process. The static
results for the base model under dead load and static wind loads (D+1.5W) are
shown in Figure 5.2. The results of the in-plane meridional force (N22) are also
displayed in Figure 5.3 with accompanying values.

(a) N22 windward (left) (b) N11 windward (left)


leeward (right) leeward (right)

Figure 5.2: In-plane forces in the base model

As can be seen, large tensile forces are present on the windward side of the merid-
ian. However, thanks to the hyperbolic shape of the chimney, these tensile forces
peak at a height of around 350m, instead of near the foundation of the chimney.
If a reinforcement percentage of 2% is applied in these tension zones then the ten-
sile stress in the reinforcement steel will top out at 245N/mm2 , well below the

56
5.3. Static analysis

2,000 0.0 −2,000.0

0
00.0 −4,000.0
−2,0
−2,000
N 22 [kN/m]

−6,000.0
−4,000
−6,000
1,000
−8,000 −8,000.0 800
600
0 30 400
60 90 120 200
150 180
angle [φ] height [m]

Figure 5.3: In-plane meridional forces (N22)

design value of 435N/mm2 . The compressive forces dominate the leeward side of
the chimney, peaking at 26.2N/mm2 . This is marginally lower than the design
compressive strength of 33.33N/mm2 for the chosen concrete quality C50/60.

5.3.2. Deformations
The deformed shape is shown in Figure 5.4, with the displacement vector sum being
plotted in Figure 5.5. The maximum displacement at the top of the chimney is
2.27m on the windward side. Most design codes do not specify a top deflection limit
for structures, except for those used in Hong Kong. There, the Code of Practice for
Structural use of Concrete (Hong Kong Buildings Department, 2013) states that
the deflection at the top of the structure should not exceed 1/500th of its height.
Given that the height of the chimney is 1000m, the deflection should not exceed
2m. However, given the fact that the chimney will not be inhabited by anyone,
this limit can be interpretated as a guideline instead of a hard limit.

Figure 5.4: Deformed chimney under under D+1.5W (magnification factor: 50x)

57
5. Analysing the base model

2.0

1.5
2.5

2 1.0

1.5
|U| [m] 1 0.5

0.5 1,000
800
0 600
0 30 400
60 90 120 200
150 180
angle [φ] height [m]

Figure 5.5: Displacement vector sum

5.3.3. Quasi nonlinear static analysis


By reducing the stiffness of tension zones, nonlinear concrete behaviour can be ap-
proximated. Using this approach, the amount of reinforcement can be determined
and elements in which cracks occur can be displayed. The in-plane forces for the
linear and the quasi nonlinear analysis are shown in Figure 5.6 and Figure 5.7.

1,000

800
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 5.6: Linear (red) vs quasi nonlinear (blue) in-plane forces N22 (meridional direction)

58
5.3. Static analysis

5,000

0
N22 [kN/m]

−5,000

−10,000
0 60 120 180
angle [φ]

Figure 5.7: Linear (red) vs quasi nonlinear (blue) in-plane forces N22 (circumferential direction)

5.3.4. Reinforcement
The minimal amount of reinforcement, as prescribed by the VGB guidelines for
natural draft cooling towers, would be 10 · 106 kg. If quasi nonlinearity and thus
crack formation is taken into account, the amount of required reinforcement steel
would be 16 · 106 kg for the base model if 2% of reinforcement were to be applied
in tension zones. The tensile stress in the reinforcement steel would then top out
at 414N/mm2 , much closer to the design value of 435N/mm2 than initially found.
The deflection would increase from 2.27m at the top to 3.24m. All of this is caused
by the cracking of stiffening rings, which make it a lot harder for the chimney to
evenly distribute its stresses around the circumference.

5.3.5. Cracking
For the base model subjected to the default load case, the cracked elements are
shown in Figure 5.8.
Some shell elements along the windward meridian do crack, which was to be ex-
pected given the in-plane force distribution (Figure 5.8). However, most of the
cracks occur in the stiffening rings while trying to keep the shell structure from
ovalising. Unfortunately, these stiffening rings are also a key element in the dy-
namic response of the chimney. If they crack, and their stiffness reduces, so do the
eigenfrequencies of the chimney, therefore making it less effective in counteracting
dynamic wind forces.
In the optimisation process, special attention has to be given to the stiffening rings
to avoid them from developing large tensile stresses. If this is deemed unfeasible,
pretensioning the rings can be considered as an alternate, but costlier solution.

59
5. Analysing the base model

Figure 5.8: Cracked elements under the default load case

5.3.6. Materials

The quantities of concrete and steel required for the construction of the base model
are displayed in Table 5.1.

Element Weight [tonne] Volume [m3 ]


Concrete shell 500720 200228
Concrete rings 88608 35443
Steel reinforcement 15961 2033
Total 605289 237704

Table 5.1: Required quantities for the construction of the base model

5.3.7. Buckling

The buckling modes for the base model have been determined with an eigenvalue
buckling analysis. The partial safety factor for wind loads, originally set at 1.5, has
been increased iteratively until an eigenvalue of 1 is found, as explained in Chapter
4. For the first buckling mode, a partial safety factor of 12.61 is found. This implies
that the buckling safety factor is λ1 = 12.61
1.5 = 8.40. The buckling safety factors
found for the second and third buckling modes are λ2 = 8.41 and λ3 = 10.16
respectively. These values are higher than those found in comparable studies (see
Appendix A.2). One of the reasons for this is that it is a lot harder to compare
buckling safety factors than eigenfrequencies since they are not only geometry- and
material-dependent, but also load-dependent. These values are useful, however, in
comparing the base model to the final optimised model in terms of buckling safety.

On top of that, the buckling mode shapes provide insight in the weak spots of
the structure in terms of buckling. As can be seen in Figure 5.9, the chimney is
weakest where it transitions from a cylinder to a hyperboloid. Note that the safety
factors in real-life will probably be a lot lower due to material nonlinearities and
imperfections in the structure.

60
5.4. Dynamic analysis

Figure 5.9: Mode shapes belonging to the first 3 buckling eigenvalues of the base model under
D+1.5W

5.4. Dynamic analysis

The dynamic analysis of shells differs from that of normal structures. Due to the
thin-walled nature of the shell structure, very complex mode shapes can occur.
Figure 5.10 shows the first six mode shapes of the base model. The first mode is
referred to as a beam bending mode shape, also known as a global mode shape;
the whole structure deflects in the same direction as if it were a cantilever beam.
The other five mode shapes are referred to as shell bending mode shapes, or local
mode shapes; the deformations show various sign changes over the height of the
chimney. The higher the eigenfrequency, the more complex the shell bending mode
shapes usually become. Oftentimes the mode shape only shows deformations at
very localised areas of the shell, leaving the rest of the chimney unaffected.

Figure 5.10: Mode shapes belonging to the first 6 eigenfrequencies of the base model

61
5. Analysing the base model

5.4.1. Modal analysis


The eigenfrequencies associated with the mode shapes seen in Figure 5.10 are shown
in Figure 5.11 and Table 5.2. As stated earlier in Chapter 3, only frequencies below
0.3Hz will have a large impact on the dynamic behaviour of the chimney. For the
base model, this implies that the first two eigenfrequencies are critical, one being
associated with a beam bending mode shape and the other one being associated
with a shell bending mode shape.

0.8

Frequency [Hz]
0.6

0.4

0.2

0
1 2 3 4 5 6
Mode
Figure 5.11: First 6 eigenfrequencies of the base model

Mode number Eigenfrequency


n [Hz]
1 0.1654
2 0.1928
3 0.3404
4 0.4775
5 0.4883
6 0.5438

Table 5.2: First 6 eigenfrequencies of the base model

5.4.2. Quasi nonlinear modal analysis


If the chimney is being subjected to heavy wind loads, some parts of the concrete
shell will crack, resulting in a reduced stiffness of the zones under tension which,
subsequently, causes a shift in the eigenfrequencies. The differences between the
eigenfrequencies of a chimney in the uncracked state and the cracked state are
shown in Figure 5.12 and Table 5.3.

Mode number Eigenfrequency Eigenfrequency


n (uncracked) [Hz] (cracked) [Hz]
1 0.1654 0.1428
2 0.1928 0.1654
3 0.3404 0.3097
4 0.4775 0.3202
5 0.4883 0.3674
6 0.5438 0.4059

Table 5.3: Comparing the eigenfrequencies of an uncracked chimney and a cracked chimney

The first impression is that all of the eigenfrequencies have shifted downwards.
This is not the case, however, because the first eigenfrequency of the uncracked

62
5.4. Dynamic analysis

0.8

Frequency [Hz]
0.6

0.4

0.2 Uncracked
Cracked
0
1 2 3 4 5 6
Mode
Figure 5.12: Comparing the eigenfrequencies of an uncracked chimney and a cracked chimney

state is actually the same as the second eigenfrequency of the cracked state. Both
these eigenfrequencies (0.1654Hz) are associated with a beam bending mode shape,
instead of a shell bending mode shape. Eigenfrequencies associated with beam
bending mode shapes are, apparently, not influenced by the reduced stiffness of
cracked elements. The first mode shape of the cracked chimney is now a shell
bending mode shape which is a very undesired effect since these mode shapes are
more prone to vortex shedding.

5.4.3. Harmonic analysis


After the modal analysis, a harmonic analysis was conducted to study the me-
chanical response of the structure when subjected to harmonic loading. Harmonic
response analyses are used to determine the static response of a structure to loads
that vary harmonically (sinusoidally) with time. By calculating the response of the
structure at a specific range of frequencies and displaying the results of the response
(in terms of the displacement at the top of the chimney), peak responses can be
identified. The response of the structure will be represented by the displacement
of the top node on the windward side. By identifying the peak frequencies and
trying to minimise them, resonance, fatigue, and other damaging effects resulting
from forced vibrations can be overcome.
By dividing the amplitude of the dynamic response by the original static response, a
function describing the dynamic amplification factor can be obtained (Figure 5.13).
This function is also known as the mechanical admittance function, which was seen
earlier in Chapter 3.

101

χm
100

10−1
0 0.1 0.2 0.3 0.4 0.5 0.6
Frequency [Hz]

Figure 5.13: Mechanical admittance function of the base model

63
5. Analysing the base model

5.4.4. Response spectrum


Now all of the variables are known to determine the spectral density of the system
response. The wind force spectrum, as described in Chapter 3, multiplied with the
square of the mechanical admittance function, gives the response spectrum (Fig-
ure 5.14). As a result of the aerodynamic admittance function, the third frequency
and higher are almost completely damped out; only the first two frequencies show
high peaks.
Especially the second eigenfrequency, associated with a shell bending mode shape,
shows high displacements when subjected to dynamic wind loading. In the opti-
misation process, special attention has to be given to these two eigenfrequencies.
By integrating the area under the response spectrum, the variance can be deter-
mined, from which a gust factor can be derived (Chapter 3.6.1). The gust factor
G calculated for the system response of the base model is 1.28.

101

100

10−1
f ·Sy (f )
Iz2 ȳ 2
10−2

10−3

10−4

10−5
0 0.1 0.2 0.3 0.4 0.5 0.6
Frequency [Hz]

Figure 5.14: Spectral density of system response for the base model

5.4.5. Vortex shedding


The Strouhal number for a cylindrical cross-section is 0.2. This means that the
critical wind speed for the first eigenfrequency of the base model will be
fe · d 0.1654 · 150
vcrit,ovaling = = = 124m/s (5.1)
St · N 0.2

This is much higher than any wind speed which might ever occur in the area. The
critical wind speed for the second eigenfrequency, however, is
fe · d 0.1928 · 150
vcrit,ovaling = = = 72.3m/s (5.2)
St · N 0.2 · 2

which is awfully close to limit of 1.25 · vm = 1.25 · 50 = 62.5m/s; i.e. 1.25 times
the characteristic mean wind velocity (averaged over 10 minutes) at the top of the
chimney (EN1991-1-4, 2010). It appears that the second eigenfrequency is not only
critical for along-wind oscillations, but also for across-wind oscillations. As was

64
5.5. Conclusions

seen in the quasi nonlinear modal analysis, cracking can cause the eigenfrequencies
to decrease. If the rings crack, the second eigenfrequency, 0.1928Hz, decreases to
0.1428. Considering the cracked state, the critical wind speed for which vortex
shedding occurs is
fe · d 0.1928 · 150
vcrit,ovaling = = = 53.6m/s (5.3)
St · N 0.2 · 2
Clearly, the rings should not be allowed to crack, otherwise failure due to vortex
shedding becomes a very real possibility.

5.5. Conclusions
Several key areas require attention in the optimisation process. First of all, tension
on the windward side of the chimney has to be reduced to avoid cracking of the
rings; if the rings crack, the eigenfrequencies decrease which would further increase
the danger of along- and across-wind oscillations. Secondly, shape and material op-
timisations should lead to higher first and second eigenfrequencies. Furthermore,
it appears that the second eigenfrequency has an even greater dynamic response
than the first eigenfrequency. Apparently the mode shape associated with the sec-
ond eigenfrequency, a shell ovalisation mode, is more in line with the deformation
behaviour of the chimney under quasi-static wind loading than the beam mode
of the first eigenfrequency. Lastly, the cost of the chimney should also not in-
crease. Preferably, the optimised model uses less material than the base model. In
summary, four different optimisation fronts can be defined:
• Reducing tension on the windward side
• Reducing material use
• Increasing the first eigenfrequency
• Increasing the second eigenfrequency

65
Creating a design tool:
6
SUMAT
In this chapter, SUMAT (Solar Updraft Modal Analysis Tool), a design tool to
quickly study several different chimney configurations will be described. The rea-
sons for designing such a tool and its inner workings will be discussed. Addition-
ally, attention will be given to the limitations of the design tool. A user manual
for SUMAT can be found in Appendix C.

6.1. The need for a design tool


The goal of this research is to optimise the structure of the chimney. In order to do
that, a great amount of chimneys with different parameter sets have to be studied
and compared. It would take too long to model all of these chimneys by hand; a
different solution has to be found to quickly model and compare results. Hence the
creation of the design tool SUMAT (Solar Updraft Modal Analysis Tool).
SUMAT is a UI (user interface) designed to quickly study different chimney con-
figurations. It is programmed in Python 3 using the Qt UI-framework. It
relies heavily on matplotlib, a 2D plotting library, for the visualisation of
analysis data in graphs.

Input parameters Generate Run ANSYS Read results


in SUMAT APDL scripts in batch mode back into SUMAT

Figure 6.1: Flowchart depicting the workings of SUMAT

The way SUMAT works is as follows: first the user can tweak the parameters of
up to five chimneys in a UI. These parameters describe every imaginable aspect of
the chimney, ranging from geometrical shape and placement of the stiffening rings
to material properties and wind loads. Once the user decides that he is satisfied

67
6. Creating a design tool: SUMAT

with the input, he can run one of the following analysis tasks: static analysis,
modal analysis and harmonic analysis. The user interface then uses all of the
parameters defined by the user and starts writing APDL (ANSYS Programmable
Design Language) scripts for the five cases. ANSYS is then booted up in the
background and runs these scripts in batch mode. Once the analysis tasks are
completed, ANSYS stores the results in .txt- and .png-files which are then being
read back into the UI, displaying the results to the user. The entire process is
shown in a simplified way in Figure 6.1.

6.2. Input parameters

The UI is one of the most critical elements in developing a design tool. A balance
has to be found between speed of input and the amount of parameters the user
can change. The more parameters of the chimney are tweakable, the deeper one
can delve into the research. On the other hand, too many parameters can become
confusing and slow down the process considerably. The choice was made to give
each chimney case its own set of parameters so that chimney configurations which
differ on more than one aspect can be compared. By giving each chimney config-
uration the default values belonging to the base model, quick studies can be done
to see the influence of certain parameter changes. A screenshot of the UI, showing
the geometry input tab, can be seen in Figure 6.2.

The input parameters for the different chimney configurations are spread over sev-
eral tabs, grouping them by geometry, material, or loading parameters.

6.3. APDL

The classic version of ANSYS is a command driven program, allowing entire models
to be built and analysed using ANSYS’ proprietary scripting language, APDL.

A small sample of an APDL script, generated by SUMAT, can be seen below. It


shows the first few lines in which material properties for concrete are set and the
thickness for the shell element is defined over the height.

mp,ex,1,37000000000
mp,prxy,1,0.2
mp,dens,1,2500
et,1,SHELL281
mat,1
sect,1,shell,,
secdata,1,1,0.0,3
secoffset,MID
*dim,val,table,6,,,Y
*taxis,val(1),1,0,200.0,400.0,600.0,800.0,1000
val(1) = 0.6,0.45,0.35,0.25,0.25,0.25

An added advantage of APDL is that some advanced features in ANSYS are only
available through the command-line interface. For instance the *GET-command
which allows the user to inquire the model to retrieve all sorts of information and
store it in a variable, allowing for selection procedures which would normally not
be possible.

68
6.4. Batch mode

Figure 6.2: SUMAT user interface, displaying the geometry tab

6.4. Batch mode

Another feature of ANSYS is that it can be run in batch mode, basically meaning
that the program can run several analysis tasks back to back without the GUI
(graphical user interface) being active. The biggest advantage of this is speed,
since most of the computation power is needed to display and continually update
the graphical output of ANSYS. Take this part away and the analysis tasks can be
completed up to several times faster than in the traditional ANSYS environment.
Table 6.1 shows the difference in computation time between GUI mode and batch
mode. Especially static analyses benefit from batch mode since the application of
the wind load profile is a very GUI intensive task. When the batch calculations
are complete, all of the required output can be stored in .txt- and .png-files.

Analysis type GUI mode batch mode Improvement [%]


Static 33.63s 7.68s 337.9%
Modal 8.56s 7.23s 18.4%
Harmonic 40.74s 33.63s 21.1%

Table 6.1: Computation time (GUI mode vs batch mode)

69
6. Creating a design tool: SUMAT

6.5. Output of results


Once the batch process of ANSYS analyses is completed, the results are being read
back into the UI. These results are displayed over various tabs, allowing the user
to quickly inspect and compare them. The three main analysis types available
in SUMAT are, as mentioned before: static, modal and harmonic. The available
results associated with these analyses types are given below:
• Static results
– Meridional in-plane forces over the height
– Meridional in-plane forces over the circumference
– Concrete and steel material usage
– Maximum compressive stress in the concrete
– Maximum tensile stress in the reinforcement steel
– Maximum deflection at the top of the chimney
– Display of cracked elements
• Modal results
– Eigenfrequencies
– Display of mode shapes
• Harmonic results
– Amplitude
– Dynamic amplification factor (mechanical admittance function)
– Response spectrum
– Gust factor
In Figure 6.3, a screenshot is shown displaying the static results tab. On the left is
a plot of the meridional in-place forces on the windward side of the chimney. On
the right the meridional in-plane forces are shown for a horizontal cross-section.
An analysis of this type only takes around 10 seconds, yet it gives the user crucial
information regarding certain design choices.

6.6. Limitations
Since the goal of this research is to study as many chimney configurations as pos-
sible in a relatively short period of time, SUMAT comes with a few limitations.
Given the available time span it would not be feasible to perform computation-
heavy nonlinear FE analyses. Therefore, some concessions have to be made in the
modelling process. These concessions are highlighted and explained below.

6.6.1. Geometrical shortcomings


The geometry which was defined in Chapter 2 is translated into a FE model by
calculating the radius (r) for one hundred points along the z axis. These points
are modelled in ANSYS as so-called keypoints, through which the geometry is
defined. One hundred keypoints for a chimney with a height of 1000m means that
the geometry is only accurate to a detail level of 10m, causing the curvature of
the geometry to not be accurately represented in the FE model. However, this

70
6.6. Limitations

Figure 6.3: SUMAT user interface, displaying the static analysis tab

curvature would be lost anyway in the meshing process. The amount of keypoints
chosen to represent the geometry was therefore closely linked to the chosen mesh
density for the base model.

The chosen mesh density was already described in the convergence study of Chapter
4 and was deemed accurate enough for the purpose of this research.

6.6.2. Material shortcomings


The material model used for concrete is linear elastic. Concrete, however, behaves
very differently under tension than under compression. As long as the concrete
shell is loaded entirely in compression then the assumption of a linear elastic model
is correct (not taking into account creep and shrinkage effects). If large tension
stresses occur, measurements will have to be taken. The influence of concrete
cracking is incorporated in SUMAT in a limited way. What happens is that once
the static analysis is completed, SUMAT checks to see which elements have tensile
stresses which exceed the tensile strength (fctd ) of concrete, and lowers the stiffness
of those elements by 70%. Once this step is completed, the analysis is run again
to see what the influence is of locally reducing the stiffness. This iterative analysis
procedure tries to emulate nonlinear analyses, while still retaining the speed of
linear analyses.

If the tensile stress exceeds the tensile strength of concrete in an element, extra

71
6. Creating a design tool: SUMAT

reinforcement is added to all elements at that height, since the wind can attack
from every possible angle. This process of selectively lowering the stiffness and
applying reinforcement is shown in Figure 6.5.

6.6.3. Loading shortcomings


Because the definition of loads is a lengthy process in ANSYS, the loads will be
applied not on a per-element basis but a per-area basis. A single area consists of one
element in the circumferential direction and 10% of all elements in the meridional
direction, since the loading fluctuates more in the circumferential direction than
it does in the meridional direction (see Chapter 3). Average wind pressures are
calculated for these areas and then applied on the chimney.

6.6.4. Accuracy of results


SUMAT only looks at the in-plane forces of the chimney. These membrane forces
are the resultant of all the internal stresses. The in-plane force divided by the
thickness is the average stress which SUMAT uses to check if a concrete element is
cracked. Because it uses the average stress in the element, it could occur that some
elements which are heavily under tension on one side and heavily under compression
on the other side are not considered as a cracked element by SUMAT. On the other
hand, elements which are considered as a cracked element might have large regions
of uncracked concrete material.
In Figure 6.4 one can see the stress distribution in a typical stiffening ring. Tensile
stresses occur along the entire perimeter of the ring, but oftentimes switching from
the inner side to the outer side. Next to that, the cracked elements according to
the algorithm used by SUMAT are shown.

(a) Principal stress distribution in a typical (b) Cracked elements according to SUMAT
stiffening ring

Figure 6.4: Comparison between actual stress distribution and in-plane force simplification

The fact that SUMAT only looks at the resultant forces might seem like an over-
simplification at first, but the goal of this research is to optimise the design of
a chimney; an optimal design is one which is mostly loaded in compression, to
minimise the amount of reinforcement required. And, in the case where large ten-
sile forces do occur in the concrete shell, the rings can be pretensioned to avoid
cracking. As long as the stiffening rings are intact, the structural behaviour of the
chimney will be similar to that of an entirely uncracked chimney.

72
6.7. Conclusions

Start

i=1

Select all elements


at ith % height

Calculate
average thickness

Calculate
crack force

yes
Select new element

yes Is element no
cracked?

Lower stiffness
of element

Select entire no
Elements left?
ring of elements

Apply extra
reinforcement

Select remaining
elements

Apply base
reinforcement

i=i+1

Figure 6.5: Flowchart depicting the reinforcement process

6.7. Conclusions
While not without its limitations, a comprehensive design tool has been developed
to aid in this research. Only a few assumptions were made initially when design-
ing the tool, giving the user more freedom in designing a chimney configuration.
Furthermore, while it costs a lot more effort to program and calibrate the design
tool, it allows for many more configurations to be studied and ultimately provides
more insight than could ever be amassed by using traditional modelling methods.

73
Sensitivity analysis
7
This chapter depicts the process which was followed to determine the influence of
the geometric and material parameters on the static and dynamic response, and
the cost of the chimney. Each parameter set consists of five chimney configurations.
One configuration represents the original base model and the other four represent
similar configurations where only the studied parameter is varied. These parameter
variations are chosen in such a way that they still exist in the domain of feasibility.

7.1. Introduction

In Appendix B, the results of the various parameter sets are shown, the conclu-
sions of which will be described here. The parameter sets fall into the following
categories: geometric parameters, material parameters, loading parameters and
damping parameters. The parameters will be scored in so-called sensitivity plots
on their ability to improve the four key problem areas defined at the end of Chapter
5. Each parameter has 4 different score ranges, one for each key problem area. A
score range can exist in the domain of -10 to 10; a positive value indicating an
improvement for that particular key problem area, and a negative value indicating
a regression for that particular key problem area. A plus sign indicates an increase
in the parameter (f.i. an increase in throat height, ring stiffness, wall thickness,
etc.), while a minus sign indicates a decrease in the parameter (f.i. a decrease in
the number of rings, concrete density, bottom radius, etc.). An example is shown in
Figure 7.1, where an arbitrary parameter has a fictional score range of -3 to 7, with
a plus sign at the top and a minus sign at the bottom. This means that increasing
the value of that parameter has a large positive effect on the key problem area,
and decreasing the value of that parameter has a small negative effect.

After all of the parameter sets are analysed and described, conclusions are derived
which can be used as the input for the optimisation process.

75
7. Sensitivity analysis

10

Score
0 Key problem area

−5

−10

Figure 7.1: Example parameter sensitivity plot

7.2. Scale

The scores which are assigned to each of the individual key problem areas exist in
the domain of -10 to 10. These ranges correlate to the following numerical values:

Reduce tension A score of zero is equal to the peak tension force in the base
model, namely 1000kN/m2 . Scores of -10 and 10 are respectively equal to
peak tension forces of −500kN/m2 and 2500kN/m2 .

Reduce material use A score of zero is equal to amount of concrete material


required for the base model, 589 · 106 kg. Scores of -10 and 10 are respectively
equal to material increments and reductions of 15%.

Improve 1st eigenfrequency A score of zero is equal to first eigenfrequency of


the base model, 0.1654Hz. Scores of -10 and 10 are respectively equal to
eigenfrequencies of 0.1154Hz and 0.2154Hz.

Improve 2nd eigenfrequency A score of zero is equal to first eigenfrequency


of the base model, 0.1928Hz. Scores of -10 and 10 are respectively equal to
eigenfrequencies of 0.0928Hz and 0.2928Hz. Note that it is easier to improve
the second eigenfrequency, hence the larger range.

An example plot is shown in Figure 7.2. The first bar from -5 to 5 indicates
that tension can be reduced up to approximately 250kN/m2 if the parameter is
increased to the maximum. The second bar, however, indicates that doing so will
cost a tremendous amount of material. Note that if a bar is coloured black, it
does not fit in the previously set range of -10 to 10, implying in this case that
the material use will decrease/increase more than 15%. Due to this uncertainty,
parameter sets with black bars should be used with caution.

The third bar is missing, suggesting that the first eigenfrequency is not influenced
by this parameter change. Lastly, the fourth bar shows that increasing this param-
eter also marginally improves the second eigenfrequency.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.2: Scale example sensitivity plot

76
7.3. Geometric parameters

7.3. Geometric parameters


Throat height Increasing the height of the throat radius has a positive effect on
the static response of the chimney. The higher the throat is located, the
less tension will be present in the shell structure. This is because of two
reasons: first of all, the increase in mass because the chimney will taper at
a slower pace, and secondly due to the increased lever arm of the chimney.
The dynamic response is influenced both positively and negatively. The first
eigenfrequency, associated with a beam bending mode shape, becomes higher
as the throat height increases. The remainder of the first few eigenfrequencies,
all associated with shell bending mode shapes, actually become lower as the
throat height increases. Considering that the first two eigenfrequencies are
the most critical in the design, increasing the throat height as a marginally
positive effect on the structural response of the chimney, coming at a cost of
more material usage.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.3: Throat height sensitivity plot

Radius throat Increasing the radius of the throat of the chimney has a negative
influence on the static response as well as the dynamic response (apart from
the first eigenfrequency). The logical conclusion would be to decrease the
radius of the throat as this would also decrease the material usage. How-
ever, considering that the radius of the throat is directly correlated to the
power output of the solar updraft tower, it would be unwise to modify this
parameter.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.4: Radius throat sensitivity plot

Radius bottom Decreasing the radius at the bottom of the chimney actually
reduces tension in the chimney, because the compressive stresses due to the
self-weight will become higher in the lower part of the chimney, compensating
for the tensile stresses caused by wind loads. The eigenfrequencies, however,
suffer if the radius at the bottom is decreased. Normally, an improvement in
the eigenfrequencies will also lead to an improvement in buckling behaviour.
This parameter seems to be the exception though, as the wider the base gets,
the likelihood of punch-through failure occurring increases, decreasing the
overall buckling safety of the structure.

77
7. Sensitivity analysis

10

5 Reducing tension

Score
Reducing material use
0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.5: Radius bottom sensitivity plot

Angle of inclination A small angle of inclination (measured from the vertical


axis) has an especially negative influence on the static behaviour. If the
transition from cylinder to hyperboloid is too abrupt and harsh, large peak
stresses will occur in this transition region. The eigenfrequencies are barely
influenced by the change in angle of inclination. A small angle will have a
minor positive effect on the first eigenfrequency, but this is mostly due to
the fact that the chimney will have less mass if the angle is very small; a
minimal angle of inclination will cause a straight line between the bottom of
the chimney and the throat while a maximal angle of inclination will curve
inwards as much as possible, requiring slightly more material (Figure 7.6).

Figure 7.6: Angle of inclination (minimum in red, maximum in blue)

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.7: Angle of inclination sensitivity plot

Wall thickness (overall) Increasing the overall wall thickness has a positive ef-
fect on the static behaviour of the chimney. More mass equals higher dead
loads, resulting in a chimney that is better able to counteract the wind loads.
However, if the wall thickness is increased too much, compressive stresses on
the leeward side of the chimney will become too high. As far as eigenfrequen-
cies go, the first mode is not influenced at all by the change in wall thickness;
the higher modes, associated with shell bending mode shapes, do show posi-
tive change as the wall thickness is decreased. If the wall is thinner, the ratio
between the stiffness of the wall and the stiffness of the rings increases, thus
improving the eigenfrequencies which rely on the stiffening rings.

78
7.3. Geometric parameters

10

5 Reducing tension

Score
Reducing material use
0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.8: Wall thickness (overall) sensitivity plot

Wall thickness (bottom) Increasing the wall thickness of the bottom half of the
chimney has a negative influence on the static behaviour. It does not help the
chimney in constraining wind loads, but it does increase compressive stresses
at the bottom of the chimney. The only positive effect of increasing the
wall thickness at the bottom of the chimney is a minor increase in the first
eigenfrequency, other than that, the benefits do not outweigh the costs.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.9: Wall thickness (bottom) sensitivity plot

Wall thickness (top) Unlike increasing the wall thickness at the bottom, in-
creasing the wall thickness at the top does have a positive effect on the static
behaviour. The thicker this part of the chimney becomes, the better the shell
can restrain wind loads. Making the chimney top-heavy does have a negative
influence on the eigenfrequencies, as well as the buckling behaviour of the
chimney, increasing the odds that punch-through failure might occur.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.10: Wall thickness (top) sensitivity plot

Number of rings (evenly spaced) The sensitivity analysis of the number of


rings reveals just how important these really are in the structural response
of the chimney. A decrease in number of rings has a very negative influence
of the static behaviour, decreasing the chimneys ability to counteract wind
loads. The fewer rings, the less cosine-like the circumferential load carrying
capacity becomes. As stated before, the first eigenfrequency is barely influ-
enced by the number of stiffening rings and apparently, neither is the third.
The second, fourth, fifth and sixth mode do show large fluctuations if the

79
7. Sensitivity analysis

number of rings is altered. Especially the higher modes, having more and
more complex mode shapes, are negatively influenced if the number of rings
is decreases. However, as was found in Chapter 5, the first two modes are
the most critical in the dynamic response of the chimney.

10

5 Reducing tension

Score
Reducing material use
0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.11: Number of rings (evenly spaced) sensitivity plot

Number of rings (spaced 100m from the top) This sensitivity analysis shows
just how important the rings at the top of the chimney are and how little
the rings at the bottom contribute to the overall load carrying capacity. Re-
moving the bottom ring has almost no influence on the static and dynamic
response of the chimney; the hyperboloid shape of the chimney ensures that
this part of the structure is stable. Removing the bottom two rings also shows
very little change in the static response of the chimney, while it does influ-
ence the eigenfrequencies of the higher modes. The same goes for removing
the bottom three rings. Removing the bottom four rings, however, yields
negative results for the static as well as the dynamic response.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.12: Number of rings (spaced 100m from the top) sensitivity plot

Number of rings (spaced 100m from the bottom) The importance of the top
stiffening rings becomes even more apparent in this sensitivity analysis. Re-
moving even one ring from the top yields very unfavourable results. If two
or more rings are removed from the top half of the chimney, it will almost
certainly fail due to high tension stresses on the windward side of the chimney.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.13: Number of rings (spaced 100m from the bottom) sensitivity plot

80
7.4. Material parameters

Aspect ratio rings By changing the aspect ratio of the rings (thus increasing
the width while decreasing the height), its moment of inertia can be altered
without changing its weight. The sensitivity analysis clearly shows that thin-
ner, wider rings with a larger moment of intertia have a positive effect on
the static as well as the dynamic behaviour. The logical response would be
to make the rings as thin as possible. This would, however, ultimately lead
to the rings collapsing under their own weight. Therefore, a maximum as-
pect ratio of 6.75:1 is chosen for the rings, based on technical drawings for
pre-existing designs (see Appendix A.3).

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.14: Aspect ratio rings sensitivity plot

7.4. Material parameters


Density of concrete Lightweight concretes benefit from lightweight aggregates,
which leads to a reduction in the density of the material. They can be used
for structural applications as well, with strengths equivalent to normal con-
crete mixtures. The problem is that by reducing the density, the dead loads
of the chimney will also decrease, thus increasing tensile stresses on the wind-
ward side. Similar results were obtained by varying the wall thickness of the
chimney; the main difference being that a lower density concrete has a better
buckling resistance than a thinner wall with regular density concrete. Fur-
thermore, not only the weight of the wall is reduced, but also of the stiffening
rings, resulting in higher eigenfrequencies for chimney configurations using
lightweight concretes.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.15: Density of concrete sensitivity plot

Concrete quality High-strength concretes have three advantages; a higher com-


pressive strength, a slightly higher tensile strength, and a higher Young’s
modulus. The compressive strength allows for thinner shell walls, the higher
tensile strength allows for fewer crack formations, while the higher Young’s
modulus will result in marginally improved eigenfrequencies all across the
board. The static response, however, does not change in the slightest when
only linear material behaviour is considered. Furthermore, the eigenfrequen-
cies are only increased marginally, thus making high-strength concretes too

81
7. Sensitivity analysis

costly to be considered as the first solution path.

10

5 Reducing tension

Score
Reducing material use
0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.16: Concrete quality sensitivity plot

Ring stiffness (material properties, all 10 rings) Increasing the ring stiffness
allows for the chimney to more evenly distribute the wind forces around the
circumference, as can be seen from the increasing cosine-like behaviour in the
circumferential direction. The eigenfrequencies associated with shell bending
mode shapes also show minor improvements. In this case, the stiffness was
increased by changing the Young’s modulus of the ring elements. This can
be accomplished in real-life by using a higher concrete quality. Note that
increasing the concrete quality only allows for a maximal stiffness increase of
20% (the stiffness of C105/C90 is only 20% higher than that of C50/60).

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.17: Ring stiffness (material properties, all 10 rings) sensitivity plot

Ring stiffness (material properties, top 5 rings) As was seen before, the top
rings are the most important in restraining the chimney counteracting dy-
namic responses. When only the stiffness of the top 5 rings is increased,
similar results are obtained as when the stiffness of all rings was increased.
It would therefore be unwise to increase costs by increasing the stiffness of
rings that contribute little to the load carrying capacity of the chimney.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.18: Ring stiffness (material properties, top 5 rings) sensitivity plot

Ring stiffness (cross-sectional area, all 10 rings) Increasing the cross-sectional


area of the rings gives similar results as increasing the Young’s modulus of
the material of the ring elements. Tension forces are reduced on the wind-
ward side of the chimney and most of the eigenfrequencies show significant

82
7.5. Loading parameters

improvements. The only difference is that by increasing the size of the rings,
the dead load also increases, leading to high compressive stresses if the rings
become too large.

10

Score 5 Reducing tension


Reducing material use
0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.19: Ring stiffness (cross-sectional area, all 10 rings) sensitivity plot

Ring stiffness (cross-sectional area, top 5 rings) Increasing the cross-sectional


area of only the top 5 rings produces similar results as when the dimensions
of all rings are increased; tension is reduced on the windward side and the
eigenfrequencies show significant gains. The main difference is that the dead
loads do not increase as much as when the cross-sectional area of all rings is
increased.

10

5 Reducing tension
Score

Reducing material use


0 Improving 1st eigenfrequency
Improving 2nd eigenfrequency
−5

−10

Figure 7.20: Ring stiffness (cross-sectional area, top 5 rings) sensitivity plot

7.5. Loading parameters

Wind speed The sensitivity analysis for wind speed shows that the chimney needs
to be built on a location where extreme wind forces do not prevail, otherwise
the structure needs to be heavily modified to be able to withstand these large
wind forces. Fortunately, most desert areas are inland, at a great distance
from the ocean, usually resulting in rather modest wind climates.
Altitude AMSL The greater the altitude of the location of the chimney, the lower
the air density will be and thus the lower the wind forces will be. Building the
chimney of the base model at an altitude of 2000m would already lower wind
forces so much that all existing tension forces in the chimney would vanish.
The problem is that locations with high altitudes are usually mountainous
areas which are almost never flat enough to construct a solar updraft tower,
let alone the issues with getting all the materials on site.
Latitude The latitude of the location of the chimney has a much larger influence
on the static response than initially thought. The Coriolis forces become
sufficiently large near the top of the chimney resulting in an increase in tensile
forces on the windward side of the chimney.

83
7. Sensitivity analysis

Return period Designing a chimney for a longer lifetime would drastically influ-
ence design decisions. Tensile forces on the windward side would become so
large that the dimensions of the stiffening rings have to be increased by a
large factor, and most likely have to be heavily pretensioned.

Pressure distribution curve The analysis shows that the influence of the wind
ribs is relatively small. If no wind ribs were applied and a different pressure
distribution curve had to be used, the static response would only show minor
changes. There would be a little more tension on the windward side in the
meridional direction and the compressive stresses at the sides of the chimney
would increase marginally, while the compressive stresses at the leeward side
of the chimney would be slightly lower.

7.6. Damping parameters

Constant damping ratio A minor increase in the damping ratio results in a


large reduction of the resonance amplitude peaks. If the dynamic response
cannot be mitigated in any other way, extra damping elements could be
installed in the structure to counteract the dynamic wind forces.

7.7. Conclusions

The above results are summarised per key problem area in Figures 7.21, 7.22, 7.23
and 7.24. By carefully examining the sensitivity plots throughout this chapter, the
following conclusions can be drawn. First of all, to reduce tension on the windward
side of the chimney, several measurements can be taken. Shape optimisations such
as increasing the throat height and decreasing the radius at the bottom of the
chimney can help in reducing tension stresses, but the most can be gained by
optimising the stiffening rings. Increasing the wall thickness also helps, but is a
lot more costly. The material spent on wall thickness can better be allocated to
increasing the dimensions of the rings. Improving the aspect ratio of the stiffening
rings also reduces tension in the shell; the rings become much stiffer due to the
increase in moment of inertia while not requiring more material.

As it turns out, optimising the first eigenfrequency is a lot more difficult than opti-
mising the second eigenfrequency; only a few parameters show actual improvement
in the first eigenfrequency. For instance, lowering the density of the concrete is a
bad idea as the tension forces in the chimney would become much larger as the
dead load will no longer be large enough to compensate them. A combination of
shape parameters will have to be used to improve the first eigenfrequency such as
increasing the throat height and redistributing the wall thickness profile so that
more mass will be at the bottom in comparison to the top, although this could also
have a negative effect on the static response of the chimney.

The second eigenfrequency, associated with a shell bending mode shape, can be
optimised easily by improving the stiffness of the rings. By modifying the aspect
ratio of the rings and adding more material to the top rings (which can be won
back by removing rings at the bottom of the chimney, where they are not really
needed), most of the dangers associated with the second eigenfrequency can be
averted.

As far as material properties go, improving the concrete quality seems like a costly
affair with no noteworthy benefits. Lightweight concretes on the other hand, do

84
7.7. Conclusions

show some benefits in terms of dynamic response, while the decreased dead load
negatively influences the static response.
The sensitivity analysis for the loading parameters shows that the chimney cannot
be constructed in windy climates, otherwise costly concessions have to be made
such as increasing the overall wall thickness, and increasing the dimensions of the
stiffening rings by a great amount. The same investments have to be made if a
SUT with a longer design life is required, due to the increase in the reference hourly
mean wind speed if the return period is increased. While the previous results were
to be expected, what came as a bit of a surprise was the large influence of the
location of the SUT on the static behaviour. Apparently, for very tall structures,
the Coriolis effect and the influence of air density cannot be neglected and prove
to be an important design parameter.
In the end, the difficulty lies in combining the parameters. While some parameter
changes show an improvement in reducing tension forces in the shell, they show
regression in the eigenfrequencies. Furthermore, some parameter changes show
improvement in the first eigenfrequency (a beam bending mode), and regression in
the second eigenfrequency (a shell bending mode), or vice versa. Given that the
first two eigenfrequencies are crucial in the dynamic response, both of these values
have to be improved. Chapter 8 will deal with finding a suitable combination of
parameters to improve the static and dynamic response of the chimney.

10

5
Score

−5

−10
t

on

qu e
10 ity
th (bo ll)

pa ced s (e nes m)

s)

s)

s)
bo t
m
gh

op

p
om

gs
ng

t
a

ng

ng
et cre
in tto

ce

to
ra
ro

al
o

in

to rin
at
ad hei

(t

ri
tt

tt
a

ri

ri
W ne ove
th

r
e
in

sp

ty io
bo
s

th

co

ea 10

5
t

cl
R us

t
(

e
oa

10 0m ly

ra

p
m

A the

f
s
iu
i

l
ec al a to
o
ss

k
th nes
hr

en

al

al
cr
o
ad

ic

ct
f

fr
al le o
T

v
R

s
on

on rea
m

pe
ck

on tie
si

pr tie
k

ar
o

C
ic

en
s
W thi
ng

fr

r
er
al

e
g

10

al
D

op
A

st ss ( ter rop
sp rin

0m
l

l
al
W

ti
f

ti
ng r o

l
ec
l
a

ia
iff s (m ria
d

-s
be

-s
ce

te

ss
ng s (

ss
be r o um

ro
in iffn (ma

ne ro
(s

(c
N

c
ri

ss
st ss
es
f

ne
ri

in ffn

iff
um be

of

i
st

st
N um

g
g

g
N

in
in
R

R
R

Figure 7.21: Reduce tension sensitivity plot

85
86
Score Score
T T
hr hr

−10
−5
0
5
10
−10
−5
0
5
10

oa oa
R t R t
ad he ad hei
i ig iu gh
us ht t
R R s
A ad th A ad th
ng i r ng i r
W us oa W us oa
al le o bo t al le o bo t
l f t l f t
W t in t W t in t
N hi
ck cl om N hi
ck cl om
7. Sensitivity analysis

N um al in N um al in
l l
um b e N th nes
s
at
i
um be N th nes
s
at
i
ic ic
be r o um
f k n
(o on be r o um
f k n
(o on
r ri b er Wa es v er r ri b er Wa es v er
of s of s
ri
ng of ll a ri
ng of ll a
ng s ( th (bo ll) ng s ( th (bo ll)
s sp rin
g
ic
k
tt s sp rin
g
ic
k
tt
(s a o (s a o
pa ced s (e nes m) pa ced s (e nes m)
ce v s ce v s
d 10 en (t
op d 10 en (t
op
10 0m ly 10 0m ly
0m sp ) 0m sp )
fr a fr a
R o m c R o m c
in f ro e in f ro e
g m th d) g m th d)
st st
R iff A th e t R iff A th e t
in ne sp e b op) in ne s p e b op )
g ec ot g ec ot
R st ss ( D t to R st ss ( D t to
in iffn ma en ra m in iffn ma en ra m
g es te s i t io ) g es te s i tio )
st s r ty st s r ty
R i i of r in R i i of r in
in ffn (ma al p
e C g in ffn (ma al p
e C g
g on co s g on co s
st ss ( ter rop cr n st ss ( ter rop cr n
iff c ia er iff c ia er
l t
et cre
e l t
et cre
e
ne ro
ss ss pr i te ne ro
ss ss pr i te
-s op es a qu -s op es a qu
(c e al (c e al
ro cti ert ll 1 it ro cti ert ll 1 it
ie 0 y ie 0 y
Figure 7.22: Reduce material use sensitivity plot

ss on ss on
-s s ri -s s ri
ec al a to ng ec al a to ng
ti p ti p

Figure 7.23: Improve first eigenfrequency sensitivity plot


on rea 5 s) on rea 5 s)
ri ri
al al
l n gs
al al
l n gs
ar ar
ea 10 ) ea 10 )
to rin to rin
p gs p gs
5 ) 5 )
ri ri
ng ng
s) s)
Score
T
hr
o

−10
−5
0
5
10
at
R he
ad ig
i us ht
R
A ad th
ng i r
W us oa
al le o bo t
l f t
W t in t
N hi
ck cl om
N um al in
l
um be N th nes
s
at
io
ic ( n
be r o um
f k
r ri be W ne ove
of ng r o al ss ra
ri f l
ng s ( th (bo ll)
s sp rin
g
ic
k
tt
(s a o
pa ced s (e nes m)
ce v s
d 10 en (t
op
10 0m ly
0m sp )
fr a
R o m ce
in fr d
g o m th )
st e
R i A the to
in ffn
e s p )
g pe bo
R st ss D ct tt
om
in iffn (ma en ra
g es te si t )
st ty io
R iff s (m ria o ri
in l C f ng
g
ne a p on co s
st ss ( ter rop cr n
iff c ia er
ne ro l et cre
e t
ss pr tie
ss -s op s qu e
(c ec e al al
ro ti r l
ss on tie 10 ity
-s s ri
ec al a to ng
ti p s)
on rea 5

Figure 7.24: Improve second eigenfrequency sensitivity plot


al al r in
ar l gs
ea 10 )
to rin
p g s)
5
ri
ng
s)

87
7.7. Conclusions
Optimisation of the base
8
model
In this chapter, the results of the sensitivity analysis are used to optimise the base
model manually using the design tool described in Chapter 6. The first step of the
optimisation process is to determine the usefulness of all of the parameters which
were researched. The second step consists of gradually introducing these parameter
changes into the base model.

8.1. Multi-objective optimisation


Multi-objective optimisation deals with optimising various objectives which may
or may not have conflicting interests. In this research, the objective functions of
the multi-objective optimisation process are the four key problem areas which were
described before. The first step of this process is finding a Pareto optimal solution
(Figure 8.1). A Pareto optimal solution is one in which any further improvement of
one of the objective functions will lead to a regression in one or more other objective
functions. Once a Pareto optimal solution is found, trade-offs can be made between
objective functions to further optimise the solution, essentially searching for another
Pareto optimal solution which favours other objective functions. These trade-offs
are made by a so-called decision maker (DM).
Several methods exist for finding an optimal solution in multi-objective optimisa-
tion processes:
No preference method No DM is required, all objective functions are treated
equally. No preference information is required.
A priori method Preference information is asked from the DM to find a Pareto
optimal solution.
A postori method The DM can choose from a representative set of Pareto op-
timal solutions which were already found.

89
8. Optimisation of the base model

Interactive method In each step of the interactive method, the DM is presented


with a Pareto optimal solution. The DM can then choose a direction in
which to steer the optimisation process, allowing for the solution to be found
iteratively.

Figure 8.1: Pareto optimal objective vectors (Adapted from Miettinen (1999))

In this case the DM is human, with the requirement that this person is an expert
in the problem domain (Miettinen, 1999). A modified version of the interactive
method is used to find a solution which satisfies all of the objective functions and
thus improves all of the key problem areas.

8.2. Solution path


The solution path in which an optimised solution is to be found consists of a step-
by-step process where parameter changes are gradually introduced into the model.
The geometric and material parameters which were analysed in the sensitivity
analysis in Chapter 7 are subdivided into four different categories, based on their
usefulness in optimising the four key problem areas. The categories are:
Primary parameters which only bring about positive change. At least one of
the four objective functions should show improvement while the other three
cannot show any regression whatsoever. These are used to find a Pareto
optimal solution.
Secondary parameters which bring about mostly positive change, only one ob-
jective function is allowed to show regression. Furthermore, the advantages
should outweigh the disadvantages; i.e. the objective functions which are to
be improved should show more gains than those which suffer from these pa-
rameter changes.
Tertiary parameters which function as a last resort measure. If one of the objec-
tive functions is found to be lacking in terms of optimisation, then gains from
other objective functions can be sacrificed by introducing these parameter
changes; this is to make sure that all four objective functions show improve-
ment when compared to the base model.
Residual parameters which should not be tampered with, either because they
only negatively influence the four objective functions or because they come
with unacceptable side-effects.
The second step consists of gradually introducing the primary, secondary and ter-
tiary parameter changes into the base model, to find another Pareto optimal so-

90
8.3. Categorising the parameters

lution which improves all of the four objective functions. As stated above, the
tertiary parameter changes will only be introduced if one of the objective functions
has not improved compared to the base model, or worse, has deteriorated due to
the introduction of secondary parameter changes. The results of each optimisation
step will be displayed in a Kiviat diagram (Figure 8.2), so that they can easily be
compared to the original base model. The scale of the Kiviat diagram is similar to
that of the sensitivity plots in Chapter 7. The Kiviat diagram for the base model is
located in the centre of each axis, indicating a score of zero. If the Kiviat diagram
moves outwards then the score is positive (with a maximum of 10), if it moves
inwards on an axis then the score is negative (with a minimum of -10).

Reduce
material use

Improve
Reduce
1st eigen-
tension
frequency

Improve
2nd eigen-
frequency

Figure 8.2: Kiviat diagram for the base model

8.3. Categorising the parameters

The parameters will now be subdivided into the aforementioned four categories,
based on the results from the sensitivity analysis.

8.3.1. Primary parameters


Aspect ratio rings Increasing the aspect ratio of the rings significantly reduces
tension on the windward side of the chimney. Furthermore, the second eigen-
frequency also shows marked improvements. This change does not affect the
material use and the first eigenfrequency.

Number of rings Removing rings from the bottom of the chimney drastically
decreases material use. Furthermore, if only one or two rings are removed
then the tension forces on the windward side and the first and second eigen-
frequency show no discernible difference. If more rings were to be removed
then some of the material gained has to be added to the stiffening rings at
the top to make up for the increase in tension. If the number of rings is
reduced further then perhaps the spacing between the rings (now fixed at
100m) should be modified to further optimise the chimney. Note that rings
should never be removed from the top of the chimney, as this would lead to
extremely large tension forces and instability problems.

91
8. Optimisation of the base model

8.3.2. Secondary parameters


Throat height Increasing the throat height has a positive effect on the reduction
of tension and improves the first and second eigenfrequency. The material
use does increase, but the benefits are larger than the drawbacks.

Ring stiffness (cross-sectional area) Increasing the stiffness of the rings by ex-
panding the cross-sectional area reduces tension on the windward side and
improves the second eigenfrequency. The increase in material use can be
counteracted to some extent by only increasing the cross-sectional area of the
top rings; this has the same effect as increasing the cross-sectional area of all
rings, but uses less material. Note that the cross-sectional area of the rings
should not be increased too much, or the first eigenfrequency will suffer.

8.3.3. Tertiary parameters


Wall thickness Adding more material to the bottom of the chimney could favour
all of the objective functions except for the reduction of material use. De-
creasing the wall thickness at the top, on the other hand, favours all objective
functions except for the reduction of tension. Reduction of the wall thickness
at the top should be done with careful deliberation as too large reductions
could lead to instability and problems with constructability aspects such as
formwork.

Radius bottom Increasing the radius at the bottom increases the first eigenfre-
quency and also decreases the deflection of the chimney, but it comes at the
cost of increased tension on the windward side and an increase in material
use.

8.3.4. Residual parameters


Radius throat While decreasing the radius of the throat leads to improvement
in three of the four objective functions, this parameter will not be considered
in the optimisation process. The reason for this is that decreasing the throat
height would decrease the energy output of the chimney as less air will be
able to travel upwards through the chimney.

Angle of inclination Decreasing the angle of inclination any further would im-
prove three out of four objective functions, however, it would also lead to
very high peak stresses in the chimney where it transitions from a cylinder
into a hyperboloid. Therefore the current setting of the angle of inclination
is considered to be the minimum for this research.

Density of concrete Using lightweight concretes would drastically increase ten-


sion on the windward side of the chimney as the dead load of the concrete
would no longer be enough to counteract the tension forces caused by wind.

Concrete quality Increasing the concrete quality would also increase the cost
of the chimney. The advantages from using a higher quality concrete do not
outweigh this increase in cost. Furthermore, higher quality concrete mixtures
have a lower pumpability (see Appendix A.3).

Ring stiffness (material properties) This parameter increases the concrete qual-
ity of the stiffening rings, as opposed to the entire concrete shell. Similar to
increasing the overall concrete quality, the minor gains are not worth the cost
and effort of using higher quality concrete mixtures.

92
8.4. Optimisation step 1

8.4. Optimisation step 1

The first optimisation step only allows for the introduction of primary parameter
changes. The primary parameters are the increase of the aspect ratio of the rings,
and the decrease of the number of rings (within certain boundaries). The sensitivity
analysis shows that the bottom two rings can be removed without any significant
impact on the static as well as the dynamic response. However, if the aspect ratio
of the rings is also increased to 6.75x1m (WxH), then tension on the windward
side is reduced by such a great amount that the bottom three rings can be removed
and the chimney is still fully loaded in compression (Figure 8.3). This allows for
a further decrease in material use; material which can later on be used to enhance
the wall thickness at certain places or increase the cross-sectional area of the rings
where needed.

1,000
Base model
Bottom 2 rings removed
800 Bottom 3 rings removed
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 8.3: Optimisation step 1 in-plane forces N22 (meridional direction)

0.6
Frequency [Hz]

0.4

0.2 Base model


Bottom 2 rings removed
Bottom 3 rings removed
0
1 2 3 4 5 6
Mode
Figure 8.4: Optimisation step 1 eigenfrequencies

As can be seen in Figure 8.4, the first eigenfrequency shows no change due to these
parameter changes. The second eigenfrequency, on the other hand, shows a lot
of improvement, increasing from 0.1920Hz to 0.2749Hz. The results of the first
optimisation step in regards to the four objective functions are shown in the Kiviat
diagram in Figure 8.5.

93
8. Optimisation of the base model

Reduce
material use

Improve
Reduce
1st eigen-
tension
frequency

Improve
2nd eigen-
frequency

Figure 8.5: Kiviat diagram for the first optimisation step

8.5. Optimisation step 2


The output from the previous step will serve as the input for this one. Now that all
of the primary parameters have been tackled and a Pareto optimal solution has been
found, trade-offs can be made using the secondary parameters to further optimise
the four objective functions. The secondary parameters are the throat height,
and increasing the cross-sectional area of the stiffening rings, more specifically, the
top rings, because increasing their dimensions proved to be the most beneficial.
The sensitivity analysis shows that increasing the cross-sectional area of the top
rings by a factor of 1.25 is the most beneficial for all four objective functions;
the first eigenfrequency shows no signs of regression, tension forces are mitigated,
the increase in material use is minimal and the second eigenfrequency shows large
improvements. Increasing the throat height to 500m also shows mostly positive
gains; the only downside is that more material is required, because it takes longer
for the larger radius of the bottom to taper off to the smaller radius of the throat.
The results are shown in Figure 8.6 and 8.7. The increase in throat height has a
minor positive influence on the first and second eigenfrequency and the increase of
the dimensions of the top rings increases the second eigenfrequency even further.
The only downside is that most of the reductions in material use from step 1 have
now been used to optimise the other three objective functions.

1,000
Optimisation step 1
Increase throat height
800 Increase dimensions top rings
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 8.6: Optimisation step 2a in-plane forces N22 (meridional direction)

94
8.5. Optimisation step 2

0.6

Frequency [Hz]
0.4

0.2 Optimisation step 1


Increase throat height
Increase dimensions top rings
0
1 2 3 4 5 6
Mode
Figure 8.7: Optimisation step 2a eigenfrequencies

Since tension forces on the windward side have reduced by such a great deal,
the effect of removing another stiffening ring from the bottom of the chimney is
researched. As Figure 8.8 shows, there will be tension forces in the foundation if
the bottom four rings are removed. However, if the rings are redistributed over the
height, having a spacing of 120m instead of 100m (so that the bottom ring will be
located at 400m, instead of 500m), the tension forces in the foundation disappear.
Figure 8.9 shows that the eigenfrequencies do not suffer if another ring is removed
from the bottom, as long as the spacing between the rings is increased. Lastly,
the results of the second optimisation step are displayed in the Kiviat diagram in
Figure 8.10. All four objective functions now show improvements when compared
to the original base model.

1,000
Optimisation step 2a
Bottom 4 rings removed
800 Rings redistributed
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 8.8: Optimisation step 2b in-plane forces N22 (meridional direction)

95
8. Optimisation of the base model

0.6

Frequency [Hz]
0.4

0.2 Optimisation step 2a


Bottom 4 rings removed
Rings redistributed
0
1 2 3 4 5 6
Mode
Figure 8.9: Optimisation step 2b eigenfrequencies

Reduce
material use

Improve
Reduce
1st eigen-
tension
frequency

Improve
2nd eigen-
frequency

Figure 8.10: Kiviat diagram for the second optimisation step

8.6. Optimisation step 3


The Kiviat diagram from optimisation step 2 shows that two of the objective func-
tions have shown significant improvement, namely the reduction in tension on the
windward side and the improvement of second eigenfrequency. The improvement of
the first eigenfrequency and the reduction of material use have been lagging behind
a bit in this optimisation process. For this reason, the third optimisation step will
focus on improving these two objective functions. The two tertiary parameters are:
modifying the wall thickness profile and the radius at the bottom.
The radius at the bottom, if increased, will improve the first eigenfrequency, but
use more material in the process. Vice versa, if the radius at the bottom is de-
creased, the first eigenfrequency will show regression, but less material will be used.
Since the goal of this optimisation step is to improve both objective functions, this
parameter cannot be used.
The wall thickness profile, on the other hand, is able to improve both of these two
objective functions. By reducing the wall thickness at the top of the chimney, the
first eigenfrequency can be increased and the material use can be decreased. As
a result of this, there will be more tension on the windward side of the chimney.
Fortunately, thanks to previous optimisations, the current configuration has a small
buffer in terms of being able to counteract tension forces in the shell. On top of that,
the wall thickness at the top of the chimney is already relatively thin, therefore it

96
8.6. Optimisation step 3

should not be reduced too much, otherwise it could lead to instability and problems
with constructability aspects such as formwork. For this reason, the wall thickness
will not be decreased by more than 5cm, so that the wall thickness will always
have a minimum value of 20cm, a value commonly found in wall thickness profiles
of NDCTs (Krätzig et al., 2009). For the base model, the wall thickness of the entire
top 50% of the chimney is 0.25cm. This optimisation step will begin by removing
5cm for the top 10% of the chimney, then by removing 5cm for the top 30% of the
chimney and lastly by removing 5cm for the entire top 50% of the chimney. The
results are shown in Figure 8.11 and 8.12.

1,000
Optimisation step 2b
Top 10% reduced to 0.2cm
800 Top 30% reduced to 0.2cm
Top 50% reduced to 0.2cm
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 8.11: Optimisation step 3 in-plane forces N22 (meridional direction)

0.6
Frequency [Hz]

0.4

Optimisation step 2b
0.2 Top 10% reduced to 0.2cm
Top 30% reduced to 0.2cm
Top 50% reduced to 0.2cm
0
1 2 3 4 5 6
Mode
Figure 8.12: Optimisation step 3 eigenfrequencies

As can be seen, if 5cm is removed from more than 30% of the top of the chimney,
tension forces will emerge once again on the windward side of the chimney. Some-
thing which was not expected, however, is the fact that the first eigenfrequency will
also start decreasing again if more than the top 30% of the wall thickness is reduced.
The first eigenfrequency peaks at 0.1867Hz if the top 30% of the wall thickness is
reduced and decreases to 0.1847Hz if the top 50% of the wall thickness is reduced.
For both these reasons, only the top 30% of the wall thickness of the chimney
will be reduced from 0.25cm to 0.2cm. The results of the third optimisation step
are displayed in the Kiviat diagram in Figure 8.13. Compared to previous Kiviat
diagrams, this one shows the most balanced optimised solution when compared to
the base model; all four of the objective functions show significant improvement.

97
8. Optimisation of the base model

Reduce
material use

Improve
Reduce
1st eigen-
tension
frequency

Improve
2nd eigen-
frequency

Figure 8.13: Kiviat diagram for the third optimisation step

8.7. Verification

The multi-objective optimisation process only considered a linear elastic analysis.


Furthermore, the modal analyses only showed the eigenfrequencies, not their ac-
companying mode shapes. To verify if the found optimised configuration actually
performs better than the base model, some additional checks are performed.
First of all, the mode shape belonging to the second eigenfrequency is now no longer
a shell bending mode shape that dominates the entire shell, it is a localised shell
mode shape with five waves around the circumference (Figure 8.14). This implies
that the second eigenfrequency is even more vulnerable for vortex shedding than
before optimisation took place, because the wave number is linearly proportional
to the critical wind speed. This can be clearly seen in the formula below where a
wave number of N = 5 would lead to a significantly lower critical wind speed than
for a wave number of N = 2.

fe · d
vcrit,ovaling = (8.1)
St · N

Figure 8.14: Mode shape belonging to the second eigenfrequency for the optimised model

98
8.8. Conclusions

Furthermore, Figure 8.15 shows that when quasi nonlinear behaviour is considered,
tension forces emerge in the foundation on the windward side of the shell. To
prevent this, an extra stiffening ring has been added at a height of 300m; the spacing
of the rings has been decreased from 120m to approximately 116.67m so that the
seven rings are evenly distributed between 300m and 1000m. Figure 8.15 clearly
shows that the extra ring ensures that the shell is loaded entirely in compression,
even when quasi nonlinear behaviour is considered. This extra ring also ensures
that the second eigenmode is once again a shell bending mode shape that dominates
the entire shell with only two waves around the circumference.

1,000
Optimisation step 3
Quasi nonlinear
800 Quasi nonlinear with extra ring
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 8.15: Verification in-plane forces N22 (meridional direction)

8.8. Conclusions
The interactive multi-objective optimisation method proved to be successful in
optimising the four objective functions. The optimisation steps can be summarised
as follows:
• Increase the aspect ratio of the rings so that the dimensions are 6.75mx1.00m
(WxH)
• Remove the bottom three rings and increase the spacing between the rings
from 100m to ≈ 116.67m (so that the bottom ring will be located at 300m)
• Increase the cross-sectional area of the top 4 (out of the remaining 6 rings)
by a factor of 1.25
• Increase the throat height from 400m to 500m
• Decrease the wall thickness of the top 30% from 0.25m to 0.2m

99
8. Optimisation of the base model

Reduce
material use

Improve
Reduce
1st eigen-
tension
frequency

Improve
2nd eigen-
frequency

Figure 8.16: Kiviat diagram for the final optimisation step

The results of the optimisation steps are shown in the Kiviat diagram in Figure 8.16.
The actual verification of the optimised model will be conducted in Chapter 9.
There the results of the optimised model will be compared to the results of those
analyses for the base model.

100
Analysing the optimised
9
model
In this chapter, the optimised model from Chapter 8 will be thoroughly analysed.
The results will be compared to those of the base model from Chapter 5. The cost
increase of each of the optimisation steps will be discussed and weighed against the
improvement in static and dynamic behaviour. Attention will also be given to the
validity and shortcomings of the chosen optimisation method.

9.1. Comparison method


To compare the optimised model with the base model, judging them solely on
the four key problem areas does not suffice; more thorough analyses have to be
conducted. More specifically, the analyses on the base model from Chapter 5 will
be repeated here for both the base model and the optimised model to determine if
the optimisation step actually led to an improved structure.

9.2. Geometry
The geometry of the optimised model and the base model are shown side by side
in Figure 9.1. Table 9.1 shows the changes in material use for both models. The
optimised model is clearly an improvement in terms of a reduction in material
use. All three aspects show a decrease in material use; especially the rings and
the amount of reinforcement required show positive gains. The concrete shell itself
only shows little improvement due to the increase in throat height which is only
compensated by the reduction in wall thickness of the top 30% of the chimney.

101
9. Analysing the optimised model

(a) Base model (b) Optimised model

Figure 9.1: The optimised model compared to the base model

Base model Optimised model Difference


Element
[tonne] [tonne] [%]
Concrete shell 500720 497060 -0.73%
Concrete rings 88608 68779 -22.38%
Concrete Total 589328 565839 -3.99%
Steel reinforcement 15961 10122 -36.58%

Table 9.1: Comparing the material use of the optimised model and the base model

9.3. Static analysis


Tension on the windward side was one of the four key problem areas of the sensi-
tivity analysis and optimisation process. The results of Chapter 8 already proved
that this problem was solved. However, tension on the windward side is not the
only aspect of the static analysis which needs to be validated. Other checks have
to performed: the deformation should not have increased, nor should compressive
stresses on the leeward side.

102
9.3. Static analysis

9.3.1. In-plane forces


Figure 9.2 and 9.3 show the difference in in-plane forces of the base model and the
optimised model. As was stated before, tension forces on the windward side have
been completely negated by the self-weight of the chimney. The maximum tensile
stress in the reinforcement steel has decreased from 245N/mm2 to 191N/mm2 ,
which is well below the design value of 435N/mm2 . Note that the reinforcement
is only activated in the stiffening rings, the shell itself is loaded entirely in com-
pression. The peak compressive stress has stayed the same at 26.2N/mm2 , still
marginally below the the design compressive strength of 33.33N/mm2 for the cho-
sen concrete quality C50/60. However, even though the peak compressive stress
has stayed the same, in general, the compressive forces on the leeward side of the
chimney have also been reduced, especially in the bottom half. Overall, the forces
are more evenly distributed around the circumference, leading to a more economical
design in terms of dead load required to compensate the wind forces.

2,000 0.0 −2,000.0

0
00.0 −4,000.0
−2,0
−2,000
N 22 [kN/m]

−6,000.0
−4,000
−6,000
1,000
−8,000 −8,000.0 800
600
0 30 400
60 90 120 200
150 180
angle [φ] height [m]

Figure 9.2: In-plane meridional forces of the base model (N22)

0.0

2,000 −2,000.0

0 0
00.
− 2,0
−4,000.0
−2,000
N 22 [kN/m]

−4,000 −6,000.0
−6,000
1,000
−8,000 800
600
0 30 400
60 90 120 200
150 180
angle [φ] height [m]

Figure 9.3: In-plane meridional forces of the optimised model (N22)

103
9. Analysing the optimised model

9.3.2. Deformations
Overall, the deformations have also been reduced as can be seen in Figure 9.4 and
9.5. Displacements on the windward side have decreased greatly while displace-
ments on the leeward side have increased slightly. This implies that the chimney
is acting more like a cantilever beam instead of a shell; if complete cantilever be-
haviour was reached then displacements on the top would be equal over the entire
circumference. A smaller displacement also indicates less fatigue and thus a longer
design working life of the chimney. The maximum displacement is now well be-
low the design criterium of 1/500 times the structural height (Hong Kong Buildings
Department, 2013), measuring in at 1.23m, as opposed to 2.27m for the base model.

2.0

1.5
2.5

2 1.0

1.5
|U| [m]

1 0.5

0.5 1,000
800
0 600
0 30 400
60 90 120 200
150 180
angle [φ] height [m]

Figure 9.4: Displacement vector sum of the base model

2.5
1.0
2

1.5
|U| [m]

0.5 0.5
1

0.5 1,000
800
0 600
0 30 400
60 90 120 200
150 180
angle [φ] height [m]

Figure 9.5: Displacement vector sum of the optimised model

104
9.3. Static analysis

9.3.3. Quasi nonlinear static analysis


While the shell itself is loaded entirely in compression, the rings themselves could
still be under tension. A quasi nonlinear analysis is conducted to see if any of the
ring elements will crack in the optimised model and, if so, what the effect is on
the static response. Figure 9.6 shows the results of this analysis. As can be seen,
the quasi nonlinear response of the optimised model is different from the static
response, but the difference is a lot smaller than for the base model. Furthermore,
even though some ring elements seem to crack which lead to a poorer response, the
self-weight of the chimney is always enough to compensate for the tension forces
induced by the wind. This implies that the base reinforcement in the shell itself
is sufficient, only the rings need additional reinforcement. Pretensioning is not
required in this case, but could be used as a countermeasure against fatigue; less
cracking leads to less deterioration of the material and thus a longer design working
life.

1,000
Base model
Optimised model
800 Base model (quasi nonlinear)
Optimised model (quasi nonlinear)
Height [m]

600

400

200

0
-2,000 2,000
N22 [kN/m]

Figure 9.6: Optimisation step 3 in-plane forces N22 (meridional direction)

9.3.4. Buckling
The buckling analyses in this research are based on eigenvalue buckling, which
means that the results are rather unconversative. However, values can be com-
pared to determine if the optimisation process was actually beneficial or not to the
buckling stability of the chimney. As can be seen in Table 9.2, the buckling safety
factors for the optimised model have increased marginally when compared to the
original base model, implying that the optimisation steps have led to a structure
that is more resistant to buckling.

Buckling mode Base model Optimised model


n λ λ
1 8.40 8.73
2 8.41 11.21
3 10.16 11.29

Table 9.2: Buckling safety factors λ of the base model and the optimised model

105
9. Analysing the optimised model

9.4. Dynamic analysis


The dynamic part of the optimisation process consisted of improving the first and
second eigenfrequency. To see what the effect of these optimisation steps is, a more
thorough dynamic analysis will have to be conducted.

9.4.1. Modal analysis


The modal analysis (Figure 9.7 and Table 9.3) reveals that the first and especially
the second eigenfrequency show improvement. Higher eigenfrequencies actually
show regression, but, considering that the frequencies above 0.3Hz are completely
damped out due to aerodynamic admittance, this can be neglected. The second
eigenfrequency was considered to be the most critical one in terms of dynamic
response, as it was also the most susceptible towards vortex shedding. Considering
that this eigenfrequency is the most improved, it can be concluded that the tuning
of the eigenfrequencies was a success. The mode shapes are shown in Figure 9.8.

0.6
Frequency [Hz]

0.4

0.2
Base model
Optimised model
0
1 2 3 4 5 6
Mode
Figure 9.7: Eigenfrequencies of the base model and the optimised model

Mode number Base model Optimised model


n [Hz] [Hz]
1 0.1654 0.1857
2 0.1928 0.3356
3 0.3404 0.3813
4 0.4775 0.3832
5 0.4883 0.4135
6 0.5438 0.4295

Table 9.3: Eigenfrequencies of the base model and the optimised model

9.4.2. Quasi nonlinear modal analysis


The quasi nonlinear analysis shows that although cracking of the rings does lead
to lowered eigenfrequencies, they are still much higher than for the quasi nonlinear
analysis of the base model (Figure 9.9). As was seen before, the eigenfrequencies
associated with beam bending modes are not influenced by the reduced stiffness
of cracked elements. However, with the base model, the second eigenfrequency
regressed so much in the quasi nonlinear analysis that it became the new first
eigenfrequency. For the optimised model, the first eigenfrequency is always the
beam bending one, regardless of whether or not a linear elastic analysis is conducted

106
9.4. Dynamic analysis

Figure 9.8: Mode shapes belonging to the first 6 eigenfrequencies of the optimised model

or a quasi nonlinear analysis. The second eigenfrequency does show some regression
for the optimised model, but not as much as for the base model.

0.6
Frequency [Hz]

0.4

0.2
Base model
Optimised model
Base model (quasi nonlinear)
Optimised model (quasi nonlinear)
0
1 2 3 4 5 6
Mode
Figure 9.9: Optimisation step 3 eigenfrequencies

9.4.3. Harmonic analysis


The harmonic analysis of the optimised model (Figure 9.10) shows what might
initially be considered to be odd behaviour. The peak of the first eigenfrequency
is actually almost twice as high as it was before and from 0.4Hz onwards the
mechanical admittance function shows a downward slope with no noticeable peaks.
These effects can be explained as follows: first of all, the higher peak suggests that
the dynamic amplification factor for that particular resonance mode is larger than
it was before. The dynamic amplification factor, however, is the result of the actual
amplitude divided by the initial displacement for 0Hz. Considering the fact that
the static deflection (thus the 0Hz displacement) decreased from 2.27m to 1.23m
(almost twice as small), the actual amplitude of that particular resonance mode
stayed constant.
The mechanical admittance function displays the dynamic amplification function
for the displacement of the top node on the windward side. The downward slope
with no peaks suggests that the eigenfrequencies in this range do not lead to mode
shapes with a displacement of the top node; indicating that the mode shapes as-

107
9. Analysing the optimised model

sociated with these eigenfrequencies show instability in the lower regions of the
chimney (Figure 9.8).
The second eigenfrequency for the optimised model does show a much smaller
peak than for the base model. Taking into account the effect of the reduced static
deflection, this suggests that the dynamic response of the second eigenfrequency is
much improved.

Base model
Optimised model

101

χm
100

10−1

0 0.1 0.2 0.3 0.4 0.5 0.6


Frequency [Hz]

Figure 9.10: Mechanical admittance function of the base model and the optimised model

9.4.4. Response spectrum


Similarly to the harmonic analysis, the peak for the first eigenfrequency of the
optimised model is higher than for the first and second eigenfrequency of the base
model. The second eigenfrequency of the optimised model is completely damped
out, however. In order to accurately compare the response spectra of both models,
the gust factor G needs to be calculated in order to determine what the dynamic
component is of the chimney’s response. The method for calculating the gust
factor was described in Subsection 3.6.1. For the base model, the gust factor was
calculated as 1.28 (see Chapter 5). For the optimised model, the same value of
1.28 is found. The gust factor, as was described before, gives a relation between
the initial static deflection ȳ and the dynamic deflection ydyn . Similar values for
the gust factor suggest that the increase in deflection due to dynamic wind actions
is 28% for both the base model and the optimised model. However, considering
the fact that the static deflection for the base model is 2.27m and 1.23m for the
optimised model, the absolute value of the dynamic component of the deflection is
smaller for the optimised model than for the base model.
The cause that the gust factor is the same for both the base model and the op-
timised model can be led back to the aerodynamic admittance function. Because
the structure is so immensely large, most of the air pressure fluctuations due to
gustiness are evened out over the surface of the chimney, thus decreasing the dy-
namic response. The sheer size of the chimney could be seen as the best possible
countermeasure against along-wind oscillations.

108
9.4. Dynamic analysis

Base model
101 Optimised model

100

10−1
f ·Sy (f )
Iz2 ȳ 2
10−2

10−3

10−4

10−5
0 0.1 0.2 0.3 0.4 0.5 0.6
Frequency [Hz]

Figure 9.11: Spectral density of system response for the base model and the optimised model

9.4.5. Vortex shedding


In Chapter 5, it was calculated that the critical wind speed for the first eigenfre-
quency of the base model was
fe · d 0.1654 · 150
vcrit,ovaling = = = 124m/s (9.1)
St · N 0.2

For the optimised model, the critical wind speed for the first eigenfrequency is
fe · d 0.1857 · 150
vcrit,ovaling = = = 139m/s (9.2)
St · N 0.2

But, as was mentioned earlier, the first eigenfrequency is not the most critical one
for vortex shedding, the second one is though. For the second eigenfrequency of
the base model, the critical wind speed was
fe · d 0.1928 · 150
vcrit,ovaling = = = 72.3m/s (9.3)
St · N 0.2 · 2

As noted before, this is awfully close to the maximal wind speed at the top of the
chimney for the reference location. For the optimised model, the second eigenfre-
quency leads to a critical wind speed of
fe · d 0.3356 · 150
vcrit,ovaling = = = 125.9m/s (9.4)
St · N 0.2 · 2

which is much higher than any wind speed which might ever occur in the area.
Therefore, the second eigenfrequency is no longer critical for vortex shedding. Even
in the cracked state, the lowered second eigenfrequency has a critical wind speed
of
fe · d 0.2363 · 150
vcrit,ovaling = = = 88.6m/s (9.5)
St · N 0.2 · 2
Given that the limit is 1.25 · vm = 1.25 · 50 = 62.5m/s; i.e. 1.25 times the charac-
teristic mean wind velocity (averaged over 10 minutes) at the top of the chimney
(EN1991-1-4, 2010). this is also considered to be safe.

109
9. Analysing the optimised model

9.5. Conclusions
In conclusion, the optimisation process has led to an overall improved structure
when compared to the original base model. All facets which were analysed have ei-
ther shown improvement or no noticeable regression. First of all, the more thorough
analyses revealed that indeed all four key problem areas have shown improvement.
The material use has decreased on all fronts, from the shell itself to the rings and
the required reinforcement. The tension on the windward side has decreased, due
to the fact that the forces are more evenly distributed around the circumference,
which also led to a decrease in compression on the leeward side. Furthermore, both
eigenfrequencies have shown improvement. While along-wind resonance does not
pose as great a threat as was initially assumed, due to the effect of aerodynamic
admittance, the results do show that the improved eigenfrequencies led to a smaller
increase in deflection as a result of dynamic wind action. Vortex shedding also no
longer poses a threat as the critical wind speeds have become much larger than
could ever occur at the reference location.

110
10Discussion
In this chapter, the conclusions from each individual chapter are discussed in light
of the results found in the optimisation step.

10.1. Geometry of the chimney


It turns out that the hyperboloid shape at the bottom of the chimney is very effec-
tive in resisting the along-wind cantilever pushover force, even without additional
stiffening rings. The rings are required in the top half of the chimney, however, to
keep the cross-section from ovalising. Material which can be saved from removing
a few rings at the bottom can be used to increase the cross-sectional area of the
top rings, where the most material is required in order to reduce tension on the
windward side and improve the second eigenfrequency. Adding too much material
will make the chimney top-heavy, negatively influencing the first eigenfrequency.
The stiffening rings themselves also benefit from a very large aspect ratio in order
to increase the moment of inertia. The maximum aspect ratio and the effect of
the protruding stiffening rings on the airflow around the chimney still have to be
investigated.

10.2. Actions working on the chimney


The most important parameter of the wind load is the reference hourly mean wind
speed. By constructing the chimney at a location where strong winds do not prevail,
a lot of material can be saved. Furthermore, locations with high altitudes lead to a
lower air density while locations near the equator lead to a smaller Coriolis effect,
both of which reduce the wind load. The problem with higher altitudes is that
this usually indicates a mountainous area which is not flat enough to construct a
solar updraft tower. Lastly, the effect of the wind ribs added on the outside of the
chimney wall is actually much smaller than anticipated. While they do distribute

111
10. Discussion

the forces more evenly around the circumference, they do not help in reducing
tension on the windward side of the chimney. As was stated before, two types
of resonance can be distinguished in these structures; along-wind resonance, and
across-wind resonance. Due to the large size of the chimney and the influence of
the aerodynamic admittance, along-wind resonance does not pose as great a threat
as initially suspected. Across-wind resonance due to vortex shedding can still lead
to failure-inducing resonance, however, and has to be dealt with accordingly. The
second eigenfrequency is the most vulnerable towards vortex shedding and thus has
to be improved.

10.3. Finite element model


The finite element model which was constructed for the purpose of this research
proved to be accurate when compared to results from previous studies. Further-
more, various checks were ran in order to verify the base fixity of the chimney,
and the accuracy of some modelling methods which were applied to reduce the
mesh density, such as using symmetry boundary conditions and locally reducing
the number of elements. However, to accurately model the chimney’s response,
nonlinear modelling techniques have to be used. The optimised model from this
research needs to be verified in terms of nonlinear buckling and nonlinear material
behaviour, before it can actually be constructed.

10.4. Results of the base model


Analyses of the base model revealed four key problem areas which required atten-
tion in the optimisation process. To avoid cracks in the shell and in the rings,
tension on the windward side of the chimney has to be reduced. If the rings crack
then the eigenfrequencies decrease which in turn would further increase the danger
of along- and across-wind oscillations. Additionally, the first and second eigenfre-
quency should be improved in order to mitigate the effects of dynamic wind load-
ing. Furthermore, it appears that the second eigenfrequency has an even greater
dynamic response than the first eigenfrequency. Lastly, the material use of the
chimney should decrease in order to lower the cost of the chimney.

10.5. The need for a design tool


In order to optimise the base model, various sensitivity analyses had to be con-
ducted. Modelling all of the chimneys for every parameter set by hand would have
taken a very long time, so for this reason, a comprehensive design tool was devel-
oped to aid in this research. This tool allows for many more configurations to be
studied and ultimately provided more insight than could ever be amassed by using
traditional modelling methods.

10.6. Sensitivity analysis


The sensitivity analysis revealed which parameters needed to be tweaked in or-
der to improve the four key problem areas. Tension on the windward side can be
reduced by optimising the aspect ratio and the dimensions of the stiffening rings,
and also to some extent by shape optimisations such as increasing the throat height

112
10.7. Optimisation process and verification of the results

and decreasing the throat height. The first eigenfrequency proved to be harder to
improve than the second eigenfrequency. Possible parameter changes are increasing
the throat height and reducing the wall thickness at the top so the chimney will
be less top-heavy. The second eigenfrequency can be optimised by adding more
material to the rings at the top. The required material can be won back by re-
moving rings at the bottom of the chimney, where they are not needed. This, in
combination with modifying the aspect ratio of the rings, can negate most of the
dangers associated with the second eigenfrequency. The other parameters which
were researched either did not lead to any significant improvements or came with
unwanted side-effects such as increased cost.

10.7. Optimisation process and verification of the


results
The process of gradually introducing parameter changes based on their usefulness
proved to be successful in optimising the four key problem areas. While the found
solution might not be the most optimal solution which exists, it does portray a
major step in the right direction of finding an optimal solution and it provides
insight in all of the individual areas which can benefit from optimisation. The
optimisation steps can be summarised as follows:
• Increase the aspect ratio of the rings so that the dimensions are 6.75mx1.00m
(WxH)
• Remove the bottom three rings and increase the spacing between the rings
from 100m to ≈ 116.67m (so that the bottom ring will be located at 300m)
• Increase the cross-sectional area of the top 4 (out of the remaining 6 rings)
by a factor of 1.25
• Increase the throat height from 400m to 500m
• Decrease the wall thickness of the top 30% from 0.25m to 0.2m
More thorough analyses revealed that the optimisation process has indeed led to
an overall improved structure when compared to the original base model. While
along-wind resonance does not pose as great a threat as was initially assumed, due
to the influence of aerodynamic admittance, the results do show that the improved
eigenfrequencies led to a smaller increase in deflection as a result of dynamic wind
action. Vortex shedding also no longer poses a threat as the improved second
eigenfrequency resulted in critical wind speeds which are much larger than could
ever occur at the chosen reference location. Future optimisations should therefore
focus more heavily finding a Pareto optimal solution which favours the second
eigenfrequency instead of the first, assuming that the accompanying mode shapes
stay the same.

113
Conclusions and
11
recommendations
In this chapter, conclusions are drawn about the conducted research and sugges-
tions for future research directions are given.

11.1. Conclusions
The sensitivity analyses and optimisation steps revealed that the key problem areas
can be improved as follows:
• Increasing the moment of inertia of the rings by changing their aspect ratio
ensures that the chimney is fully loaded in compression while also improving
the second eigenfrequency.
• An increase in the throat height further improves the reduction of tension on
the windward side and improves the first eigenfrequency.
• A reduction in wall thickness at the top of the chimney also improves the
first eigenfrequency while simultaneously reducing material use.
• The stiffening rings at the bottom of the chimney serve little to no purpose.
Removing them leads to another reduction in material use while some of
the material gained can be used to improve the second eigenfrequency by
increasing the dimensions of the top rings, consequently reducing tension
even further.
The results of the optimisation process have been displayed in the Kiviat diagram
in Figure 11.1. More thorough analyses of the optimised model revealed that:
• Along-wind resonance does not pose as great a threat as was initially assumed,
due to the effect of aerodynamic admittance. The results do show, however,
that the improved eigenfrequencies led to a smaller increase in deflection as
a result of dynamic wind action.

115
11. Conclusions and recommendations

Reduce
material use

Improve
Reduce
1st eigen-
tension
frequency

Improve
2nd eigen-
frequency

Figure 11.1: Kiviat diagram for the optimised model (blue) and the base model (red)

• Vortex shedding also no longer poses a threat as the improved second eigen-
frequency resulted in critical wind speeds which are much larger than could
ever occur at the chosen reference location.
• Contrary to what was expected, the second eigenfrequency, associated with
a shell bending mode shape, is more critical than the first, associated with a
beam bending mode shape, in terms of dynamic response.
• Geometric and material parameter optimisations can lead to such a reduction
in tension that the minimum reinforcement percentages, according to VGB
guidelines for NDCTs, are deemed sufficient.
• Optimisations of the four key problem areas also led to a reduction in the
static deflection and a minor improvement of the buckling safety factors. As
is often the case, shape optimisations which benefit the eigenfrequencies often
also benefit the buckling stability.

11.2. Recommendations
While the analyses revealed that the chimney structure of the base model was
indeed optimised in this research, more research is required in order to determine
whether or not the found optimised solution is actually feasible. Especially in terms
of nonlinear behaviour and constructability, more knowledge is required to verify
the design. Furthermore, the optimisation process could be expanded as to include
more aspects of the chimney, such as the airflow in the chimney, a more expansive
cost model, and a more thorough implementation of a location model. To gain
more insight in the behaviour of the chimney and to further optimise the design,
the following future research directions are proposed:
• Nonlinear (buckling) analysis: reinforced concrete is a nonlinear mate-
rial. To accurately model such a material, effects such as tension cracking,
tension stiffening, the nonlinear stress-strain relationship in compression, the
yielding of reinforcement and the nonlinear concrete-steel bond have to be
taken into account. The problem with nonlinear analyses is that they take
a long time to set-up and compute; therefore making them unsuited for the
type of research which was conducted in this thesis. An option would be to
incorporate nonlinear analyses sporadically; the optimisation process would

116
11.2. Recommendations

still be mostly based on linear elastic analyses, but every few iterations, the
process would be steered by intermittent nonlinear analyses. Furthermore,
in order to accurately model the buckling behaviour of a chimney, eigenvalue
buckling does not suffice. The results of eigenvalue buckling (also known as
classic Euler buckling) are very unconservative, making them unsuited for ac-
tual design calculations. Nonlinear buckling is much more accurate; it works
in such a way that it gradually increases the load during a static analysis
until a small increase in load will lead to a large deflection in the structure.
In order for this to work, nonlinear material behaviour needs to be taken into
account.
• Multi-objective optimisation: the optimisation process in this research
only focused on a few aspects: the static and dynamic response of the chim-
ney, and the cost of the chimney based on the amount of material required.
However, to truly find an optimal structure, many more aspects have to be
considered such as: the ideal flow behaviour in- and outside the chimney, the
costs associated with constructing the chimney and the materials required,
nonlinear material behaviour. Combining all of these aspects in a multi-
objective optimisation framework could result in a more well thought-out
design. The largest difficulty would lie in creating models for all of these
aspects and giving suitable weights to every single one of these aspects in
order to optimise them. A nontrivial multi-objective optimisation problem
does not have a single solution which optimises every aspect; an optimisation
of one aspect would lead to a regression in another aspect. This effect could
already be seen partly in this research. If one were to consider even more
aspects then the optimisation process could no longer be done manually.
• Directional design for a specified location: the optimised chimney con-
figuration from this research is suitable for any type of location around the
globe, given certain maximum wind speed restrictions. In order to fully op-
timise the chimney for a specific location, directional design could be applied
to further minimise the material use. By using local wind measurements, a
wind rose can be produced which shows a dominant wind direction. More
material can then be allocated to the dominant direction while less material
is needed in the other directions.
• Airflow around the stiffening rings: while some research has already
been conducted in this area (Alberti, 2006), it is still unknown what the
influence of the dimensions and placement of the stiffening rings is on the
behaviour of the airflow around the chimney. A parameter driven study
combined with computational fluid dynamics could provide insight in the
influence of the stiffening rings on the flow behaviour; this knowledge could
then be used to make more well thought-out decisions when configuring a
chimney in SUMAT.
• In-depth research on the constructability aspects: this thesis presents
some background information on the constructability aspects of the chimney
of a solar updraft tower, however, this is only the tip of the iceberg. To be
able to realise a chimney of this height, more research has to be conducted
on aspects such as: minimal wall thickness in terms of stability and concrete
coverage, a design for the foundation, but most importantly the shape and
dimensions of the stiffening rings. In order to maximise the potential of
the rings, the aspect ratio has to be increased as much as possible, yet the
cantilevering action could lead to instability and the stiffening rings could
collapse under their own weight. Existing technical drawings for the stiffening
rings show that the rings could be partly supported on steel beams protruding
from the chimney wall, but more research is required in this area.

117
11. Conclusions and recommendations

• Other materials/construction typologies: this research focused on con-


structing the chimney of a solar updraft tower in traditional reinforced con-
crete. However, other options exist such as construction the chimney as a steel
gridshell, a cable-net structure, or even more innovative typologies such as
floating chimneys (Papageorgiou, 2010) and inflatable chimneys (Putkaradze
et al., 2013). Various new developments in concrete technology could also lead
to more cost-effective designs. High-strength concretes and even UHPFRC
(Ultra High Performance Fiber Reinforced Concrete) are becoming cheaper
by the day, perhaps eventually becoming more cost-effective than traditional
reinforced concrete. Similarly, technological advancements in prefabricated
sections could eventually lead to these types of structures to be constructed
from prefab elements.

118
Bibliography
Alberti, L. (2006). Flow around cylindrical towers: the stabilising role of vertical
ribs.
ANSYS (2013). Help Documentation.
ASCE7-02 (2002). Minimum Design Loads for Buildings and Other Structures.
Technical report.
Bachmann, H., Ammann, W. J., Eisenmann, J., Floegl, I., Hirsch, G. H., Klein,
G. K., Lande, G. J., Mahrenholtz, O., Natke, H. G., Nussbaumer, H., Rainer,
J. H., and Verlag, B. (1995). Vibration Problems in Structures - Practical Guide-
lines.
Backström, T. V. and Gannon, A. J. (2000). Compressible Flow Through Solar
Power Plant Chimneys. 122(August 2000).
Bathe, K.-J. (1982). Finite Element Procedures in Engineering analysis.
Borri, C., Lupi, F., and Niemann, H.-j. (2011). Innovative modelling of dynamic
wind action on Solar Updraft Towers at large heights. In Proceedings of the 8th
International Conference on Structural Dynamics, pages 3567–3574.
Bottenbruch, H. (2010). Prosperity for all in 100 years through the Solar Chimney
Power Plant. SCPT 2010 Proceedings.
Braam, C. R. and Lagendijk, P. (2010). CB2 Constructieleer Gewapend Beton.
Burton, T., Jenkins, N., Sharpe, D., and Bossanyi, E. (2011). Wind energy hand-
book.
Chapelle, D. and Bathe, K.-J. (2003). The Finite Element Analysis of Shells.
Chowdhury, I. and Dasgupta, S. P. (2009). Dynamics of Structure and Foundation
- A Unified Approach - 1. Fundamentals.
Deaves, D. M. and Harris, R. I. (1978). A Mathematical Model of the Structure of
Strong Winds. Number 76 in CIRIA report. CIRIA.
Dowrick, D. J. (2007). Earthquake Risk Reduction.
Dyk, C. V. (2008). A Methodology for Radical Innovation – illustrated by applica-
tion to a radical Civil Engineering structure. PhD thesis.
Dyrbye, C. and Hansen, S. (1997). Wind loads on structures.
EN1991-1-4 (2010). Actions on structures - part 1-4: general actions - wind actions.
Technical Report 2005.
EN1992-1-1 (2011). Design of concrete structures - Part 1-1: General rules and
rules for buildings. Technical Report 2004.
Fluri, T., Pretorius, J., Dyk, C. V., Backström, T. V., Kröger, D., and Zijl, G. V.
(2009). Cost analysis of solar chimney power plants. Solar Energy, 83(2):246–256.
Gould, P. L. (2005). Cooling Tower Structures. In Handbook of Structural Engi-
neering, Second Edition, chapter 27, pages 1–42.

119
Bibliography

Harte, R. (2013). Stability and nonlinear behaviour of RC solar updraft towers.


In Proceedings of the 5th International Conference on Structural Engineering,
Mechanics and Computation, pages 953–958.

Harte, R., Graffmann, M., and Krätzig, W. B. (2013). Optimization of Solar


Updraft Chimneys by Nonlinear Response Analysis. Applied Mechanics and
Materials, 283:25–34.

Harte, R. and Krätzig, W. B. (2011). On structural characteristics of solar chim-


ney power technology. In Proceedings of the 8th International Conference on
Structural Dynamics, pages 3561–3566.

Harte, R. and Van Zijl, G. P. A. G. (2007). Structural stability of concrete wind


turbines and solar chimney towers exposed to dynamic wind action. Journal of
Wind Engineering and Industrial Aerodynamics, 95(9-11):1079–1096.

Helmus, M., Warkus, N., and Lorek, M. (2010). Solar Chimney in Southern Africa.
SCPT 2010 Proceedings.

Holmes, J. (2007). Wind Loading Structures - Second Edition.

Hong Kong Buildings Department (2013). Code of Practice for Structural Use of
Concrete. Technical report.

Inman, D. and Singh, R. (2001). Engineering vibration.

ISO-4354 (2009). Wind actions on structures. Technical report.

Kappos, A. (2002). Dynamic Loading and Design of Structures.

Krätzig, W. B., Harte, R., Andres, M., Eckstein, U., and Wörmann, R. (2013).
Große Schalentragwerke für Energieanlagen: Von Naturzugkühltürmen zu Kami-
nen solarer Aufwindkraftwerke. In Nothnagel, R. and Twelmeier, H., editors,
Baustoff und Konstruktion, pages 53–66. Springer Berlin Heidelberg, Berlin, Hei-
delberg.

Krätzig, W. B., Harte, R., and Montag, U. (2009). From large natural draft cooling
tower shells to chimneys of solar upwind power plants. In Proceedings of the
International Association for Shell and Spatial Structures (IASS) Symposium
2009, number October, pages 2–17.

Krätzig, W. B., Harte, R., and Wörmann, R. (2008). Large shell structures for
power generation technologies. In Proceedings of the 6th International Conference
on Computation of Shell and Spatial Structures, number May, pages 1–4.

Krishna Raju, N. (2008). Prestressed Concrete 4th edition.

Landsberg, H. (1981). The urban climate. Academic Press.

Lauß at, L. (2013). Constructability-aspects for reinforced concrete solar updraft


power plants chimneys : From construction technology and logistics to sched-
ules and costs – data-requirements for economical evaluation. In Proceedings
of the 5th International Conference on Structural Engineering, Mechanics and
Computation, volume 1982, pages 971–976.

Lohaus, L. and Abebe, Y. A. (2010). Concrete Concepts for Solar Chimneys. SCPT
2010 Proceedings.

Miettinen, K. (1999). Nonlinear multiobjective optimization.

Mills, D. (2004). Advances in solar thermal electricity technology. Solar Energy,


76(1-3):19–31.

120
Bibliography

NEN-EN1992-1-1 (2011). Design of concrete structures - Dutch National Annex.


Technical report.
Niemann, H.-j., Lupi, F., Höffer, R., Hubert, W., and Borri, C. (2009). The Solar
Updraft Power Plant: Design and Optimization of the Tower for Wind Effects.
In Proceedings of the 5th European & African Conference on Wind Engineering,
pages 1–12.
Onyango, F. and Ochieng, R. (2006). The potential of solar chimney for application
in rural areas of developing countries. Fuel, 85(17-18):2561–2566.

Papageorgiou, C. D. (2010). Floating Solar Chimney Technology.


Pariseau, W. (2011). Design analysis in rock mechanics.
Putkaradze, V., Vorobieff, P., Mammoli, A., and Fathi, N. (2013). Inflatable free-
standing flexible solar towers. Solar Energy, 98:85–98.

Rousseau, J.-p. (2005). Dynamic evaluation of the solar chimney.


Ruscheweyh, H. (1982). Dynamische Windwirkung an Bauwerken. Bauverlag.
Sawka, M. (2004). Solar Chimney - Untersuchungen zur Strukturintegrität des
Stahlbetonturms.

Schindelin, H. (2002). Entwurf eines 1500 m hohen Turms eines Solar-


Aufwindkraftwerkes - Parameteruntersuchung zur Geometrieoptimierung.
Schlaich, J. and Bergermann, R. (2010). The Solar Updraft Tower : Das Aufwind-
kraftwerk Motivation and Concept. SCPT 2010 Proceedings.

Schlaich, J., Bergermann, R., Schiel, W., and Weinrebe, G. (2005). Design of Com-
mercial Solar Updraft Tower Systems—Utilization of Solar Induced Convective
Flows for Power Generation.
Schlaich, J., Weinrebe, G., and Bergermann, R. (2013). Solar Updraft Towers.
In Richter, C., Lincot, D., and Gueymard, C. A., editors, Solar Energy, pages
658–687. Springer New York, New York, NY.
Strauß, J. (2010). Optimised Erection of Giga Towers. SCPT 2010 Proceedings.
VGB-R-610E (2010). Structural Design of Cooling Towers: VGB Guideline on the
Structural Design, Calculation, Engineering and Construction of Cooling Towers.
VGB PowerTech.
Vrouwenvelder, A. (2014). CIE5145 Random Vibrations Lecture Notes.
Weinrebe, G. and Bergermann, R. (2010). Realization and Costs of Solar Updraft
Towers. SCPT 2010 Proceedings.

121
Literature review
A
A.1. Solar updraft tower

Background
The world is on the verge of an energy crisis. Industrialised nations are polluting
the climate with their use of fossil fuels. These fossil fuels that Western coun-
tries thrive on are running out, yet the total energy demand only keeps growing
(Figure A.1). At the same time, Third World countries live in poverty because
while their population keeps growing, they cannot afford the energy to sustain that
growth.

15000
[MTOE/year]

10000

5000

1970 1980 1990 2000 2010 2020 2030 2040 2050


Oil Gas Coal Nuclear Renewable

Figure A.1: World energy production (dataset from www.2052.info)

123
A. Literature review

Because of this, the problems related to poverty, pollution and the depletion of fossil
fuels can only be solved by using renewable energy sources on a global scale. The
poorer countries on the southern hemisphere, especially African countries, have an
advantage here over the richer countries which are mostly situated in the northern
hemisphere; the annual solar irradiation in these poorer countries is much higher
which makes them ideal candidates for housing large scale affordable solar power
plants.
These solar power plants come with several advantages for poorer countries, for
example:
• They provide new jobs for the construction of these large scale projects.
• They can benefit from their own product.
• Since electric energy can be transported over very large distances without
noteworthy losses they can export their energy product to the richer countries.
One of the proposed concepts for a solar power plant is the solar updraft tower.
Two important elements of a solar updraft tower are a glass collector roof with
a diameter of up to 7km and a concrete chimney with proposed heights of up to
1500m. The construction materials for these two elements can be manufactured
from sand and stone which is directly available on site (Schlaich and Bergermann,
2010).

Working principle
The solar updraft tower consists of three essential elements, some of which have
already been applied in other forms of power generation but now come together in
a unique way. The three elements can be seen in Figure A.2.

Figure A.2: Working principle (Krätzig et al., 2009)

Solar air collector A large circular transparent roof functions as the collector
area. The collector area is open at the perimeter to allow cold air to enter.
The roof allows solar radiation to heat up the air underneath the roof and
trap it there, functioning like a greenhouse. The power output of a solar
updraft tower is dependent on the size of this collector area and as such,
air collectors with a diameter of up to 7km have been proposed. The canopy
material of the air collector is likely to be toughened glass, because of induced
vortices near the chimney which would cause instability in weaker materials
such as plastic sheets (Mills, 2004).

124
A.1. Solar updraft tower

Chimney/Tower A chimney stands at the centre of the air collector, forming


an airtight connection. Since hot air is lighter than cold air, the hot air
that is trapped in the air collector will want to rise up in the chimney. This
causes suction in the chimney, thus drawing in even more hot air from the
air collector and causing cold air to enter the air collector from the outer
perimeter. The taller the chimney, the higher the pressure difference due to
the stack effect; chimneys with a height of up to 1500m have been proposed.
Wind turbines The airflow caused by the air collector and chimney drives pressure-
staged turbines at the base of the chimney. The mechanical energy in the
turbines is then converted into electrical energy by conventional generators.
The efficiency of a solar updraft tower can be expressed in the following form:
Pelectrical
ηg = (A.1)
G · Acollector

This gives the ratio between the amount of electricity generated [W ] and the
amount of solar radiation G[W/m2 ] which hits the collector area Acollector [m2 ].
The energy conversion rate of this process from solar energy to electrical energy
is around 3%. Losses occur due to the fact that there is a temperature drop of
around 10 degrees Celsius from the bottom to the top in a 1000m tall chimney.
Large amounts of air have to move up which causes gravitational losses. The fact
that the air which exits the chimney at the top is warmer than the ambient air also
results in thermal energy loss. Lastly, as the air enters the collector area, it expands
due to the solar radiation with little increase in pressure. Most of the solar energy
is lost in the simple expansion of air before it even reaches the turbines (Onyango
and Ochieng, 2006).

Key aspects
Even though the energy conversion ratio might not be as high as that of other
energy plants which harness solar energy, solar updraft towers have some very
special features of their own (Schlaich et al., 2005):
• Solar updraft towers operate 24 hours a day on solar power alone. During the
day, radiation energy is stored into the soil underneath the air collector and
released into the collector at night. Airflow through the chimney is caused
by temperature differences between the air collector and the ambient air, so
with lower ambient temperatures at night the solar updraft tower will still
generate electricity. The efficiency at night can be improved by placing water
bags on the ground in the air collector, thus increasing thermal storage.
• Solar updraft towers work with direct and diffuse radiation which is beneficial
for tropical countries with frequently overcast skies.
• A solar updraft tower does not need cooling water. This is a key benefit for
countries of high solar radiation which also have to deal with a low supply of
water.
• Due to the fact that a solar updraft tower has very few moving parts, it
requires little to no maintenance.
• The main resources needed for construction are sand and stone which are
readily available on most suitable sites
The solar updraft tower comes with some disadvantages as well though:
• The solar air collector requires a very large area (up to 7km in diameter) to
be considered efficient. Because of this, solar updraft towers can only be built

125
A. Literature review

where land prices are low such as desert and savannah areas. Note that the
area must also be very flat to be considered.
• Investment– and construction costs are very high, causing investors to shy
away from a feasible but unproven technology, at least on a large scale.
• Areas with frequent sand storms or earthquakes should also be avoided, due
to high maintenance and construction costs.

Pilot project
In 1979, the German Ministry for Research and Technology granted the German
firm Schlaich Bergermann & Partners the sum of 3.5 million DM (around e1.8
million) for a feasibility study of a solar updraft tower. Instead, the firm decided
to invest that money in a prototype plant in Manzanares, Spain (Figure A.3).

Figure A.3: Prototype in Manzanares, Spain (Schlaich et al., 2013)

The prototype with a height of 200m and an output of 20kW, allowed them to verify
their calculations in a real life situation (Figure A.4). The original plan was to build
the prototype in 1980, perform measurements in the following two years and then
dismantle it out of safety reasons. The skin of the prototype plant was comprised
of corrugated metal sheet, being held upright by stay cables, positioned at four
different heights in three different directions. The grant given by the German
Ministry did not cover corrosion protection of the cables, so failure of the chimney
was imminent. Yet it was not until 1990 that the cables finally gave in and the
structure collapsed, causing the prototype to be decommissioned.

Figure A.4: Measured output vs calculated output (Schlaich et al., 2013)

The prototype allowed for the firm to test a variety of different materials for the
transparent or translucent covering of the solar air collector. Several types of
plastic sheet material and glass were selected to find out which one was the most
durable and cost effective in the long run. The glass panels were deemed the most
efficient since they were able to resist heavy storms for many years and proved to
be self-cleaning due to rainfall.

126
A.1. Solar updraft tower

Location
The main requirement for the site of a solar updraft tower is a high annual solar
irradiation. Without it, the power plant would never come close to realising its full
potential. There are, however, several other factors which influence the suitability
of a location for the construction of a SUT. First of all, the location has to be rather
flat if a collector with a diameter of 7km were to be constructed. Locations which
are prone to hurricanes, typhoons, cyclones and other forms of natural disasters
are also rejected. Earthquakes also belong in this list, but as has been said before,
the chimney has a natural safety margin against seismic loading due to its low first
eigenfrequency. Still, seismic regions have to be avoided since earthquakes could
activate the higher modes of a solar updraft tower and cause unwanted stresses.
Other factors also come into play, such as the political stability of the region,
the availability of skilled workers and construction materials, but these factors are
rather hard to quantify.
An initial study has been carried out to determine suitable locations for a solar
updraft tower. Four basic parameters (annual solar irradiation, mean wind speed,
elevation, and seismic activity) are weighed against each other to determine the
suitability of a location (Figure A.5).

(a) Annual solar irradiation (b) Mean wind speed

(c) Elevation (d) Seismic activity

Figure A.5: Global map data for various parameters

First of all, only locations with an annual solar irradiation of 2200kW/m2 are
considered. Of the locations that remain, those with a negative elevation are filtered
out, since building a SUT in the ocean is not an option. The same goes for locations
with a high elevation; mountainous areas are rarely flat enough to build a SUT,
never mind the fact that it would be a logistical nightmare to transport all of the
labourers and materials up there. Lastly, areas with a high mean wind speed and
areas which are prone to earthquakes are also filtered out. The result is a so-called
’best location map’, which can be seen in Figure A.6. Since this map neglects a lot

127
A. Literature review

of very important factors, it is not really a map for the ’best’ location of a solar
updraft tower. Yet it is a first step in the right direction in filtering out areas which
are considered unsuitable and a rough estimation in determining which areas are
suitable.

Figure A.6: Best location map

As can be seen, large parts of the continent of Africa are suitable for the deployment
of a SUT. Especially Niger in Central Africa and Botswana in the south score well.
Other locations that perform well are Western Asia (Yemen, Oman, Saudi Arabia)
and Australia.

Financial feasibility
Perhaps the largest obstacle standing in the way of the solar updraft tower are the
large costs associated with it. A small scale power plant will not be economically
viable while a large one will require huge amounts of capital which scares off in-
vestors since the technology has not been proven on a large scale. However, there
are several other motives to consider for investing in solar updraft towers, namely:
• Creation of jobs in third world countries
• Maximising the potential of third world countries, allowing them to rely less
on foreign exchange
• Saving other resources such as water and non-renewable sources (oil, coal,
gas, etc.) (Bergermann, Weinrebe)
Furthermore, once the first plant has been built and its efficiency has been proven,
subsequent plants can benefit from economy of scale; reduction of costs by building
a great number of them.
The large costs associated with solar updraft towers can be split into the following
three components:
• Total investment cost

128
A.1. Solar updraft tower

• Operation and maintenance cost

• Financing conditions which can be summarised in an average interest rate

The total investment cost can then be split into the percentages seen in Figure A.7.
A 200MW coal-fired power plant (CFPP) costs around e300 million to construct
while a solar updraft tower with a similar energy production will cost around e750
million (Bottenbruch, 2010; Weinrebe and Bergermann, 2010). Regardless of this
large difference in initial cost, when looking at it in the long run it becomes clear
that solar updraft towers have the upper hand over traditional methods of energy
generation such as CFPPs. This advantage can best be illustrated by comparing
the LEC (levelised electricity cost) of SUTs and CFPPs.

Engineering and coordination: 7%


Chimney: 25%
Turbines: 19%

Cost of a solar
updraft tower

Collector: 49%
Figure A.7: Total investment costs for a solar updraft tower

Figure A.8 shows such an example, where both plants have been completely paid
off after 20 years. The LEC of the solar updraft tower is high at first due to the
annuity payments. Fossil fuels are becoming increasingly scarce and thus their
prices are rising, causing the LEC of the SUT and the CFPP to be equal after 20
years. Here a shift occurs, both plants have been paid for but the CFPP still relies
on fossil fuels which have become the largest contributor to its LEC. The solar
updraft tower on the other hand needs no external resources and only operation
and maintenance costs determine its LEC from here on out. On top of that, coal-
fired power plants have a lower life expectancy than solar updraft towers and have
to be replaced every 30 years, increasing the LEC once again due to investment
costs associated with the construction of this second plant. Here in lies the strength
and the weakness of SUTs, they are financially beneficial in the long run, but most
investors want to see returns in the short run. Furthermore, CFPPs emit 0.95kg
CO2 /kWh while a SUT does not emit CO2 at all (disregarding the fact that the
construction process of a SUT will be driven by fossil fuels). This also lowers the
LEC of electricity produced by solar updraft towers since they do not need to pay
for carbon credits; a carbon credit is a generic term which is basically a permit for
companies to emit carbon dioxide (Fluri et al., 2009).

The main parameters that define the output of a solar updraft tower are the col-
lector diameter and the chimney height. For both these parameters investigations
have been done to see what the optimal value is in terms of cost per kW installed.
In Figure A.9 it can be seen that the optimal diameter is around 7000-7500m
and the optimal height is somewhere in the range of 1000-1100m (Weinrebe and
Bergermann, 2010).

129
A. Literature review

Figure A.8: LEC costs for a solar updraft tower vs a coal-fired power plant (Schlaich et al., 2005)

(a) Optimal collector diameter (b) Optimal chimney height

Figure A.9: Optimal dimensions for a solar updraft tower (Weinrebe and Bergermann, 2010)

A.2. Structural principles


Over the years, several structural systems for solar updraft towers have been pro-
posed, ranging from free-standing reinforced concrete shells to guyed tubes made
of corrugated sheeting, cladded cable-net structures and even more unconventional
suggestions such floating chimneys (Papageorgiou, 2010) and inflatable chimneys
(Putkaradze et al., 2013). Most of these structural systems have been tried and
tested in the construction of natural draught cooling towers (NDCTs). NDCTs are
the closest relative of solar updraft towers, structurally speaking, and can therefore
provide a good starting point for the structural design.

Natural draft cooling towers (NDCTs)


NDCTs utilize the same principle as solar updraft tower, buoyancy via a tall chim-
ney to move large quantities of air upwards. Their goal for doing so might be
different, but the mechanics are the same. The highest cooling towers in existence
top out at around 200m, for example the one in Niederaussem, while the proposed

130
A.2. Structural principles

height for solar updraft towers is in the range of 1000-1500m. Suffice it to say, a
mere extrapolation of the structure of a NDCT will not be feasible. Structures of
such a height come with a whole slew of new issues and differ greatly in their static,
as well as their dynamic behaviour.
In most cases, NDCTs are constructed as reinforced concrete hyperboloid shell
structures. Some exceptions exist such as the cable-net NDCT which was con-
structed in Schmehausen (Figure A.10).

(a) RC cooling tower (b) Cable-net cooling tower in Schmehausen

Figure A.10: Examples of NDCTs (en.wikipedia.org)

From NDCTs to SUTs


Solar updraft towers are closely related to the shape- and strength-optimised design
of NDCTs. In Figure A.11 several pre-designs for different SUT sizes are shown.
The smallest one, with a height of 500m, is just on the edge of profitability, when
being compared to parabolic solar reflector plants. The 1000m tall chimney served
as a reference design for the base model used in this research. Combined with a
collector diameter of 6km it will produce a peak power of 200MW.

Figure A.11: From the largest NDCT to the chimneys of solar updraft towers (Krätzig et al.,
2009)

The actions working on a SUT are very similar to those working on a NDCT.
The largest difference between the two is that SUT have several stiffening rings

131
A. Literature review

distributed evenly over the height. Compared to a typical cantilever beam in


bending, a tall, unstiffened shell structure behaves very differently (Figure A.12).
The thinner the shell structure becomes, the harder it gets for the shell to evenly
distribute the meridional forces around the circumference. The stiffening rings
assist the shell structure in evenly distributing these forces, leaning towards an
optimal cosine-distribution.

Figure A.12: Circumferential distribution of meridional stresses in a beam and a shell (Schlaich
et al., 2013)

These stiffening rings also improve the buckling behaviour of the shell structure.
For an unstiffened shell the instability modes would dominate the entire shell (Fig-
ure A.13). Ring stiffeners prevent these global modes from occurring by localising
the buckling modes (Krätzig et al., 2009). These stiffening rings not only improve
the buckling behaviour of the chimney, but also the dynamic behaviour. SUT chim-
neys have very low eigenfrequencies, such that strong winds can lead to resonance
of the shell. Seismic actions operate in a much higher frequency spectrum, thus
not forming a great danger.

Figure A.13: First three instability modes of a 1000m tall chimney under (D+We +Wi ) (Krätzig
et al., 2009)

132
A.3. Construction process

A.3. Construction process


The construction process of a solar updraft tower is a daunting challenge on its own.
Never before has a structure of this height been erected. This section attempts to
define the main challenges arising in constructing such a tall structure.

Concrete temperature
Solar updraft towers can only be profitable when placed in hot climates. In these
hot climates, special considerations have to be given to the temperature of the
concrete. Because of the heat, the hydration process of the cement is so fast that
it could lead to premature stiffening of the freshly poured concrete. Large concrete
structures generate a lot of heat due to the hydration process and, because the
structure is so massive, this generated heat has trouble dissipating to the outside
environment. If the rate of the cooling down process is too high, thermal cracks can
form. In addition, because the structure is constantly exposed to high temperatures
and possibly direct sunlight, the surface of the concrete can dry up too quickly,
increasing the chance of cracking. All of these effects can, in time, decrease the
characteristic strength and durability of the applied concrete.
Advances in concrete technology have led to measures which can be taken to min-
imise the aforementioned risks. Some of these measures are:
• The usage of concrete with a very low cement content in combination with
ground granulated blast furnace slag and pulverised fly ash
• The usage of cements which have a lower heat generating capacity than reg-
ular cement (for instance Natural pozzolan)
• The usage of superplasticisers and a large quantity of retarding admixtures
The codes specify that the maximum allowable temperature in freshly poured con-
crete should not exceed 30◦ C. In regions with a high ambient temperature, the
average temperature of the concrete mixture can be as high as 49◦ C. There exists
the option to artificially lower the concrete temperature by using ice and ice water
cooling technologies. Figure A.14 shows the amount of water and ice required to
cool 1m3 of concrete. When 170L of chilled water is used, the temperature de-
creases from 49◦ C to 37◦ C, which is still higher than the 30◦ C which the codes
prescribe. However, when 100L of ice water is combined with 70kg of flake ice, the
concrete temperature can be decreased to 29◦ C.
Ice and ice water cooling technologies prove to be very effective in reducing the
temperature of concrete in hot climates. Lastly, a few other solutions to keep the
temperature of the concrete down are:
• Keeping the machinery, pumping lines and aggregates away from direct sun-
light by using some form of shading
• Refrain from casting concrete during the daytime

Concrete pumping
Another challenge lies in pumping the concrete to the great heights required for
solar updraft towers. The only somewhat comparable project is the Burj Khal-
ifa, a skyscraper in Dubai with a height of 828m (Figure A.15). There they used
three stationary pumps, each applying 150 bars of pressure, to transport concrete
to a record height of 606m. The solar updraft tower, with its height of 1000m,
would require a tremendous amount of pressure to transport the concrete all the

133
A. Literature review

Figure A.14: Concrete cooling with ice and water (www.kti-plersch.com)

way to the top. Another solution would be to use staged pumping, meaning that
aside from the stationary pumps at the base of the chimney, there would be sev-
eral pumps positioned at intermediate levels to continue pumping concrete to the
desired height.
Aside from pressure, another important aspect in pumping concrete is the choice of
pumping lines. Larger pipes are usually the preferred choice for mass concreting,
yet the downside is that the concrete spends a longer time in the pipes, leading
to stiffening of the concrete and a reduced effectiveness of additives such as super-
plasticisers. Narrow pipes are the preferred choice for transporting concrete over
long distances or heights, since the concrete is able to move at a much faster pace
through the pipe. The issue with these more narrow pipes is that there is an in-
creased risk of blockage and a shorter lifespan for the pipes themselves due to wall
friction against the concrete.
A small example for different pipe diameters is shown in Table A.1. In it, three
pipe diameters are compared with regard to the time it takes to pump concrete all
the way to the top of a 1000m tall solar updraft tower. Two cases are described to
show the fluctuations in pumping operation. In case one the pump functions sat-
isfactorily, with a pumping rate of 12(m3 /h). In case two the pump is functioning
in the worst possible conditions, having a pumping rate of 8(m3 /h).

Pipe diameter Concrete in Q Duration Q Duration


Option
[mm] pipe [m3 ] [m3 /h] [min] [m3 /h] [min]
1 100 7,9 12 40 8 59
2 125 12,3 12 62 8 92
3 150 17,7 12 89 8 133

Table A.1: Pumping duration (Lohaus and Abebe, 2010)

Concrete cannot stay in the pipe for more than an hour in such hot climates,
otherwise premature stiffening will occur, possibly damaging the entire set-up.

134
A.3. Construction process

Figure A.15: Burj Khalifa (www.skyscraperpage.com)

Given this fact, option 1 and 2 both seem like feasible solutions under satisfactory
operating conditions. Unfortunately, once conditions deteriorate, option 2 is no
longer able to deliver the concrete within the hour, potentially having very costly
consequences. Therefore, only option 1 is a viable solution for a 1000m tall chimney.

Another problem in pumping concrete to such great heights is the fact that the con-
crete mixture starts to lose its stability because of the enormous pumping pressure.
This effect, also known as forced bleeding, causes the separation of the grout or ce-
ment paste from the aggregates. This unsaturated suspension can lead to blockage
in the pipe once the coarser aggregates start to clump together. This problem is
only exacerbated by the friction in the pipe, leading to the grout being pulled out
of the mixture. High-strength concrete has a lower water-cement ratio than regular
concrete mixtures and is therefore even more susceptible towards forced bleeding.

In conclusion, for a good pumpability, the concrete needs to have an optimal level of
cohesion. Too little cohesion and the mixture starts to separate, too much cohesion
and the concrete cannot move smoothly through the pipe due to friction (Lohaus
and Abebe, 2010).

An alternative to concrete pumping would be the installation of high-speed eleva-


tors which could put huge concrete masses into position in time (Strauß, 2010).
However, the elevator systems which are currently on the market, such as the Ali-
mak Scando 650 FC-S (Figure A.16) can only transport 1m3 of concrete per shift.
Assuming a transportation height of 500m, one shift will take approximately 15
minutes; 2.5 minutes for loading, 5 minutes to travel to the top, 2.5 minutes for
casting and 5 minutes to travel back down again. This implies that 4m3 can be
transported per hour, which is very little compared to the pumping installation at
the Burj Khalifa which was designed to pump 30m3 per hour (Lauß at, 2013).

135
A. Literature review

Figure A.16: Alimak Scando 650 FC-S (www.alimakhek.com)

Formwork
Already stated was that only traditional concrete construction methods are a viable
option for the construction of a solar updraft tower. Currently, pre-fabricated
sections cannot be produced in the sizes which would be required for such a large
structure. Hence, the only realistic solution is the usage of in-situ concrete in
formwork (Helmus et al., 2010).
Two types of formwork are common in the construction of tall structures, namely
climbing formwork and gliding formwork.

Climbing formwork
Climbing formwork, as its name implies, is a type of formwork which climbs with
the building while it increases in height. A distinction is made between self-climbing
formwork and crane-climbing formwork. In the case of self-climbing formwork, the
structure is able to climb on its own using hydraulic jacks. In the case of the latter,
the formwork around the structure is elevated with the help of one or more cranes,
once the concrete has hardened. Crane-climbing formwork is not very suitable for
extremely tall constructions, since at such heights the cranes have trouble operating
even under minor winds.
The climbing formwork itself is firmly anchored in the concrete. The platform
also doubles as a working space for the construction workers; an attached plat-
form hanging underneath the climbing formwork allows for construction workers
to perform concrete finishing work.

Gliding formwork
Gliding formwork is similar to climbing formwork, with the main difference being
that the pouring process is continuous instead of periodic. Hydraulic jacks elevate
the structure at a constant pace, ranging from 15cm/h to 30cm/h. The speed
is mostly dependent on the time needed for concreting and placing reinforcement

136
A.4. Comparison of vertical wind-profiles

bars. If the formwork moves too fast, the concrete at the bottom will not have
hardened enough to be able to stand on its own. Also, it is very important that the
concreting process is not interrupted, otherwise large problems will arise regarding
the strength and stability of the resulting structure. Similarly to the climbing
formwork, suspended scaffolds below the working platform enables construction
workers to treat the concrete surface. Altitude is not a limiting factor for gliding
formwork; the platform is always firmly anchored to the structure itself, being able
to withstand wind speeds of 164km/h.
A large advantage of gliding formwork is that it is not limited to orthogonal building
forms, the geometry of the cross-section can be changed during the gliding process.
A disadvantage is that when using gliding formwork the shell thickness needs to be
increased slightly (Harte and Krätzig, 2011).
In conclusion, gliding formwork is better suited for the construction of a solar
updraft tower, due to the fact that the geometry of the formwork can be changed
during the casting process (Sawka, 2004).

Pretensioning
The chimney of a solar updraft tower is characterised by a very thin shell thickness,
held together by stiffening rings distributed evenly over the height. If wind causes
large tension forces in the shell, these stiffening rings might require pretensioning.
The most common method for pretensioning cylindrical structures is a traction
machine, also called a ’merry go round’, developed by the Preload Engineering
Company. The traction machine, hanging from a trolley, runs along the top of the
cylindrical wall. The pretensioning wire is run through a steel plate with a small
hole in it while it is wound on the structure to achieve the desired tension. As a
safety measure, the wires are anchored to the structure using clips protruding from
the wall to ensure that, if a wire were to snap, the winding will not detach. The
winding speed is around 4.5m/s, thanks to technological advancements over the
years (Krishna Raju, 2008).

A.4. Comparison of vertical wind-profiles


Most wind design profiles are only valid for altitudes up to 200m since there is
a lack of meteorological data for wind speeds above 300m (Harte and Van Zijl,
2007). In this appendix, several wind profiles will be compared to one another to
find out which one is most suitable for modelling the wind actions on a 1000m tall
chimney. The wind profiles which will be compared are the log profile, the corrected
log profile, and the power law profile. All three have their own advantages and
disadvantages and all three are still used commonly nowadays. Eurocode 1991,
for instance, uses the log profile, while ASCE-07 uses the power law profile. The
international standard for wind actions, ISO-4354, uses the corrected log profile.
First of all, a quick overview of the individual wind profiles will be given, after
which they will be compared to one another using similar input parameters.

Log profile
In the boundary layer, two length scales play an important role, namely the surface
roughness and the boundary layer height. The surface roughness dominates the
wind speed near the ground while the boundary layer height determines the wind
speeds in the top half of the boundary layer, near the free flow regime. The

137
A. Literature review

logarithmic profile, described in this paragraph, only takes the surface roughness
into account. It accurately portrays the wind profile up to 300m above ground for
high wind speeds (> 25m/s).
 
u∗ z
Vz = ln (A.2)
k z0
where k is von Kármán’s constant (0.4) and u∗ is the so-called friction velocity
which is defined by r
τ0
u∗ = (A.3)
ρ
In this formula, ρ is the density of air and τ0 is the shear stress at the ground
surface. For storm profiles, the friction velocity is usually in the range of 1 − 2m/s.
As mentioned earlier, EC 1 uses the logarithmic profile for the mean wind velocity
up to 200m above ground (Dyrbye and Hansen, 1997).

Corrected log profile


The log profile is not very accurate at altitudes above 300m. Deaves and Harris
(1978) developed a modified version of this log profile which incorporates correction
terms depicting wind effects which are prevalent at high altitudes, such as the
Coriolis effect. The corrected log profile accurately fits data retrieved from wind
measures and also covers surface roughness changes. It is given by
(      2  3  4 )
u∗ z z z 4 z 1 z
Vz = ln + 5.75 − 1.88 − + (A.4)
k z0 zG zG 3 zG 4 zG

where the last three terms only come into play at altitudes above 300m. In it, zg
is the gradient height and fc is the Coriolis parameter, which is dependent on the
latitude of the site.
The Coriolis effect is a force which is caused by the rotation of the Earth. At the
equator the Coriolis effect is zero, but in the Northern hemisphere, wind forces
deflect to the right while in the Southern hemisphere, wind forces deflect to the
left. This is illustrated in Figure A.17.

Power law profile


While the log profile has a valid theoretical basis, the conditions on which it is based
are rarely ever met in real life. Furthermore, since the logarithm of a negative value
does not exist, the profile cannot be used for negative altitudes. It is also relatively
hard to integrate a logarithmic function. For these reasons, some wind engineers
prefer to use the power law profile.
The power law profile has no real theoretical basis, but it is fairly accurate when
compared to the log profile. It is also easily integrated over the height which makes
it a lot simpler to determine the bending moment at the base of a structure. The
power law profile can be written as
 α
z
Vz = Vref (A.5)
zref

where the α-exponent will change with the roughness of the terrain. If the rough-
ness length z0 is known, the α-exponent can be determined with (Holmes, 2007)
 
1
α= (A.6)
ln(zref /z0 )

138
A.4. Comparison of vertical wind-profiles

Figure A.17: Coriolis effect (www.charlesburrows.com)

For several terrains, the roughness length z0 and α-exponent are given in Table A.2.

Terrain z0 α
Open country, flat 0.05 0.16
Suburban settlement 0.30 0.28
Inner cities 3.00 0.40

Table A.2: α-values for different types of terrain (Landsberg, 1981; Ruscheweyh, 1982)

Comparison

The three profiles which were previously described are now being compared against
one another. To compare the two log profiles with the power law profile, matching
parameters have to be chosen. The chosen roughness length z0 is 0.05m. The
α-exponent for this roughness length with a reference height zref of 10m is 0.16.
The friction velocity u∗ is set at 2.0m/s and the latitude of the comparison site is
15◦ North.

The three profiles are plotted in Figure A.18. As can be seen, they all show similar
results near the ground, but as the altitude increases they start to deviate from one
another. The power law profile overestimates the wind speed from an altitude of
100m and higher. The log profile and the corrected log profile show similar results
up to an altitude of 200m, but after that the Coriolis effect causes the two to grow
apart. Given these results, and the fact that the chimney described in this research
is 1000m tall, only the corrected log profile suffices. The other two profiles, while
admittedly easier to use, show large deviations at high altitudes which makes them
too inaccurate to be used for the modelling of wind actions on a solar updraft
tower.

139
A. Literature review

1,000
Log
Corrected log
900
Power law

800

700

600
Height [m]

500

400

300

200

100

0
1 1.5 2 2.5
Normalised wind speed (Vz /Vref )

Figure A.18: Comparison of wind profiles

A.5. Damping
General information
The resonance in a structure, caused by an external excitation, eventually dies
out. This is a result of a phenomenon known as damping. The resonance ampli-
tude decreases due to various types of energy dissipation which are present in the
structural system (Figure A.19).
The dynamic differential equation for a damped, forced harmonic oscillator (Fig-
ure A.20) is (Kappos, 2002)

M Ü + C U̇ + KU = F (t) (A.7)

where C is the damping ratio and U̇ is the velocity vector. This damping ratio
is the result of various damping components. The three most common forms of
damping are: (Bachmann et al., 1995; Rousseau, 2005)
Viscous damping dissipates energy proportional to the velocity of the structure,
usually as a result of the drag force of a surrounding gas.
Coulomb damping is a result of friction forces between dry surfaces, and falls
in the category of material damping.
Hysteretic damping also known as structural damping is caused by internal
friction forces in the material.

140
A.5. Damping

Figure A.19: Free damped vibration as a result of an initial displacement (Bachmann et al.,
1995)

Figure A.20: SDOF oscillator (Kappos, 2002)

To model these effects, several options are available. The most commonly known
method is Rayleigh damping (Figure A.21). In this method, two parameters
are specified, the α- and the β-parameter. α-damping, also known as mass-
damping, represents the viscous component. It is very effective in damping low-
frequency system-level oscillations. β-damping, also known as stiffness-damping,
represents the hysteretic component, and it is very effective in damping high-
frequency component-level oscillations (Chowdhury and Dasgupta, 2009).
The equation to determine the damping matrix [C] using Rayleigh damping is

[C] = α[M ] + β[K] (A.8)

Where [M ] is the mass matrix and [K] is the stiffness matrix.


The simplest method of applying damping to a structural system is to use a con-
stant damping ratio, specified as a decimal value, which represents the ratio of the
actual damping to the critical damping
c
ζ= (A.9)
cc

where c is the actual damping and cc is the critical damping.


Since this research focuses mostly on tuning the eigenfrequencies of the chimney
of a solar updraft tower, the choice is made to use the simplified method of apply-
ing a constant damping ratio to the structure. Rayleigh damping more accurately

141
A. Literature review

Figure A.21: Rayleigh damping

portrays the damping of higher eigenfrequencies, but because the first few eigenfre-
quencies will be the dominant ones in regards to dynamic wind loading, this effect
is neglected in this research.
Furthermore, since damping is the result of many different effects working together,
it is very hard to find a proper calculation method for the amount of damping
present in a structure. A commonly used approach is to measure the damping
in existing structures and use those values as a guideline for the damping in new
structures.

Logdec
As was stated before, several mechanisms exist which are the cause of damping in a
structure. When measuring damping in a structure, the amount present is usually
expressed as an equivalent viscous damping. Damping in existing structures can be
measured with the logdec (logarithmic decrement). The logdec is the natural log
of the ratio of two peaks in an underdamped structural system. In Figure A.22, a
damped vibration is shown in which two successive peaks, x1 and x2 are specified.

Figure A.22: Damped vibration (Bachmann et al., 1995)

To calculate the logdec of two successive two peaks, the following formula can be
used, with x1 = x(t), x2 = x(t + nT ), and n as the number of intermediate peaks
(n = 1 if the peaks are immediately following one another) (Inman and Singh,

142
A.5. Damping

2001)

1 x(t)
δ= ln (A.10)
n x(t + nT )

Once the logdec is known, the constant damping ratio can be calculated using
1
ζ=q (A.11)
2π 2

1+ δ

If the constant damping ratio is zero, then the structure will resonate to infinity, if
it is one then the structure will be critically damped. Also note how the frequency
of the peak shifts slightly depending on the amount of damping in a structure
(Figure A.23). The eigenfrequencies of a damped structure show a minor devia-
tion from an identical, but undamped structure. However, for very low values of
damping, this frequency shift can be neglected.

Figure A.23: Forced response of a system for different damping ratios (Bachmann et al., 1995)

In the Eurocode, δ-values are given for several structural systems (Table A.3). Since
the chimney in this research is built using reinforced concrete, a logdec value of 0.10
is used (note that the RC chimneys described in this table are a lot more slender
than the chimney described in this research and therefore the aforementioned logdec
value is used). A logdec of 0.10 implies ± 1.5% of critical damping.

Structural system δ
Reinforced concrete buildings 0.10
Steel buildings 0.05
Mixed structures (concrete + steel) 0.08
Reinforced concrete towers and chimneys 0.03

Table A.3: δ-values for different structural systems (EN1991-1-4, 2010)

143
Sensitivity analysis
B
Almost all of the pages in this chapter have the lay-out as depicted below. Axes
and tick labels are only shown in the samples below to improve the readability of
the pages; the scale of the plot axes is always the same.

1,000 5,000
N22 [kN/m]
Height [m]

800
600 0
400 −5,000
200
0 −10,000
-2,000 2,000 0 60 120 180
N22 [kN/m] angle [φ]
(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.1: Sample plot in-plane forces N22

0.8
Frequency [Hz]

0.6
0.4
0.2
0
1 2 3 4 5 6
Mode
Figure B.2: Sample plot eigenfrequencies

145
B. Sensitivity analysis

B.1. Geometric parameters


The following sets of geometric parameters are analysed and scored in terms of
static response, modal response and the costs based on the amount of material
required.
Throat height 300m, 350m, 400m, 450m, 500m
Radius throat 65m, 70m, 75m, 80m, 85m
Radius bottom 120m, 130m, 140m, 150m, 160m
Angle of inclination Min (9.2◦ ), Low (11.4◦ ), Med (13.6◦ ), High (15.8◦ ), Max
(18.0◦ )
Wall thickness (overall)
• Thinner (0.5m; 0.35m; 0.25m; 0.15m; 0.15m; 0.15m)
• Thin (0.55m; 0.4m; 0.3m; 0.2m; 0.2m; 0.2m)
• Default (0.6m; 0.45m; 0.35m; 0.25m; 0.25m; 0.25m)
• Thick (0.65m; 0.5m; 0.4m; 0.3m; 0.3m; 0.3m)
• Thicker (0.7m; 0.55m; 0.45m; 0.35m; 0.35m; 0.35m)
Wall thickness (bottom)
• Thinner (0.5m; 0.35m; 0.25m; 0.25m; 0.25m; 0.25m)
• Thin (0.55m; 0.4m; 0.3m; 0.25m; 0.25m; 0.25m)
• Default (0.6m; 0.45m; 0.35m; 0.25m; 0.25m; 0.25m)
• Thick (0.65m; 0.5m; 0.4m; 0.25m; 0.25m; 0.25m)
• Thicker (0.7m; 0.55m; 0.45m; 0.25m; 0.25m; 0.25m)
Wall thickness (top)
• Thinner (0.6m; 0.45m; 0.35m; 0.15m; 0.15m; 0.15m)
• Thin (0.6m; 0.45m; 0.35m; 0.2m; 0.2m; 0.2m)
• Default (0.6m; 0.45m; 0.35m; 0.25m; 0.25m; 0.25m)
• Thick (0.6m; 0.45m; 0.35m; 0.3m; 0.3m; 0.3m)
• Thicker (0.6m; 0.45m; 0.35m; 0.35m; 0.35m; 0.35m)
Number of rings (evenly spaced)
• 10 (100m; 200m; 300m; 400m; 500m; 600m; 700m; 800m; 900m; 1000m)
• 9 (111m; 222m; 333m; 444m; 556m; 667m; 778m; 889m; 1000m)
• 8 (125m; 250m; 375m; 500m; 625m; 750m; 875m; 1000m)
• 7 (143m; 286m; 429m; 571m; 714m; 857m; 1000m)
• 6 (167m; 333m; 500m; 667m; 833m; 1000m)

146
B.1. Geometric parameters

Number of rings (spaced 100m from the top)


• 10 (100m; 200m; 300m; 400m; 500m; 600m; 700m; 800m; 900m; 1000m)
• 9 (200m; 300m; 400m; 500m; 600m; 700m; 800m; 900m; 1000m)
• 8 (300m; 400m; 500m; 600m; 700m; 800m; 900m; 1000m)
• 7 (400m; 500m; 600m; 700m; 800m; 900m; 1000m)
• 6 (500m; 600m; 700m; 800m; 900m; 1000m)
Number of rings (spaced 100m from the bottom)
• 10 (100m; 200m; 300m; 400m; 500m; 600m; 700m; 800m; 900m; 1000m)
• 9 (100m; 200m; 300m; 400m; 500m; 600m; 700m; 800m; 900m)
• 8 (100m; 200m; 300m; 400m; 500m; 600m; 700m; 800m)
• 7 (100m; 200m; 300m; 400m; 500m; 600m; 700m)
• 6 (100m; 200m; 300m; 400m; 500m; 600m)
Aspect ratio rings 3.40x2.00m, 3.85x1.75m, 4.50x1.50m, 5.40x1.25m, 6.75x1.00m

147
B. Sensitivity analysis

Throat height

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.3: Throat height in-plane forces N22

300m 350m 400m 450m 500m

Figure B.4: Throat height eigenfrequencies

300m 350m 400m 450m 500m


Reducing tension -6 -3 0 3 6
Reducing material use 2 1 0 -1 -2
Improving 1st eigenfrequency -3 -1 0 1 3
Improving 2nd eigenfrequency -1 0 0 0 1

Table B.1: Throat height sensitivity analysis

148
B.1. Geometric parameters

Radius throat

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.5: Throat radius in-plane forces N22

65m 70m 75m 80m 85m

Figure B.6: Throat radius eigenfrequencies

65m 70m 75m 80m 85m


Reducing tension 2 1 0 -1 -2
Reducing material use 5 2 0 -2 -5
Improving 1st eigenfrequency -3 -1 0 2 -1
Improving 2nd eigenfrequency 6 3 0 -2 -2

Table B.2: Throat radius sensitivity analysis

149
B. Sensitivity analysis

Radius bottom

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.7: Bottom radius in-plane forces N22

120m 130m 140m 150m 160m

Figure B.8: Bottom radius eigenfrequencies

120m 130m 140m 150m 160m


Reducing tension 6 3 0 -2 -4
Reducing material use 4 2 0 -2 -4
Improving 1st eigenfrequency -2 -1 0 1 2
Improving 2nd eigenfrequency 0 0 0 0 0

Table B.3: Bottom radius sensitivity analysis

150
B.1. Geometric parameters

Angle of inclination

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.9: Angle of inclination in-plane forces N22

Min Low Med High Max

Figure B.10: Angle of inclination eigenfrequencies

Min Low Med High Max


Reducing tension -8 -3 0 1 3
Reducing material use 1 1 0 -1 -1
Improving 1st eigenfrequency 3 1 0 0 -1
Improving 2nd eigenfrequency 1 0 0 0 -1

Table B.4: Angle of inclination sensitivity analysis

151
B. Sensitivity analysis

Wall thickness (overall)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.11: Wall thickness (overall) in-plane forces N22

Thinner Thin Default Thick Thicker

Figure B.12: Wall thickness (overall) eigenfrequencies

Thinner Thin Default Thick Thicker


Reducing tension -10 -5 0 5 10
Reducing material use 10 7 0 -7 -10
Improving 1st eigenfrequency 1 0 0 0 -1
Improving 2nd eigenfrequency 3 2 0 -2 -3

Table B.5: Wall thickness (overall) sensitivity analysis

152
B.1. Geometric parameters

Wall thickness (bottom)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.13: Wall thickness (bottom) in-plane forces N22

Thinner Thin Default Thick Thicker

Figure B.14: Wall thickness (bottom) eigenfrequencies

Thinner Thin Default Thick Thicker


Reducing tension 1 0 0 0 -1
Reducing material use 8 4 0 -4 -8
Improving 1st eigenfrequency -3 -2 0 2 3
Improving 2nd eigenfrequency -1 0 0 0 1

Table B.6: Wall thickness (bottom) sensitivity analysis

153
B. Sensitivity analysis

Wall thickness (top)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.15: Wall thickness (top) in-plane forces N22

Thinner Thin Default Thick Thicker

Figure B.16: Wall thickness (top) eigenfrequencies

Thinner Thin Default Thick Thicker


Reducing tension -8 -4 0 4 8
Reducing material use 7 4 0 -4 -7
Improving 1st eigenfrequency 4 2 0 -1 -3
Improving 2nd eigenfrequency 3 1 0 -1 -2

Table B.7: Wall thickness (top) sensitivity analysis

154
B.1. Geometric parameters

Number of rings (evenly spaced)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.17: Number of rings (evenly spaced) in-plane forces N22

10 9 8 7 6

Figure B.18: Number of rings (evenly spaced) eigenfrequencies

10 9 8 7 6
Reducing tension 0 -3 -7 -10 -10
Reducing material use 0 1 2 3 4
Improving 1st eigenfrequency 0 0 0 1 1
Improving 2nd eigenfrequency 0 -1 -1 -2 -2

Table B.8: Number of rings (evenly spaced) sensitivity analysis

155
B. Sensitivity analysis

Number of rings (spaced 100m from the top)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.19: Number of rings (spaced 100m from the top) in-plane forces N22

10 9 8 7 6

Figure B.20: Number of rings (spaced 100m from the top) eigenfrequencies

10 9 8 7 6
Reducing tension 0 0 -1 -2 -5
Reducing material use 0 1 2 3 4
Improving 1st eigenfrequency 0 0 0 0 0
Improving 2nd eigenfrequency 0 0 0 0 0

Table B.9: Number of rings (spaced 100m from the top) sensitivity analysis

156
B.1. Geometric parameters

Number of rings (spaced 100m from the bottom)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.21: Number of rings (spaced 100m from the bottom) in-plane forces N22

10 9 8 7 6

Figure B.22: Number of rings (spaced 100m from the top) eigenfrequencies

10 9 8 7 6
Reducing tension 0 -6 -10 -10 -10
Reducing material use 0 1 2 3 4
Improving 1st eigenfrequency 0 0 -7 -10 -10
Improving 2nd eigenfrequency 0 -3 -2 -6 -10

Table B.10: Number of rings (spaced 100m from the bottom) sensitivity analysis

157
B. Sensitivity analysis

Aspect ratio rings

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.23: Aspect ratio rings in-plane forces N22

3.40x2.00m 3.85x1.75m 4.50x1.50m 5.40x1.25m 6.75x1.00m

Figure B.24: Aspect ratio rings eigenfrequencies

3.4x2m 3.85x1.75m 4.5x1.5m 5.4x1.25m 6.75x1m


Reducing tension -10 -5 0 5 10
Reducing material use 0 0 0 0 0
Improving 1st eigenfrequency -2 -1 0 0 0
Improving 2nd eigenfrequency -3 -2 0 4 8

Table B.11: Aspect ratio rings sensitivity analysis

158
B.2. Material parameters

B.2. Material parameters


The following sets of material parameters are analysed and scored in terms of static
response, modal response and the Material uses based on the amount of material
required.
Density of concrete 1700kg/m3 , 1900kg/m3 , 2100kg/m3 , 2300kg/m3 , 2500kg/m3
Concrete quality C50/60, C60/75, C70/85, C80/95, C90/105
Ring stiffness (material properties, all 10 rings) 1x, 1.05x, 1.1x, 1.15x, 1.2x
Ring stiffness (material properties, top 5 rings) 1x, 1.05x, 1.1x, 1.15x, 1.2x
Ring stiffness (cross-sectional area, all 10 rings) 1x, 1.25x, 1.5x, 1.75x, 2x
Ring stiffness (cross-sectional area, top 5 rings) 1x, 1.25x, 1.5x, 1.75x, 2x

159
B. Sensitivity analysis

Density of concrete

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.25: Density of concrete in-plane forces N22

1700kg/m3 1900kg/m3 2100kg/m3 2300kg/m3 2500kg/m3

Figure B.26: Density of concrete eigenfrequencies

1700 1900 2100 2300 2500


Reducing tension -10 -10 -7 -4 0
Reducing material use 0 0 0 0 0
Improving 1st eigenfrequency 7 4 3 2 0
Improving 2nd eigenfrequency 4 3 2 1 0

Table B.12: Density of concrete sensitivity analysis

160
B.2. Material parameters

Concrete quality

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.27: Concrete quality in-plane forces N22

C50/60 C60/75 C70/85 C80/95 C90/105

Figure B.28: Concrete quality eigenfrequencies

C50/60 C60/75 C70/85 C80/95 C90/105


Reducing tension 0 0 0 0 0
Reducing material use 0 0 0 0 0
Improving 1st eigenfrequency 0 1 2 2 3
Improving 2nd eigenfrequency 0 0 1 1 2

Table B.13: Concrete quality sensitivity analysis

161
B. Sensitivity analysis

Ring stiffness (material properties, all 10 rings)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.29: Ring stiffness (material properties, all 10 rings) in-plane forces N22

1x 1.05x 1.1x 1.15x 1.2x

Figure B.30: Ring stiffness (material properties, all 10 rings) eigenfrequencies

1x 1.05x 1.1x 1.15x 1.2x


Reducing tension 0 1 1 2 2
Reducing material use 0 0 0 0 0
Improving 1st eigenfrequency 0 0 0 0 0
Improving 2nd eigenfrequency 0 0 1 1 2

Table B.14: Ring stiffness (material properties, all 10 rings) sensitivity analysis

162
B.2. Material parameters

Ring stiffness (material properties, top 5 rings)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.31: Ring stiffness (material properties, top 5 rings) in-plane forces N22

1x 1.05x 1.1x 1.15x 1.2x

Figure B.32: Ring stiffness (material properties, top 5 rings) eigenfrequencies

1x 1.05x 1.1x 1.15x 1.2x


Reducing tension 0 1 1 1 2
Reducing material use 0 0 0 0 0
Improving 1st eigenfrequency 0 0 0 0 0
Improving 2nd eigenfrequency 0 0 1 1 2

Table B.15: Ring stiffness (material properties, top 5 rings) sensitivity analysis

163
B. Sensitivity analysis

Ring stiffness (cross-sectional area, all 10 rings)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.33: Ring stiffness (cross-sectional area, all 10 rings) in-plane forces N22

1x 1.25x 1.5x 1.75x 2x

Figure B.34: Ring stiffness (cross-sectional area, all 10 rings) eigenfrequencies

1x 1.25x 1.5x 1.75x 2x


Reducing tension 0 8 10 10 10
Reducing material use 0 -2 -5 -8 -10
Improving 1st eigenfrequency 0 -1 -1 -2 -2
Improving 2nd eigenfrequency 0 2 4 6 9

Table B.16: Ring stiffness (cross-sectional area, all 10 rings) sensitivity analysis

164
B.2. Material parameters

Ring stiffness (cross-sectional area, top 5 rings)

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.35: Ring stiffness (cross-sectional area, top 5 rings) in-plane forces N22

1x 1.25x 1.5x 1.75x 2x

Figure B.36: Ring stiffness (cross-sectional area, top 5 rings) eigenfrequencies

1x 1.25x 1.5x 1.75x 2x


Reducing tension 0 6 10 10 10
Reducing material use 0 -2 -3 -4 -5
Improving 1st eigenfrequency 0 0 -1 -1 -2
Improving 2nd eigenfrequency 0 2 4 6 9

Table B.17: Ring stiffness (cross-sectional area, top 5 rings) sensitivity analysis

165
B. Sensitivity analysis

B.3. Loading parameters


The following sets of loading parameters are analysed in terms of their static re-
sponse. Note that the modal response is based on the unloaded state, hence any
load variations will not influence the eigenfrequencies of the structure.
Wind speed 25m/s, 30m/s, 35m/s, 40m/s, 45m/s
Altitude AMSL 300m, 500m, 1000m, 1500m, 2000m
Latitude 0◦ , 10◦ , 15◦ , 20◦ , 30◦
Return period 25 years, 50 years, 100 years, 200 years, 500 years
Pressure distribution curve K1,0, K1,1, K1,2, K1,3, K1,5

166
B.3. Loading parameters

Wind speed

25m/s
30m/s
35m/s
40m/s
45m/s

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.37: Wind speed in-plane forces N22

Altitude AMSL

300m
500m
1000m
1500m
2000m

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.38: Altitude in-plane forces N22

167
B. Sensitivity analysis

Latitude

0◦
10◦
15◦
20◦
30◦

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.39: Latitude in-plane forces N22

Return period

25 years
50 years
100 years
200 years
500 years

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.40: Return period in-plane forces N22

168
B.3. Loading parameters

Pressure distribution curve

K1,0
K1,1
K1,2
K1,3
K1,5

(a) Meridional direction (b) Circumferential direction (H = 350m)

Figure B.41: Pressure distribution curve in-plane forces N22

169
B. Sensitivity analysis

B.4. Damping parameters


The following set of damping parameters is analysed in terms of its harmonic
response.
Constant damping ratio 0.5%, 1%, 1.5%, 2%, 2.5%

170
B.4. Damping parameters

Constant damping ratio

0.6
0.5

2.5%
0.4

2%
Frequency (Hz)

1.5%
0.3

1%
0.2

0.5%
0.1
0
102

101

100

10−1

Dynamic Amplification Factor

Figure B.42: Constant damping ratio harmonic analysis

171
SUMAT user manual
C
Introduction
SUMAT is a design tool created specifically to aid in this research. It allows for
the user to analyse several chimney configurations simultaneously and compare
their results. The tool was created in Python 3, using the PyQT framework for
the graphical shell. It requires a Python 3 distribution which includes NumPy,
matplotlib with the Basemap add-on, and PyQT 4. ANSYS needs to be installed
on the user’s computer to use the full feature set of SUMAT. On the flash drive
which accompanies this report, a portable distribution of Python 3 can be found
which includes all of the aforementioned libraries and a shortcut to automatically
run SUMAT.

Figure C.1: SUMAT logo, modelled after the 4th mode shape

173
C. SUMAT user manual

Geometry tab

Figure C.2: SUMAT user interface, displaying the geometry tab

In the ’geometry’ tab, the user can specify the geometric parameters of up to
5 different chimneys. The geometric parameters are divided into the following
headings:
Shape Define the overall shape of the chimney.
Stiffeners Define the amount and the spacing of the stiffening rings.
Wall thickness Define the wall thickness profile at 6 discrete positions. Interme-
diate values are interpolated
Mesh size Choose the amount of elements used in the meshing process. Also
allows the user to choose between a full modelling of the chimney or a half
modelling with symmetry boundary conditions.

174
Materials tab

Figure C.3: SUMAT user interface, displaying the materials tab

In the ’materials’ tab, the user can specify the material parameters for the five
different sets. The material parameters are divided into the following headings:
Base materials Define the density, Young’s modulus and tensile strength of the
concrete and reinforcement steel.
Ring stiffness Allows the user to increase the stiffness of the stiffening rings, ei-
ther by increasing the cross-sectional area of the rings or the Young’s modulus
of the material. Lastly, the user can choose if the stiffness of all of the rings
has to be increased, or just the stiffness of a select few.
Reinforcement Choose the reinforcement percentages in the meridional and cir-
cumferential direction. These percentages are not used to define the actual
stiffness of the material, but to be able to calculate the total amount of
reinforcement steel required. The user can also choose the amount of rein-
forcement to be added in tension zones. This value is also used to calculate
the maximum steel stress for the static analysis.
Soil Normally the chimney is modelled as being completely fixed. By modelling
the soil, the chimney will be supported on springs with stiffnesses derived
from the soil stiffness.

175
C. SUMAT user manual

Loads and Options tab

Figure C.4: SUMAT user interface, displaying the loads and options tab

In the ’loads and options’ tab, the user can specify the wind and dead loads for the
five different sets. The loads and options parameters are divided into the following
headings:
Loads Choose the mean reference wind speed and its averaging interval, the lati-
tude and altitude to calculate the influence of the Coriolis effect and the effect
of the decrease in air density. Different circumferential pressure distributions
are also made available to the user. Lastly, the partial safety factors can be
specified per set.
Static By using an interactive approach, nonlinear material behaviour can be
approximated by lowering the stiffness of tension zones.
Modal The user can choose the number of eigenmodes to be extracted. Like
the static analysis, the modal analysis allows for an iterative calculation to
determine the influence of cracking on the eigenfrequencies. The parameters
for the Von Kármán PSD can also be specified. Lastly, the Strouhal number,
which is used to calculate the influence of vortex shedding, can be changed.
Harmonic Choose the number of calculation steps and the range of frequencies
which should be analysed. A constant value for the structural damping can
also be set.

176
Preview tab

Figure C.5: SUMAT user interface, displaying the preview tab

In the ’preview’ tab, the shape, the placement of the stiffening rings and the wall
thickness profile for the five sets are displayed so that the user can quickly verify
if no mistakes were made in this process.

177
C. SUMAT user manual

Static tab

Figure C.6: SUMAT user interface, displaying the static tab

The ’static’ tab displays the following results:


Meridional direction Plots the in-plane meridional force in the meridional di-
rection. The user can choose between separate plots for wind and dead loads,
or a cumulative plot which displays the combined load case.
Circumferential direction Plots the in-plane meridional force in the circumfer-
ential direction. The user can choose at which height the cross-section should
be made (in increments of 5%).
Other Numeric values are given for the total weight of the concrete, the weight of
the rings, the amount of reinforcement steel applied, the maximum compres-
sive stresses in the concrete, the maximum tensile stresses in the steel, and
the displacement at the top.

178
Cracking tab

Figure C.7: SUMAT user interface, displaying the cracking tab

The ’cracking’ tab displays in which elements the tensile stress has exceeded the
tensile strength of concrete.

179
C. SUMAT user manual

Buckling tab

Figure C.8: SUMAT user interface, displaying the buckling tab

The ’buckling’ tab can be used to iteratively determine for which load factor the
eigenvalue is one, which can be reverse engineered into a buckling safety factor.

180
Eigenfrequencies tab

Figure C.9: SUMAT user interface, displaying the eigenfrequencies tab

The ’eigenfrequencies’ tab displays the following results from the modal analysis:
Modal analysis Plots the eigenfrequencies for the five sets.
Von Kármán PSD Plots the eigenfrequencies on the Von Kármán PSD.
Other Numeric values are given for the eigenfrequencies.

181
C. SUMAT user manual

Mode Shapes tab

Figure C.10: SUMAT user interface, displaying the mode shapes tab

The ’mode shapes’ tab displays the mode shapes of the chimneys. By calculating
the displacement vectors of the top nodes it also tries to determine if the mode is
a beam bending mode or a shell bending mode. Based on this result, the critical
speed for vortex shedding can be calculated.

182
Harmonic tab

Figure C.11: SUMAT user interface, displaying the harmonic analysis tab

The ’harmonic’ tab displays the results of the harmonic analysis. Results can be
plotted as an absolute amplitude, as a dynamic amplification factor (by dividing
the displacement for any given frequency by the displacement for 0Hz), and as a
response spectrum (using the Von Kármán PSD). From the response spectrum, the
gust factor can be calculated by computing the variance of the dynamic response
using the composite trapezoidal rule.

183
C. SUMAT user manual

Location tab

Figure C.12: SUMAT user interface, displaying the location tab

The ’location’ tab is able to display various types of map date, such as annual solar
irradiation, seismic activity, mean wind speed, and elevation. It helps the user in
finding a suitable location for constructing a solar updraft tower. The map data
can be combined by weighting the data. The result is a so-called ’best location’
map which shows ideal locations for a solar updraft tower. If a location is chosen,
the latitude and elevation values can be used in the ’loads and options’ tab.

184
Settings tab

Figure C.13: SUMAT user interface, displaying the settings tab

Several options can be adjusted in the ’settings’ tab:


Settings Locate the installation directory of ANSYS and specify which directory
should be used as the working directory (useful when running SUMAT from
a slow flash drive).
Legend labels Change the legend labels for the five sets.
Pro Mode Allows SUMAT to use pre-existing data (useful to show the workings
of SUMAT on computers without ANSYS installed).

185
List of Figures
D
1 Pre-existing design for a 1000m tall SUT (courtesy of Krätzig &
Partner, Bochum) (Niemann et al., 2009) . . . . . . . . . . . . . . . iv

1.1 Working principle (Krätzig et al., 2009) . . . . . . . . . . . . . . . . 1


1.2 Pre-existing design for a 1000m tall SUT (courtesy of Krätzig &
Partner, Bochum) (Niemann et al., 2009) . . . . . . . . . . . . . . . 2
1.3 Dynamic excitation frequencies of structures by wind and earth-
quake (Holmes, 2007) . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Flowchart depicting the research methodology . . . . . . . . . . . . . 5

2.1 Deriving the geometric parameters of the chimney . . . . . . . . . . 10


2.2 Example of directional design . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Hyperbola . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Technical drawings for the ring stiffeners (Harte and Krätzig, 2011;
Niemann et al., 2009) . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.1 Mean wind profile with turbulence component (Bachmann et al., 1995) 18
3.2 Simplified representation of the roughness length (Dyrbye and Hansen,
1997) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3 Roughness length z0 for several terrain types (ISO-4354, 2009) . . . 21
3.4 Pressure distribution for chimney with ribs (left) and with a smooth
surface (right) (Niemann et al., 2009) . . . . . . . . . . . . . . . . . 22
3.5 Pressure distributions for the design of NDCTs (Gould, 2005) . . . . 22
3.6 Surface roughness and type of pressure distribution (Gould, 2005) . 23
3.7 Density of air as a function of altitude . . . . . . . . . . . . . . . . . 24
3.8 Free stream velocity pressure . . . . . . . . . . . . . . . . . . . . . . 26
3.9 Classification of dynamic effects from wind (Bachmann et al., 1995) 26
3.10 Spectral densities and magnification functions (Bachmann et al., 1995) 27
3.11 Von Kármán PSD . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.12 Aerodynamic admittance function . . . . . . . . . . . . . . . . . . . 29
3.13 Wind force spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.14 Composite trapezoidal rule (en.wikipedia.org) . . . . . . . . . . . . . 30

186
List of Figures

3.15 Vortex shedding in transcritical regime (Sawka, 2004) . . . . . . . . 31


3.16 Across-wind vibration amplitude for reduced wind speed ur = u/dfe
(Bachmann et al., 1995) . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.17 Ovalising shapes for cylindrical shell structures (Ruscheweyh, 1982) . 32

4.1 The eight steps of creating a FE model: defining keypoints, draw-


ing splines, creating areas, setting mesh divisions, meshing, setting
boundary conditions, adding stiffening rings, applying loads . . . . . 36
4.2 SHELL281 element (ANSYS, 2013) . . . . . . . . . . . . . . . . . . . 37
4.3 SHELL281 elements share their nodes with BEAM189 elements . . . 37
4.4 Mesh without refinement (left) and with refinement (right) . . . . . 38
4.5 Full chimney (left) and half chimney with symmetry boundary con-
ditions (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.6 Bi-linear stress-strain diagram for concrete (EN1992-1-1, 2011) . . . 40
4.7 Idealised (A) and design (B) stress-strain diagrams for reinforcing
steel (EN1992-1-1, 2011) . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.8 Different degrees of mesh densities, ranging from very coarse (left)
to very fine (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.9 Comparing the first 25 eigenfrequencies . . . . . . . . . . . . . . . . 42
4.10 Different degrees of mesh densities, ranging from very coarse (left)
to very fine (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.11 Comparing the first 25 eigenfrequencies for the ’+’-variants . . . . . 43
4.12 Comparing the calculated eigenfrequencies with those found in lit-
erature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.13 Layered model for a reinforced concrete shell (Gould, 2005) . . . . . 45
4.14 Nonlinear material properties of reinforced concrete (Gould, 2005) . 45
4.15 Linear elastic analysis compared to quasi nonlinear analysis (1, 2
and 3 iterations) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.16 Fictional Young’s modulus for different reinforcement percentages . . 47
4.17 Adjusting variable loads to find an eigenvalue of 1.0 (ANSYS, 2013) 48
4.18 Equivalent nodal allocations for a) 2-D elements, b) 3-D elements,
c) triangular 3-D elements (ANSYS, 2013) . . . . . . . . . . . . . . . 49
4.19 FE model with the foundation springs . . . . . . . . . . . . . . . . . 49
4.20 Comparing the in-plane meridional forces for several soil stiffnesses
with the fixed foundation . . . . . . . . . . . . . . . . . . . . . . . . 50
4.21 Comparing the eigenfrequencies for several soil stiffnesses with the
fixed foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

5.1 In-plane forces coordinate system . . . . . . . . . . . . . . . . . . . . 56


5.2 In-plane forces in the base model . . . . . . . . . . . . . . . . . . . . 56
5.3 In-plane meridional forces (N22) . . . . . . . . . . . . . . . . . . . . 57
5.4 Deformed chimney under under D+1.5W (magnification factor: 50x) 57
5.5 Displacement vector sum . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.6 Linear (red) vs quasi nonlinear (blue) in-plane forces N22 (merid-
ional direction) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.7 Linear (red) vs quasi nonlinear (blue) in-plane forces N22 (circum-
ferential direction) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.8 Cracked elements under the default load case . . . . . . . . . . . . . 60
5.9 Mode shapes belonging to the first 3 buckling eigenvalues of the base
model under D+1.5W . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.10 Mode shapes belonging to the first 6 eigenfrequencies of the base
model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.11 First 6 eigenfrequencies of the base model . . . . . . . . . . . . . . . 62
5.12 Comparing the eigenfrequencies of an uncracked chimney and a
cracked chimney . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.13 Mechanical admittance function of the base model . . . . . . . . . . 63

187
List of Figures

5.14 Spectral density of system response for the base model . . . . . . . . 64

6.1 Flowchart depicting the workings of SUMAT . . . . . . . . . . . . . 67


6.2 SUMAT user interface, displaying the geometry tab . . . . . . . . . 69
6.3 SUMAT user interface, displaying the static analysis tab . . . . . . . 71
6.4 Comparison between actual stress distribution and in-plane force
simplification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.5 Flowchart depicting the reinforcement process . . . . . . . . . . . . . 73

7.1 Example parameter sensitivity plot . . . . . . . . . . . . . . . . . . . 76


7.2 Scale example sensitivity plot . . . . . . . . . . . . . . . . . . . . . . 76
7.3 Throat height sensitivity plot . . . . . . . . . . . . . . . . . . . . . . 77
7.4 Radius throat sensitivity plot . . . . . . . . . . . . . . . . . . . . . . 77
7.5 Radius bottom sensitivity plot . . . . . . . . . . . . . . . . . . . . . 78
7.6 Angle of inclination (minimum in red, maximum in blue) . . . . . . 78
7.7 Angle of inclination sensitivity plot . . . . . . . . . . . . . . . . . . . 78
7.8 Wall thickness (overall) sensitivity plot . . . . . . . . . . . . . . . . . 79
7.9 Wall thickness (bottom) sensitivity plot . . . . . . . . . . . . . . . . 79
7.10 Wall thickness (top) sensitivity plot . . . . . . . . . . . . . . . . . . 79
7.11 Number of rings (evenly spaced) sensitivity plot . . . . . . . . . . . . 80
7.12 Number of rings (spaced 100m from the top) sensitivity plot . . . . . 80
7.13 Number of rings (spaced 100m from the bottom) sensitivity plot . . 80
7.14 Aspect ratio rings sensitivity plot . . . . . . . . . . . . . . . . . . . . 81
7.15 Density of concrete sensitivity plot . . . . . . . . . . . . . . . . . . . 81
7.16 Concrete quality sensitivity plot . . . . . . . . . . . . . . . . . . . . . 82
7.17 Ring stiffness (material properties, all 10 rings) sensitivity plot . . . 82
7.18 Ring stiffness (material properties, top 5 rings) sensitivity plot . . . 82
7.19 Ring stiffness (cross-sectional area, all 10 rings) sensitivity plot . . . 83
7.20 Ring stiffness (cross-sectional area, top 5 rings) sensitivity plot . . . 83
7.21 Reduce tension sensitivity plot . . . . . . . . . . . . . . . . . . . . . 85
7.22 Reduce material use sensitivity plot . . . . . . . . . . . . . . . . . . 86
7.23 Improve first eigenfrequency sensitivity plot . . . . . . . . . . . . . . 86
7.24 Improve second eigenfrequency sensitivity plot . . . . . . . . . . . . 87

8.1 Pareto optimal objective vectors (Adapted from Miettinen (1999)) . 90


8.2 Kiviat diagram for the base model . . . . . . . . . . . . . . . . . . . 91
8.3 Optimisation step 1 in-plane forces N22 (meridional direction) . . . . 93
8.4 Optimisation step 1 eigenfrequencies . . . . . . . . . . . . . . . . . . 93
8.5 Kiviat diagram for the first optimisation step . . . . . . . . . . . . . 94
8.6 Optimisation step 2a in-plane forces N22 (meridional direction) . . . 94
8.7 Optimisation step 2a eigenfrequencies . . . . . . . . . . . . . . . . . 95
8.8 Optimisation step 2b in-plane forces N22 (meridional direction) . . . 95
8.9 Optimisation step 2b eigenfrequencies . . . . . . . . . . . . . . . . . 96
8.10 Kiviat diagram for the second optimisation step . . . . . . . . . . . . 96
8.11 Optimisation step 3 in-plane forces N22 (meridional direction) . . . . 97
8.12 Optimisation step 3 eigenfrequencies . . . . . . . . . . . . . . . . . . 97
8.13 Kiviat diagram for the third optimisation step . . . . . . . . . . . . . 98
8.14 Mode shape belonging to the second eigenfrequency for the opti-
mised model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.15 Verification in-plane forces N22 (meridional direction) . . . . . . . . 99
8.16 Kiviat diagram for the final optimisation step . . . . . . . . . . . . . 100

9.1 The optimised model compared to the base model . . . . . . . . . . 102


9.2 In-plane meridional forces of the base model (N22) . . . . . . . . . . 103
9.3 In-plane meridional forces of the optimised model (N22) . . . . . . . 103
9.4 Displacement vector sum of the base model . . . . . . . . . . . . . . 104
9.5 Displacement vector sum of the optimised model . . . . . . . . . . . 104

188
List of Figures

9.6 Optimisation step 3 in-plane forces N22 (meridional direction) . . . . 105


9.7 Eigenfrequencies of the base model and the optimised model . . . . . 106
9.8 Mode shapes belonging to the first 6 eigenfrequencies of the opti-
mised model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
9.9 Optimisation step 3 eigenfrequencies . . . . . . . . . . . . . . . . . . 107
9.10 Mechanical admittance function of the base model and the optimised
model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
9.11 Spectral density of system response for the base model and the op-
timised model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

11.1 Kiviat diagram for the optimised model (blue) and the base model
(red) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

A.1 World energy production (dataset from www.2052.info) . . . . . . . 123


A.2 Working principle (Krätzig et al., 2009) . . . . . . . . . . . . . . . . 124
A.3 Prototype in Manzanares, Spain (Schlaich et al., 2013) . . . . . . . . 126
A.4 Measured output vs calculated output (Schlaich et al., 2013) . . . . 126
A.5 Global map data for various parameters . . . . . . . . . . . . . . . . 127
A.6 Best location map . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
A.7 Total investment costs for a solar updraft tower . . . . . . . . . . . . 129
A.8 LEC costs for a solar updraft tower vs a coal-fired power plant
(Schlaich et al., 2005) . . . . . . . . . . . . . . . . . . . . . . . . . . 130
A.9 Optimal dimensions for a solar updraft tower (Weinrebe and Berg-
ermann, 2010) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
A.10 Examples of NDCTs (en.wikipedia.org) . . . . . . . . . . . . . . . . . 131
A.11 From the largest NDCT to the chimneys of solar updraft towers
(Krätzig et al., 2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
A.12 Circumferential distribution of meridional stresses in a beam and a
shell (Schlaich et al., 2013) . . . . . . . . . . . . . . . . . . . . . . . 132
A.13 First three instability modes of a 1000m tall chimney under (D+We +Wi )
(Krätzig et al., 2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
A.14 Concrete cooling with ice and water (www.kti-plersch.com) . . . . . 134
A.15 Burj Khalifa (www.skyscraperpage.com) . . . . . . . . . . . . . . . . 135
A.16 Alimak Scando 650 FC-S (www.alimakhek.com) . . . . . . . . . . . . 136
A.17 Coriolis effect (www.charlesburrows.com) . . . . . . . . . . . . . . . 139
A.18 Comparison of wind profiles . . . . . . . . . . . . . . . . . . . . . . . 140
A.19 Free damped vibration as a result of an initial displacement (Bach-
mann et al., 1995) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
A.20 SDOF oscillator (Kappos, 2002) . . . . . . . . . . . . . . . . . . . . . 141
A.21 Rayleigh damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
A.22 Damped vibration (Bachmann et al., 1995) . . . . . . . . . . . . . . 142
A.23 Forced response of a system for different damping ratios (Bachmann
et al., 1995) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

B.1 Sample plot in-plane forces N22 . . . . . . . . . . . . . . . . . . . . . 145


B.2 Sample plot eigenfrequencies . . . . . . . . . . . . . . . . . . . . . . 145
B.3 Throat height in-plane forces N22 . . . . . . . . . . . . . . . . . . . . 148
B.4 Throat height eigenfrequencies . . . . . . . . . . . . . . . . . . . . . 148
B.5 Throat radius in-plane forces N22 . . . . . . . . . . . . . . . . . . . . 149
B.6 Throat radius eigenfrequencies . . . . . . . . . . . . . . . . . . . . . 149
B.7 Bottom radius in-plane forces N22 . . . . . . . . . . . . . . . . . . . 150
B.8 Bottom radius eigenfrequencies . . . . . . . . . . . . . . . . . . . . . 150
B.9 Angle of inclination in-plane forces N22 . . . . . . . . . . . . . . . . 151
B.10 Angle of inclination eigenfrequencies . . . . . . . . . . . . . . . . . . 151
B.11 Wall thickness (overall) in-plane forces N22 . . . . . . . . . . . . . . 152
B.12 Wall thickness (overall) eigenfrequencies . . . . . . . . . . . . . . . . 152
B.13 Wall thickness (bottom) in-plane forces N22 . . . . . . . . . . . . . . 153

189
List of Figures

B.14 Wall thickness (bottom) eigenfrequencies . . . . . . . . . . . . . . . . 153


B.15 Wall thickness (top) in-plane forces N22 . . . . . . . . . . . . . . . . 154
B.16 Wall thickness (top) eigenfrequencies . . . . . . . . . . . . . . . . . . 154
B.17 Number of rings (evenly spaced) in-plane forces N22 . . . . . . . . . 155
B.18 Number of rings (evenly spaced) eigenfrequencies . . . . . . . . . . . 155
B.19 Number of rings (spaced 100m from the top) in-plane forces N22 . . 156
B.20 Number of rings (spaced 100m from the top) eigenfrequencies . . . . 156
B.21 Number of rings (spaced 100m from the bottom) in-plane forces N22 157
B.22 Number of rings (spaced 100m from the top) eigenfrequencies . . . . 157
B.23 Aspect ratio rings in-plane forces N22 . . . . . . . . . . . . . . . . . 158
B.24 Aspect ratio rings eigenfrequencies . . . . . . . . . . . . . . . . . . . 158
B.25 Density of concrete in-plane forces N22 . . . . . . . . . . . . . . . . . 160
B.26 Density of concrete eigenfrequencies . . . . . . . . . . . . . . . . . . 160
B.27 Concrete quality in-plane forces N22 . . . . . . . . . . . . . . . . . . 161
B.28 Concrete quality eigenfrequencies . . . . . . . . . . . . . . . . . . . . 161
B.29 Ring stiffness (material properties, all 10 rings) in-plane forces N22 . 162
B.30 Ring stiffness (material properties, all 10 rings) eigenfrequencies . . . 162
B.31 Ring stiffness (material properties, top 5 rings) in-plane forces N22 . 163
B.32 Ring stiffness (material properties, top 5 rings) eigenfrequencies . . . 163
B.33 Ring stiffness (cross-sectional area, all 10 rings) in-plane forces N22 . 164
B.34 Ring stiffness (cross-sectional area, all 10 rings) eigenfrequencies . . 164
B.35 Ring stiffness (cross-sectional area, top 5 rings) in-plane forces N22 . 165
B.36 Ring stiffness (cross-sectional area, top 5 rings) eigenfrequencies . . . 165
B.37 Wind speed in-plane forces N22 . . . . . . . . . . . . . . . . . . . . . 167
B.38 Altitude in-plane forces N22 . . . . . . . . . . . . . . . . . . . . . . . 167
B.39 Latitude in-plane forces N22 . . . . . . . . . . . . . . . . . . . . . . . 168
B.40 Return period in-plane forces N22 . . . . . . . . . . . . . . . . . . . 168
B.41 Pressure distribution curve in-plane forces N22 . . . . . . . . . . . . 169
B.42 Constant damping ratio harmonic analysis . . . . . . . . . . . . . . . 171

C.1 SUMAT logo, modelled after the 4th mode shape . . . . . . . . . . . 173
C.2 SUMAT user interface, displaying the geometry tab . . . . . . . . . 174
C.3 SUMAT user interface, displaying the materials tab . . . . . . . . . . 175
C.4 SUMAT user interface, displaying the loads and options tab . . . . . 176
C.5 SUMAT user interface, displaying the preview tab . . . . . . . . . . 177
C.6 SUMAT user interface, displaying the static tab . . . . . . . . . . . . 178
C.7 SUMAT user interface, displaying the cracking tab . . . . . . . . . . 179
C.8 SUMAT user interface, displaying the buckling tab . . . . . . . . . . 180
C.9 SUMAT user interface, displaying the eigenfrequencies tab . . . . . . 181
C.10 SUMAT user interface, displaying the mode shapes tab . . . . . . . . 182
C.11 SUMAT user interface, displaying the harmonic analysis tab . . . . . 183
C.12 SUMAT user interface, displaying the location tab . . . . . . . . . . 184
C.13 SUMAT user interface, displaying the settings tab . . . . . . . . . . 185

190
List of Tables
E
3.1 Correction factor for other return periods (ASCE7-02, 2002) . . . . . 20
3.2 Equations for the pressure distribution curves H(θ) (Gould, 2005) . 23
3.3 Surface roughness and type of pressure distribution (Gould, 2005) . 24

4.1 Error margins for the intermediate and intermediate+ mesh . . . . . 43


4.2 Error margins for calculated model compared to values found in
literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3 Spring stiffnesses of the foundation (Dowrick, 2007) . . . . . . . . . . 48
4.4 Comparing the eigenfrequencies for several soil stiffnesses with the
fixed foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.5 Young’s modulus of various rock types (Pariseau, 2011) . . . . . . . 51

5.1 Required quantities for the construction of the base model . . . . . . 60


5.2 First 6 eigenfrequencies of the base model . . . . . . . . . . . . . . . 62
5.3 Comparing the eigenfrequencies of an uncracked chimney and a
cracked chimney . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

6.1 Computation time (GUI mode vs batch mode) . . . . . . . . . . . . 69

9.1 Comparing the material use of the optimised model and the base
model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
9.2 Buckling safety factors λ of the base model and the optimised model 105
9.3 Eigenfrequencies of the base model and the optimised model . . . . . 106

A.1 Pumping duration (Lohaus and Abebe, 2010) . . . . . . . . . . . . . 134


A.2 α-values for different types of terrain (Landsberg, 1981; Ruscheweyh,
1982) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
A.3 δ-values for different structural systems (EN1991-1-4, 2010) . . . . . 143

B.1 Throat height sensitivity analysis . . . . . . . . . . . . . . . . . . . . 148


B.2 Throat radius sensitivity analysis . . . . . . . . . . . . . . . . . . . . 149
B.3 Bottom radius sensitivity analysis . . . . . . . . . . . . . . . . . . . . 150

191
List of Tables

B.4 Angle of inclination sensitivity analysis . . . . . . . . . . . . . . . . . 151


B.5 Wall thickness (overall) sensitivity analysis . . . . . . . . . . . . . . 152
B.6 Wall thickness (bottom) sensitivity analysis . . . . . . . . . . . . . . 153
B.7 Wall thickness (top) sensitivity analysis . . . . . . . . . . . . . . . . 154
B.8 Number of rings (evenly spaced) sensitivity analysis . . . . . . . . . 155
B.9 Number of rings (spaced 100m from the top) sensitivity analysis . . 156
B.10 Number of rings (spaced 100m from the bottom) sensitivity analysis 157
B.11 Aspect ratio rings sensitivity analysis . . . . . . . . . . . . . . . . . . 158
B.12 Density of concrete sensitivity analysis . . . . . . . . . . . . . . . . . 160
B.13 Concrete quality sensitivity analysis . . . . . . . . . . . . . . . . . . 161
B.14 Ring stiffness (material properties, all 10 rings) sensitivity analysis . 162
B.15 Ring stiffness (material properties, top 5 rings) sensitivity analysis . 163
B.16 Ring stiffness (cross-sectional area, all 10 rings) sensitivity analysis . 164
B.17 Ring stiffness (cross-sectional area, top 5 rings) sensitivity analysis . 165

192

You might also like