Download as pdf or txt
Download as pdf or txt
You are on page 1of 107

EXAMENSARBETE INOM SAMHÄLLSBYGGNAD,

AVANCERAD NIVÅ, 30 HP
STOCKHOLM, SVERIGE 2019

Numerical and experimental


dynamic analyses of Hägernäs
pedestrian bridge.
- Including seasonal effects.

AIDA IBISEVIC

HASANHÜSEYIN UGUR

KTH
SKOLAN FÖR ARKITEKTUR OCH SAMHÄLLSBYGGNAD
Numerical and experimental dynamic
analyses of Hägernäs pedestrian bridge.
- Including seasonal effects.

Aida Ibišević
Hasanhüseyin Ugur

June 2019
TRITA-ABE-MBT-19473
c Aida Ibišević & Hasanhüseyin Ugur 2019
Royal Institute of Technology (KTH)
Department of Civil and Architectural Engineering
Division of Structural Engineering and Bridges
Stockholm, Sweden, 2019
Acknowledgements

This thesis was conducted at the department of Structural Engineering and Bridges
at KTH Royal Institute of Technology, supervised by Professor Roberto Crocetti and
examined by Professor Jean-Marc Battini. We would like thank you for your endless
support throughout our thesis. Our weekly discussions were incredibly fruitful in
guiding us in our work during the past months. We would also like to show our
appreciation to Professor Costin Pacoste, for providing immense insight in to the
world of dynamics. Thank you for your time and knowledge.
It would not be possible for us to conduct our experiments without the expertise of
Mr. Stefan Trillkott, laboratory engineer at the department. Therefore we would
like to send our sincere gratitude towards you, Stefan, for sticking with us through
long days during both cold and warm weather.
To Professor Lola Martínez Rodrigo; thank you for guiding us throughout our meet-
ings, but also for helping us in conducting our experiments. Know that your enthu-
siasm and participation throughout this master thesis is greatly appreciated.
To Fangzhou Liu and Ph.D. Andréas Andersson, thank you for your assistance in
guiding us throughout our numerical and experimental analyses.
To Fredrik Hernborg, thank you for assisting us in constructing the Finite Element
Model. We also thank KE-gruppen for their warm hospitality.
To John Hallak Neilson, thank you for being a colleague that provided guidance and
meaningful insight throughout the entirety of the thesis. Know that your help during
the experimental process was vital and greatly appreciated but most importantly
thank you for being a colleague and for being willing to share this important chapter
of our lives.
To PhD candidate Daniel Colmenares Herrera, thank you for your assistance and for
your participation during the course of our thesis. It has been a pleasure conducting
all of the experiments in collaboration with you.
Finally, thank you to our family and friends, for not only supporting us in the process
of our Master Thesis, but throughout all five years spent at KTH. Know that your
endless support has meant the world to us.
Sincerly, Aida Ibišević & Hasanhüseyin Ugur

iii
Abstract

Wood as a construction material has in recent years increased, in particular concern-


ing pedestrian bridges. By utilizing wood, the ecological footprint can be reduced,
and the material can be designed to comply with the increasing aesthetic demand
bridge designers are facing. However, as the material weighs little with respect to
its bearing capacity, combined with design becoming more slender, human induced
vibrations are becoming a problem.
Having this in mind, the objective of the thesis is to conduct a case study on an
existing timber pedestrian bridge and assess its dynamic parameters by means of
experiential testing and numerical modelling. The case study concerns the Hägernäs
bridge, an arch bridge located in Hägernäs, Täby. The thesis also considers seasonal
effects by conducting experiments on two separate occasions. In addition, the thesis
evaluates influencing parameters on the dynamic behaviour by conducting a sensi-
tivity analysis. To aid the above mentioned objective, a literature review covering
similar type of analysis is conducted. The literature review also studies the seasonal
effect, mainly from the asphalt layer, as its stiffness contribution is temperature
dependant.
The results from the dynamic parameters showed that not all modes fall above the
recommended values concerning damping ratio (with values above 1-1.5%). How-
ever, all modes fulfill design criteria concerning the magnitude of the natural fre-
quencies.
Furthermore, results showed that the natural frequencies are highly temperature
dependant. The measured values during warm weather (+17◦ C) resulted in lower
values than those from the cold weather experiment (-10◦ C). Moreover, the greatest
difference, by 21% was on the 1st transverse mode and the lower difference was on
the 1st vertical mode, that decreased by merely 5%. Moreover, the damping ratio
was calculated and it was not possible to find any correlation between warm and
cold temperature. Instead, decreased temperature caused some modes to increase in
damping ratio and others to decrease. The most affected mode was once again the
1st transverse mode which increased by 146% going from warm to cold temperature.

v
Finally, modelling the bridge resulted in a stiffer model and thus higher frequencies.
Several aspects where therefore taken into consideration to reflect both the mode
shapes and frequencies from the experiments. The sensitivity analysis concluded
that the behaviour of the arch was vital in order to reach similar results, in particular
all connections comprehending the arches.
Keywords: Pedestrian Timber Arch Bridge, Seasonal Effect, Modal Analysis, Dy-
namic Analysis, Finite Element Method, Experiments

vi
Sammanfattning

Användningen av trä som konstruktionsvirke i byggandet av gång- och cykelbroar


har på senare år ökat. En av anledningarna till detta är att materialets ekologiska
fotavtryck är mindre i förhållande till andra bärande material. Vidare uppfyller
materialet de alltmer estetiska krav brokonstruktörer måste förhålla sig till. Materi-
alets bärförmåga är hög i förhållande till dess vikt. Detta, i kombination med att allt
slankare konstruktioner byggs, medför att överskridande vibrationer, till följd av för-
bipasserande trafikanters last, vibrerar med samma egenfrekvens som trästrukturen.
I och med detta uppstår potentiella problem som måste kartläggas och åtgärdas.
Syftet med vårt examensarbete är således att utforma en fallstudie där vi kartläg-
ger en befintlig bros dynamiska egenskaper med hjälp av numeriska modeller och
experiment. Vidare ämnas undersöka om/hur dessa egenskaper förändras beroende
på utomhustemperatur. Fallstudien behandlar Hägernäs bro, en bågbro som står
i Hägernäs, Täby. Projektet beaktar även den temperaturberoende effekten på de
modala parametrarna, något som studeras genom att experiment utförts på bron
vid två separata tillfällen med olika utomhustemperatur. I syfte att kartlägga de
aspekter med störst inflytande på brons dynamiska egenskaper har även en käns-
lighetsanalys utförts. Vidare, i jämförande syfte har en litteraturstudie utförts som
beaktar liknande analyser där befintliga träbroar blivit experimenterade på och nu-
meriska modellerade. Litteraturstudien beaktar även temperatur effekten med fokus
på asfalten, då dess styvet är temperaturberoende.
Angående resultatet från de dynamiska parametrarna i förhållande till de rekom-
menderade värdena från Eurocode, hade den första transverala moden något mindre
dämpning. Däremot uppfyllde samtliga moder kravet på de naturliga frekvensernas
storhet.
Vidare påvisade resultatet från experimenten att de naturliga frekvenserna är tem-
peraturberoende. De uppmätta resultaten under de varma omständigheterna (+17◦ C)
resulterade i lägre värden än de från experimenten under kalla omständigheter (-
10◦ C). Den egenmod som till högst grad blev påverkad, var den första transversala
moden som minskade med hela 21%, tillskillnad från den första vertikala egenmoden
för vilken temperaturen påverkade minst – en minskning på endast 5%. Dessutom
beräknades dämpningen från de bägge experimenten där det inte gick att finna en
korrelation grundat på temperaturskillnad. Istället var temperatureffekten olika
beroende på egenmod, där vissa minskade och andra ökande med en ökad temper-
atur. Det var däremot återigen möjligt att dra slutsats om att den första transversala
moden påverkades mest, där en ökning på hela 146% förekom.

vii
Slutligen resulterade den numeriska modellen i en alldeles för styv modell vid jäm-
förelse med de uppmätta värdena från testerna. Känslighetsanalysen beaktade där-
för ett flertal parametrar där utfallet behandlade både frekvenser och modformer.
Resultatet från denna analys påvisade att brobågen, inklusive dess koppling till an-
dra bärande delar har störst inflytande för att erhålla resultat likt de från testerna.
Nyckelord: Träbroar, Bågbroar, Temperatur effekt, Dynamisk analys, Finita Ele-
ment Metoden, Experiment

viii
Contents

Acknowledgements iii

Abstract v

Sammanfattning vii

List of Figures xiii

List of Tables xvii

Nomenclature xix

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Hägernäs bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Structural components . . . . . . . . . . . . . . . . . . . . . . 3
1.2.2 Design criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Aims and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Literature Review 7
2.1 Validating FE-modeling with experiments . . . . . . . . . . . . . . . 7
2.1.1 Timber bridge over Bata’s channel . . . . . . . . . . . . . . . 8
2.1.2 The Larden bridge . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.3 Gois footbridge . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Seasonal effect due to the asphalt layer . . . . . . . . . . . . . . . . . 15

ix
3 Method 17
3.1 FE-Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.1 Global coordinate system . . . . . . . . . . . . . . . . . . . . . 18
3.1.2 Individual structural components . . . . . . . . . . . . . . . . 19
3.1.3 Material properties . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.4 Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.5 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.1 Changes implemented . . . . . . . . . . . . . . . . . . . . . . 31
3.2.2 Tuned model . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.3 Experimental analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3.1 Measuring process . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3.2 Setup: Experiment winter season . . . . . . . . . . . . . . . . 37
3.3.3 Setup: Experiment summer/spring season . . . . . . . . . . . 38
3.3.4 Data processing . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4 Results 43
4.1 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1.1 Winter experiment: March 5, 2019, -10◦ C . . . . . . . . . . . 44
4.1.2 Spring Experiments: March 17, 2019, 17◦ C . . . . . . . . . . . 48
4.1.3 Comparison experiments between seasons . . . . . . . . . . . . 52
4.1.4 Hammer input in FEM . . . . . . . . . . . . . . . . . . . . . . 53
4.2 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2.1 Frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2.2 2D Mode shapes . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2.3 3D Mode shapes . . . . . . . . . . . . . . . . . . . . . . . . . 57

5 Discussion 59
5.1 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.2 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

x
5.2.1 Hammer input in FEM . . . . . . . . . . . . . . . . . . . . . . 62
5.2.2 Modelling the asphalt . . . . . . . . . . . . . . . . . . . . . . 62

6 Conclusions 63
6.1 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.2 FE-model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.3 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.3.1 Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.3.2 Modelling the asphalt layer . . . . . . . . . . . . . . . . . . . 64

Bibliography 67

A Phase angles 69

B Technical drawings 73

xi
List of Figures

Figure 1.1 Hägernäs bridge. . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


Figure 1.2 Simplified elevation and plan view, structural components. . . . . 3

Figure 2.1 Timber bridge over Bata’s channel. . . . . . . . . . . . . . . . . . 8


Figure 2.2 The Larden bridge. . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Figure 2.3 Plan view, elevation view and section views of the Larden bridge. 11
Figure 2.4 Gois footbridge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Figure 3.1 Flowchart depicting the workflow in Abaqus/Brigade+. . . . . . . 17


Figure 3.2 FE-model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Figure 3.3 Left arch component . . . . . . . . . . . . . . . . . . . . . . . . . 19
Figure 3.4 Right arch component . . . . . . . . . . . . . . . . . . . . . . . . 19
Figure 3.5 Final part of the left side of the arch. . . . . . . . . . . . . . . . . 20
Figure 3.6 Final part of the right side of the arch. . . . . . . . . . . . . . . . 20
Figure 3.7 Modelling to avoid overlapping elements . . . . . . . . . . . . . . 20
Figure 3.8 Results including thickness . . . . . . . . . . . . . . . . . . . . . . 20
Figure 3.9 Bridge section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Figure 3.10 Abaqus model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Figure 3.11 Location of the simplified cross-section . . . . . . . . . . . . . . . 21
Figure 3.12 Beam 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Figure 3.13 Beam 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Figure 3.14 (a) and (b) presents the connection from the drawing whereas
(c) presents the connection in the FE-model with a tie-constraint rigidly
connecting the two components. . . . . . . . . . . . . . . . . . . . . . . . 23
Figure 3.15 Design of the hanger to steel beam connection. . . . . . . . . . . . 24

xiii
Figure 3.16 Hanger to beam connection. . . . . . . . . . . . . . . . . . . . . . 24
Figure 3.17 Section views presenting the connection between the steel beams
and arch. Note that the red lines in (b) represents the tie constraint. . . 25
Figure 3.18 Design of the deck plate connecting to the glued laminated tim-
berinated beams. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Figure 3.19 Section views of the attachment between the longitudinal beams
and the steel beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Figure 3.20 The design of the crown of the arch. . . . . . . . . . . . . . . . . 27
Figure 3.21 Boundary condition for the support at the upper level of the slope
from both drawings and the FE-model. . . . . . . . . . . . . . . . . . . . 27
Figure 3.22 Boundary condition for the support at the lower level of the slope
from both drawings and the FE-model. . . . . . . . . . . . . . . . . . . . 28
Figure 3.23 Boundary conditions supports at both the ends of the arches and
cross bracing (left side). . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Figure 3.24 Boundary conditions supports at both the ends of the arches and
cross bracing (right side). . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Figure 3.25 Plane view of boundary conditions at both the ends of the arches
and cross bracing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Figure 3.26 Process of acquiring data from experiments. . . . . . . . . . . . . 36
Figure 3.27 Illustrative figure of the accelerometers mounted on a block. The
picture is taken during the winter experiment. . . . . . . . . . . . . . . . 36
Figure 3.28 Plan view showcasing the setup for the winter experiment. . . . . 37
Figure 3.29 Plan view showcasing the setup for the summer/spring experiment. 38
Figure 3.30 Time domain with and without a low-pass filter. . . . . . . . . . . 39
Figure 3.31 Shifted frequency domain caused by a large peak in the time domain. 40
Figure 3.32 Response from a cleared signal. . . . . . . . . . . . . . . . . . . . 40
Figure 3.33 Flowchart depicting the data processing. . . . . . . . . . . . . . . 41

Figure 4.1 FRF vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . 44


Figure 4.2 Frequency domain vertical accelerometers. . . . . . . . . . . . . . 44
Figure 4.3 FRF vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . 45
Figure 4.4 Frequency domain vertical accelerometers. . . . . . . . . . . . . . 45
Figure 4.5 FRF vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . 46

xiv
Figure 4.6 Frequency domain vertical accelerometers. . . . . . . . . . . . . . 46
Figure 4.7 FRF transverse accelerometers. . . . . . . . . . . . . . . . . . . . 47
Figure 4.8 FRF vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . 48
Figure 4.9 Frequency domain vertical accelerometers. . . . . . . . . . . . . . 48
Figure 4.10 FRF vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . 49
Figure 4.11 Frequency domain vertical accelerometers. . . . . . . . . . . . . . 49
Figure 4.12 FRF vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . 50
Figure 4.13 Frequency domain vertical accelerometers. . . . . . . . . . . . . . 50
Figure 4.14 FRF transverse accelerometers. . . . . . . . . . . . . . . . . . . . 51
Figure 4.15 Frequency response during both seasons. . . . . . . . . . . . . . . 52
Figure 4.16 Frequency domain from a hammer input from two different models
from the sensitivity analysis. . . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 4.17 2D mode shapes depicting the 1st vertical mode. . . . . . . . . . . 54
Figure 4.18 2D mode shapes depicting the 2nd vertical mode. . . . . . . . . . 55
Figure 4.19 2D mode shapes depicting the 3rd vertical mode. . . . . . . . . . . 55
Figure 4.20 2D mode shapes depicting the 1st transverse mode. . . . . . . . . 56
Figure 4.21 2D mode shapes depicting the 1st torsional mode. . . . . . . . . . 56
Figure 4.22 1st vertical mode. . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Figure 4.23 2nd vertical mode. . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Figure 4.24 3rd vertical mode. . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Figure 4.25 1st transverse mode. . . . . . . . . . . . . . . . . . . . . . . . . . 58
Figure 4.26 1st torsional mode. . . . . . . . . . . . . . . . . . . . . . . . . . . 58

Figure A.1 Vertical accelerometers. . . . . . . . . . . . . . . . . . . . . . . . 70


Figure A.2 Transverse accelerometers. Note that A14 measures in the longi-
tudinal direction and that A11 was turned off. . . . . . . . . . . . . . . . 71

xv
List of Tables

Table 1.1 Propoerties structural components. . . . . . . . . . . . . . . . . . . 3

Table 3.1 Initial material input in FE-model . . . . . . . . . . . . . . . . . . 22


Table 3.2 Changes considered in the sensitivity analysis. . . . . . . . . . . . 31
Table 3.3 Position, direction and placement of each accelerometer. . . . . . . 37
Table 3.4 Position and placement of hammer and jump tests. . . . . . . . . . 37
Table 3.5 Postion, direction and placement of each accelerometer. . . . . . . 38
Table 3.6 Postion and placement of hammer and jump tests. . . . . . . . . . 38

Table 4.1 Resulting natural frequencies. . . . . . . . . . . . . . . . . . . . . . 52


Table 4.2 Resulting damping ratios. . . . . . . . . . . . . . . . . . . . . . . . 53
Table 4.3 Frequencies compared to the warm weather experiments. . . . . . . 54

xvii
xviii
Nomenclature

Physics Constants

ζ Damping ratio [−]

a Accelerations [m/s2 ]

F Force [kN ]

f Natural frequency [Hz]

t Time [s]

Abbreviations

A Accelerometer

AVT Ambient Vibration Test

BC Boundary Condition

DAQ Data Acquisition Software

DOF Degree Of Freedom

FEA Finite Element Analysis

FEM Finite Element Modelling

FFT Fast Fourier Transformation

FRF Frequency Response Function

GL-Beams Glued Laminated Timber Beams

GLT Gluled Laminated Timber

HPB Half-Power Bandwidth method

LVL Laminated Veneer Lumber

PSD Power Spectral Density

xix
Chapter 1

Introduction

1.1 Background

The most common material used in constructing footbridges is concrete. However, in


recent years wood is increasingly used as a construction material due to its ecological
benefits and is subsequently more frequently applied in bridge design (EEA (2015)).
The combination of high bearing capacity relative to its self weight makes timber
suitable in pedestrian bridge design. The material can be used to produce different
types of bridges ranging from beam bridges to arch bridges, and has the ability to
satisfy the increasing aesthetic demand bridge designers are facing, by incorporating
steel detailing. In addition, wooden bridges allow for an industrial manufacturing
process in which large parts of the structure is produced and transported to site
for easy erection and assembling. Furthermore, to comply with more ecological and
economic standards, bridge designers are choosing slender designs (Pousette and
Fjellström (2009)). The combination of a slender, light and long footbridge result
in uncomfortable vibrations for crossing pedestrians and is thus a comfort issue.
The induced vibrations are related to the natural frequencies of the system, i.e.
the frequency that the structure vibrates with during resonance. Furthermore, the
amplitude of vibrations during resonance is determined by the amount of damping in
the system, and consequentially determining the damping ratio for such structures
is of utmost importance.
One of the most famous examples of human induced vibrations is the Millennium
bridge in London. Due to its slender design severe vibration issues in the lateral
direction occurred. The phenomena resulted in an ongoing research topic about the
dynamic behaviour of bridges (Sétra (2006)). Moreover, it has been proven that
assessing footbridges dynamically is difficult and thus field studies are required for
further understanding (Zivanovic et al. (2007))

1
CHAPTER 1. INTRODUCTION

1.2 Hägernäs bridge

Considering the background, a case study of an existing pedestrian wooden arch


bridge with steel detailing is conducted. The bridge is located in Hägernäs, Täby,
north of Stockholm above highway E18. It has a deck that is 42 meters long with a
free width of 3.5 meters, see Figure 1.1.

(a) Elevation view.

(b) View standing on the bridge deck. (c) Side view.

Figure 1.1: Hägernäs bridge.

The bridge was built by Veidekke AB upon request by Täby municipality. Fur-
thermore, the steel and timber detailing was constructed by Moelven AB, and all
concrete detailing by Konfem Byggkonsult AB. Hägernäs bridge was finished in 2007
and was designed in accordance to the Swedish road administrations code Bro 2004.
However, since 2011, design is regulated according to Eurocode and thus the focus
of this thesis lies on its regulations.

2
1.2. HÄGERNÄS BRIDGE

1.2.1 Structural components

The structural components comprise of a bridge deck made up from two laminated
veneer lumber (LVL) plates glued together. Below the deck plate lies a grid con-
sisting of 5 longitudinal and thirteen 13 glued laminated timber beams. The deck
plate is then carried by the five longitudinal beams. These are subsequently resting
at the ends on concrete abutments as well as on four steel beams, two of which are
directly attached to the glued laminated timber arches. The other two steel beams
are attached to steel hangers which in turn is attached to the arches. The arches are
cast in concrete abutments, as are two steel cross bracing acting as lateral support
that are connected to the steel beams. All components and their properties can be
seen in Figure 1.2 and Table 1.1.

Figure 1.2: Simplified elevation and plan view, structural components.

Table 1.1: Propoerties structural components.

Numbering acc. Cross sectional


Description Material Classification Amount
to Figure 1.2 properties
1 Arch GLT GL30c (L40) 380x900 2
2 Hanger Steel S355 HEA200 4
3 Beam 1 Steel S355 HEA280 4
4 Beam 2 Steel S355 HEA280 4
5 Cross bracing Steel S355 M30 4
6 Cross beam GLT GL30c (L40) 165x315 13
7 Main beam GLT GL30c (L40) 140x585 5
8 Deck plate LVL Kerto-Q 3530x63 2

3
CHAPTER 1. INTRODUCTION

1.2.2 Design criteria

Vibration induced issues are considered in the serviceability limit state. According
to code, all footbridges shall perform to cause no discomfort for pedestrians crossing
the bridge. To measure the comfort, vibrations are considered in terms of eigenfre-
quencies (natural frequencies). Eurocode sets two conditions to satisfy the comfort
criteria (CEN (2001))

1. Eigenfrequency condition

• Vertical eigenfrequencies shall be greater than 5 Hz.


• Horizontal (transverse) and torsional vibrations shall be greater than 2.5
Hz.

2. Amplitude condition. For structures not satisfying the 1st criteria, vibration
amplitudes need to be verified. Verification is performed by determining the
vibration amplitude (accelerations) under specific dynamic loading. The max-
imum amplitude must be

• 0.7 m/s2 for vertical movements,


• 0.2 m/s2 for horizontal movements during normal usage and
• 0.4 m/s2 during exceptional circumstances.

However, the code remarks on the uncertainties in the criteria, and thus designers
often refer to other source material such as the Setra guidelines when designing
bridges with eigenfrequencies lower than the 1st criteria. For wooden bridges, the
code states that when no values have been verified, the damping ratio can be taken
as (CEN (2004))

• ζ = 0.010 for structures without mechanical joints and,

• ζ = 0.015 for structures with mechanical joints.

4
1.3. AIMS AND SCOPE

1.3 Aims and scope

The first objective is to conduct a literature review. It covers previous publications


similar to Hägernäs bridge that have compared the dynamic behaviour of an ana-
lytical model to experiments. In addition, the review covers publications studying
the effect temperature has on the modal parameters due to the asphalt layer.
The main objective, however, is to evaluate the mode shapes and frequencies of
Hägernäs bridge by applying finite element analysis (FEA) and comparing the re-
sults to experimental measurements. Subsequently, the comparison determines the
validity in the FE-model. A sensitivity analysis is then conducted to evaluate the
effect different boundary conditions, constraints and material inputs have on the dy-
namic behaviour. Considering that pedestrians are subjected to discomfort through
vibrations in the bridge deck, focus lies on its behaviour. The scope is limited by
evaluating the natural frequencies and mode shapes by applying modal analysis in
the FE-model and thus solving the eigenvalue problem.
In addition to verifying the mode shape and frequency, the purpose of the exper-
iments is to determine the damping ratio. To further limit the scope, excitation
occurs by means of jumping on the bridge deck, striking the deck with a hammer
and performing ambient vibration tests (AVT). Another objective is to evaluate the
seasonal effect on the modal parameters. Thus, two experiments were conducted,
one during the winter period and another during the spring/summer period.

1.4 Limitations
• No running or walking is considered.

• Humidity is not taken in to consideration.

• Non-linear effects are not regarded in the Finite Element Model.

5
Chapter 2

Literature Review

A literature review is conducted in which the objective(s) of the thesis are explored
by reviewing previous studies. These studies are meant to serve as a basis for the
applied methods and the outcome of the results. Therefore, the review initially
treats studies in which different experimental methods are applied to validate a
finite model for civil engineering structures. Therefrom, previous studies in which
timber footbridges have been dynamically evaluated and analyzed are considered.
Subsequently, these studies contain various modelling approaches as well as different
methods for conducting experiments and are thus of interest for both the modelling
and experimental part of the thesis. Finally, studies in which the seasonal effects
due to the asphalt layer on the modal parameters is reviewed to provide an insight
as to what the cause of the outcome from the experimental analysis may be.

2.1 Validating FE-modeling with experiments

As the complexity of structures increase over time, determining and monitoring its
behaviour for further understanding increases as well. In dynamic analysis, the need
for experimentally evaluating the behaviour is essential and has thus improved over
time. Cunha and Caetano (2006) explore the evolution of experimental analysis
within the field of civil engineering. The authors initially present the input-output
method for dynamic evaluation. This method comprises of determining the modal
parameters by calculating the frequency response function (FRF), mainly by ex-
citing the structure with an impulse hammer. However, with the introduction of
A/C converters, determining the dynamic response is nowadays possible by output-
only methods, such as reconstructing the response through excitation from wind,
also known as an ambient vibration test (AVT). By presenting several experiments
conducted on mainly bridges, the authors conclude that determining the modal re-
sponse with output-only methods is sufficient for dynamically assessing a bridge and
that these methods can be used in finite element validation (Cunha and Caetano
(2006)).

7
CHAPTER 2. LITERATURE REVIEW

Moreover, according to He and Fu (2001) applying finite element modelling as a


tool for designers has throughout the years been proven to be essential in structural
design, and the application of modal analysis in an FE-analysis is imperative in a
dynamic assessment. The authors proceed in explaining how a finite analysis merely
is an approximation of the dynamical applications and a modal analysis is unable to
capture uncertain properties such as damping and non-linear behaviour. Thus, the
need to determine the validity by means of experiments is essential to complement
the finite analysis.

2.1.1 Timber bridge over Bata’s channel

A timber-based bridge is studied in a publication by Čecháková et al. (2012), in


which experimental results and a representative FE-model is compared. The purpose
of the publication was to give an accurate description of the structural behavior
of the bridge, and to create a model that reflects the degradation of the glued
laminated timber. The workflow consisted in performing experimental analysis prior
to the modelling to have the experimental results as a complement in the modelling
procedure.

Bridge description

Contrary to the footbridge investigated in this report, Čecháková et al. (2012) in-
vestigates a road bridge across the Bata’s channel, designed to carry heavy loads
from lorries and farm machines, see Figure 2.1.

Figure 2.1: Timber bridge over Bata’s channel.

The bridge is simply supported, resting on two glued laminated timber beams of
class GL24h with a cross-section 280x1800 mm. The beams have a span of 11.94 m,
an axle span of 5.56 m and are interconnected with seven steel frames of class S355.
Five longitudinal glued laminated timber beams of class GL28h are placed on top
of the flanges of the interconnecting beams. A deck plate made up from transverse

8
2.1. VALIDATING FE-MODELING WITH EXPERIMENTS

timber oak beams with the dimensions 100x160 mm is resting on top of the timber
beams.
The main connection of interest is that between the main glued laminated timber
beams and the steel frames. The connection is made up from steel plates and twelve
M20 bolts. All other timber to steel connections are mainly achieved by utilizing
nuts, bolts, washers, plates and welds. The timber parts are connected with wood
screws.

Experiments

Two sets of experiments were conducted: one static and one dynamic. The structure
is statically experimented on by evaluating the stresses in the main glued laminated
timber beams by loading the structure with the use of a vehicle. The stresses
were measured using strain gauges at the top and bottom fibre of the main glued
laminated timber beams at midspan. The dynamic evaluation was conducted by
crossing the bridge with a vehicle moving at a speed of 15 km/h, with the purpose
of exciting the eigenmodes.

FE-model

An FE-model in Scia Engineer is constructed where all structural components were


modelled with respect to their inter-relational placement. Connections with eccen-
tricities were created with the use of rigid links. The main glued laminated timber
beams were modelled as orthotropic plates to account for the orthotropic behavior.
Separate beams were modeled for each timber plank that makes up the deck to
exclude a deck plate behaviour.

Results & Conclusions

The experimental results from the static tests resulted in an 82 % accuracy. However,
the dynamic tests failed to evoke any eigenfrequency and thus no information about
the accuracy in the dynamic behaviour was concluded. Finally, the authors express
that further analysis (which already is in progress) is necessary to realistically model
the degradation and behaviour of the connection between the glued laminated timber
beams and the steel frames.

9
CHAPTER 2. LITERATURE REVIEW

2.1.2 The Larden bridge

Another timber-based bridge is studied by Rønnquist and Wollebaek (2019) that


entailed modelling and experimentally assessing the dynamic and static behaviour of
the bridge. The intention of the publication is to review the structural integrity as it
is experiencing lateral induced vibrations causing discomfort for passing pedestrians,
especially during concerts in the nearby area.

Bridge Description

The Larden bridge is a slender glued laminated timber footbridge located in Norway
above Numedalslågen river. Its longest free span is 92 m with a total length of 130
m, see Figure 2.2.

Figure 2.2: The Larden bridge.

The structure is divided into three parts(A-B, B-C and C-D), see Figure 2.3. A-B
and C-D is identical at its bottom parts in which they consist of glued laminated
timber trusses strengthened by horizontal glued laminated timber beams below the
deck, see section 2-2. Part B-C is at its bottom part strengthened using both glued
laminated timber trusses and steel trusses. The top part (section 1-1) consist of two
glued laminated timber beams (main beams) attached to a steel truss (L-shape). On
top of the main beam rests glued laminated timber cross beams (100x200 mm) at an
axial distance of 2.2 m. Four longitudinal glued laminated timber beams (100x160
mm) rest on top of the crossbeams and on top of the longitudinal beams is the deck
plate, that is made up from timber planks (45x120 mm). The top part (section 1-1)
is consistent along the entire length of the structure.

10
2.1. VALIDATING FE-MODELING WITH EXPERIMENTS

Figure 2.3: Plan view, elevation view and section views of the Larden bridge.

Furthermore, attachments between main beams and crossbeams provided minimal


stiffness in the transverse direction, and for that purpose the steel truss is located
in transverse direction connecting to the main beams. The steel truss was attached
by utilizing screws. To account for the difference in thermal expansion between the
two materials, screws were placed in elongated holes. However, as a consequence,
the stiffness contribution becomes difficult to determine.

Experiments

The bridge was experimented on by deploying full scale measurements by means


of three methods: (1) Stamping or jumping at midspan (2) People standing with
wide footing rhythmically rocking their body in the transverse direction and (3)
Applying an electrically driven exciter consisting of counter rotating masses. The
purpose was to apply the first method to acquire the vertical modes, the second
method to acquire the horizontal and transverse modes and the third to excite
the bridge with a known harmonic force and amplitude. All measurements were
conducted by measuring accelerations in the horizontal and vertical directions. The
type of experiments conducted for the static experiments were not conveyed in this
publication but the results were mentioned for comparison reasons.

11
CHAPTER 2. LITERATURE REVIEW

FE-model

All models were created using Ansys. All static models were in 2D whereas the
dynamic models were created in 3D. Two dynamic models were considered, one using
a more constricting version of Ansys and another which allows for a more detailed
model. Both models utilize the free vibration analysis to obtain natural frequencies
and mode shapes. Unfortunately, these models did not manage to capture the
frequency. The authors therefore proceeded in modelling the bridge using FrameIT,
a program developed by one of the authors during his doctoral thesis. FrameIT
allows for more versatile and detailed modelling. The model in FrameIT focuses on
correctly modelling the mass and stiffness. Therefore, the railings were placed as
masses at the same position as it is in drawings and the deck plates were assumed
to have almost no stiffness contribution and was therefore also modelled as a mass.
All other structural components were modelled as beam and bar elements. Material
properties were according to the authors assigned as "normal" and rigid links and
hinges were deployed for the connections.
To try to create a model close to the experimental values, the connective stiffness
between the main beams and the steel truss were altered. The same method was
applied to the connection between the main beams and the cross beams which
resulted in a so called "tuned model".

Results & Conclusions

The results from the static experiments compared to the 2D static analysis were
very accurate. However, as mentioned, the 3D models built in Ansys managed to
capture the mode shapes fairly well, but had for instance a 14% overestimation of
the 1st horizontal mode. The (untuned) model made in FrameIT was also too stiff,
but by tuning the stiffness in the connections, the authors managed to capture the
1st horizontal mode precisely.
The concluding remarks include the level of difficulty that lies in modelling the
stiffness in the connections (the authors denotes the stiffness to be "softer" than the
applied rigid links). These connections were mainly the steel to timber connections
as well as the connection between the deck structure and the glued laminated timber
beams. Finally, the authors reflect upon the fact that a very simple static 2D model
results in very accurate results whereas the fairly complex 3D models (untuned)
fail to capture the dynamic behaviour. Thus, Rønnquist and Wollebaek (2019)
conclude that a dynamic evaluation requires a more detailed modelling procedure
in comparison to a static evaluation.

12
2.1. VALIDATING FE-MODELING WITH EXPERIMENTS

2.1.3 Gois footbridge

In another study, Cruz et al. (2009) conducted a dynamic evaluation and analysis
on the Gois footbridge, located in Gois, Portugal. Gois footbridge is a timber based
arch bridge. The purpose of the publication is to evaluate the condition of the
bridge and to investigate potentially harmful dynamic effects from pedestrian-bridge
interactions.

Bridge description

Gois footbridge has a similar design to Hägernäs footbridge, see Figure 2.4. Glued
laminated timber of class GL24h is used for the bearing structural components.
Timber planks make up the 2.15x31 m deck that is resting on a grid made up
from glued laminated stringers and crossbeams. The longitudinal edges of the grid
is connected to the two glued laminated longitudinal beams with a span of 30.6
m. Two three-hinged glued laminated timber arches suspend the deck with twelve
hangers on each side. The arches are connected to one another using glued laminated
timber beams at the position between the hangers. Lateral support for the arches
and the deck is provided by steel struts.

Figure 2.4: Gois footbridge.

Experiments

The authors proceed with the experiments by conducting dynamic tests with respect
to four different cases. The setup consists of 8 accelerometers. A preliminary dy-
namic test was performed in the first case. Ambient vibration tests was performed
for the second and third test, with an additional mass over the deck for the latter.
In the fourth case, the bridge was deliberately excited with pedestrians.

13
CHAPTER 2. LITERATURE REVIEW

FE-modelling

The authors acknowledge the important influence connections have on the overall
structural behavior of timber-based footbridges. They also address the fact that
there is a deficiency of quantitative information regarding this subject.
The initial modelling approach (performed in SAP2000) consists of each connection
being released from rotation in the main local direction, with the exception of the
handrail in which every degree of freedom (DOF) was released. Hinged connections
were used for the supports in the arches and the deck. Each structural part was
modelled apart from the timber planks that were modelled as an added mass.
The obtained natural frequencies and mode shapes were compared to those from the
FE-analysis. The comparison is conducted by applying the Mass Matrix Method
which yields a scale factor that is used for comparison. The results show that the
experimental values were higher for each mode (the six first modes are considered)
and that subsequently the model was too stiff. A model tuning was therefore per-
formed with the purpose of obtaining a model closer to the experimental values.
These changes mainly applied to the stiffness of the main connections. This was
performed by conducting a sensitivity analysis in which 9 critical points (connec-
tions) were identified. The degrees-of-freedom for each identified point is released
one DOF at a time. Based on the results from the ATV, 30 parameters were changed
in order to come closer to the experimental values.

Results & Conclusions

Tuning the model resulted in very accurate results with a maximum difference in
eigenfrequency being 8.37 %. However, the untuned model was stiffer in comparison
to the 3rd test for all modes with the maximum difference being for the 4th mode at
59 %. The results clearly show that a higher mode tend to result in less accurate
results. As mentioned, the tuning mainly consisted of identifying the stiffness in
the connections and thus the authors conclude its vital part in correctly modelling
a timber-based structure for dynamic evaluation purposes.

14
2.2. SEASONAL EFFECT DUE TO THE ASPHALT LAYER

2.2 Seasonal effect due to the asphalt layer

Damping ratio is essential for reducing the dynamic amplification factor during reso-
nance and the natural frequencies are vital in design criteria. In the field of structural
dynamics, several studies reviewing the modal parameters for timber bridges have
been conducted. A study performed by Hamm (2007) determined the damping ra-
tio for 19 wooden bridges with different span lengths and structural systems. All
bridges had a 1st vertical frequency under 4 Hz, and was thus below the 1st condition
in design criteria. The study concluded that the bridges with asphalt pavement had
a larger energy dissipation (ζ = 1.1% − 3.3%) than without (ζ = 0.2% − 1.35%).
Furthermore, in a study conducted by Gülzow et al. (2015) a wooden bridge was
tested in terms of its dynamic properties. The study determined the damping ratio
and natural frequencies by means of forced and ambient vibration tests prior to and
after the asphalt pavement had been constructed. The results from the ambient
vibration tests resulted in a damping ratio between 0.9 - 2.4 % and natural frequen-
cies between and 6.13 - 13.8 Hz without the asphalt and a damping ratio between
1.7 - 4.1 % and natural frequencies between 5.25 - 12.34 Hz with the asphalt. The
forced tests resulted in a damping ratio between 1.6 - 1.8 % and natural frequencies
between and 6.23 - 24.2 Hz without the asphalt and a damping ratio between 2.1 -
2.6 % and natural frequencies between 5.19 - 19.71 Hz with the asphalt.
The studies above determine the effect asphalt has on the damping ratio and the
natural frequencies and that its contribution should not be neglected. In order to
further study the asphalt effect, a laboratory study was conducted by Gsell et al.
(2010) in which a simply supported and cable stayed wooden bridge were tested
and modeled. The bridges were tested in room temperature by means of different
vibration tests, in which the damping ratio and natural frequency was determined,
with and without pavement. The beneficial contribution of asphalt on the damping
and its softening effect on the natural frequencies was yet again verified. Gsell et al.
(2010) proceeded in modelling the stiffness effect of the asphalt in FEM with two
different assumptions: (1) No shear transfer between the asphalt and timber deck
and (2) Full shear force transfer between the asphalt and timber deck. Furthermore,
the damping properties was modelled using Rayleigh damping. To account for the
temperature effect on the shear modulus, the shear loss was calculated in accordance
to coaxial tests for temperatures at -5◦ C and +40◦ C. Results conclude that full shear
transfer and Rayleigh damping overestimates the damping ratio, whereas no shear
effect and Rayleigh damping is an underestimation. Subsequently, the authors con-
clude that the effect temperature has on the asphalt, combined with its viscoelastic
properties, is crucial for correctly determining the damping ratio. The study con-
ducted by Gsell et al. (2010) was further investigated by Feltrin et al. (2011) in
which the same specimens were excited using a shaker from which accelerations, air
temperature and pavement temperature were measured. Tests were then conducted
with and without the asphalt layer in which the 1st natural frequency and damping
ratio were evaluated. Two timber panels were tested, one during the winter season
and another during the summer season. Results showed that the summer tests had
a 4.3 % increased natural frequency and a 413 % increased damping ratio with the

15
CHAPTER 2. LITERATURE REVIEW

pavement. For the summer tests, having no pavement decreased the natural fre-
quencies by 32% and the damping ratio by merely 50 %. However, the difference in
air temperature and asphalt temperature was 20◦ C in the summer tests as opposed
to 0◦ C in the winter tests.
In an attempt to numerically model the effect of the asphalt layer, Feltrin et al.
(2011) created two numerical models in which one was modelled with full shear effect
and the other with no shear effect. The viscoelastic properties were determined from
torque tests at different temperatures and loading frequencies. Tests showed that the
biggest influence factor on stiffness, for temperatures ranging from -10◦ C to 40◦ C,
was the storage modulus which changed by a magnitude factor of two, whereas the
shear loss modulus changed by a factor of one. To calculate the natural frequency,
the authors assumed that the structure was a simple beam with pinned boundary
conditions. Results showed that shear effect can be noted at low temperatures, since
the full-shear-model had a natural frequency of 6.4 Hz, the no-shear-transfer-model
3.58 Hz and the experiments 5.33 Hz. At high temperatures, modelling with or
without shear effect resulted in a similar frequency. These were, however, closer
to the values measured without the effect of the pavement. In terms of damping,
the summer results showed some shear effect since the no-shear-transfer-model had
a damping of 0.55% whereas the full-shear-transfer-model had a damping ratio of
2.4% and the measured values were 0.8% without and 1.2% with the pavement.
The authors proceeded in conducting measurements and tests on two existing timber
bridges with different types of deck constructions to account for the real behaviour.
Tests measured the modal parameters by means of ambient vibration tests and
jump tests. These tests were conducted at both high and low temperatures. Results
showed once again a decreasing natural frequency with higher temperature whereas
the damping ratio remained unchanged. To study the seasonal effect in detail, jump
tests every 30 minutes were conducted on one of the bridges for an entire day.
Accelerometers, air temperature and asphalt temperature were measured. Natural
frequency and damping ratio were then determined for each temperature. The
natural frequency decreased with an increased air temperature and had thus a strong
correlation. The correlation between damping ratio and air temperature, however,
was minimal. Final conclusions entailed that the asphalt layer has an effect on
the modal parameters of a timber bridge. Furthermore, the authors conclude that
the asphalt layer has greatest impact on the eigenfrequencies at high temperatures.
Therefore, the authors conclude that numerically modelling the seasonal effect can
be performed with two methods: (1) Modelling the asphalt layer as an added mass to
capture the lower frequency range and (2) Modelling the asphalt layer with full shear
effect to capture the higher frequency range. Finally, the damping ratio from both a
theoretical model and laboratory tests concluded that a higher temperature results
in a higher damping ratio. The tests on a real structure, however, turned out to not
support these results. Therefore, the authors conclude that other mechanisms, such
as joints, materials etc. combined with the uncertainties in the methods deployed
for determining the damping restricts the possibility of evaluating the effect of the
asphalt layer on the damping ratio.

16
Chapter 3

Method

In the following section, the method used to study the objectives are presented. The
thesis is divided into 3 separate parts consisting of the modelling procedure for the
FE-model, the sensitivity analysis and the experimental process.

3.1 FE-Model

All finite element modelling is performed with the use of Brigade+/Abaqus, in which
the workflow is conducted using modules. These can be seen in Figure 3.1.

Part Property Assembly Step Interaction

Visualization Job Mesh Load

Figure 3.1: Flowchart depicting the workflow in Abaqus/Brigade+.

17
CHAPTER 3. METHOD

3.1.1 Global coordinate system

Figure 3.2 illustrates a finale version of the FE-model, including the global coordinate
system. In the following sections on the FE-modelling as well as in Section 3.2
depicting the process of the sensitivity analysis, referencing the global coordinate
system commonly occurs. Thus, this section clarifies the global coordinate system
with relation to the global geometry of the bridge.

YZ

Z
Y

Figure 3.2: FE-model.

Global geometry

x-axis: the direction along the length of the bridge deck (longitudinal direction).
y-axis: the direction along the width of the bridge deck (transverse direction).
z-axis: the direction along the height of the bridge deck (vertical direction).

Commonly used terminology

In-plane rotation: rotation about the y-axis.


Out of plane rotation: rotation about the x-axis.

18
3.1. FE-MODEL

3.1.2 Individual structural components

In this section, the modelling procedure for each structural component is depicted.
This includes geometry and element type. The main purpose of this section is to
present the modelling procedure by presenting significant aspects. These mainly
entail assumptions that deviates from design, but are crucial to maintain a model
such that it is within reason with regards to detailing.

Deck plate

The deck plate is modelled as a shell element with the same length as the longitudinal
beams, 42.068 m. The width of the deck is 3.53 m and it is modelled to include the
slope of 7.5%, that the real structure has. The deck plate consists of two laminated
veneer timber (LVL) plates with a thickness of 63 mm that are glued and screwed
together. In the FE-model, a simplification is made in which the deck is modelled
with a thickness of 126 mm. This approach is valid since there is full interaction
between the plates. To account for the asphalt layer on top of the deck plate, a
higher density in the material definition of the deck is applied, with the purpose
being to have representative weight on the 126 mm thick deck plate. Since the deck
plate is orthotropic, the material is defined as an orthotropic plate.

Arches

In the creation of the arches, a sketch was made in AutoCAD to represent the correct
curvature and length of the arch. The sketch was imported in to Abaqus, in which
a 3D beam element was created. Since the arch is non-symmetrical with differing
distance between connections, the arch was divided in to two parts with respect
to the crown. The imported sketches, which represent the left side and right side
of the crown are presented in Figures 3.5 and 3.6, respectively. The vertical lines
represent connection points which are used to connect all structural components
correctly with regard to their placement. The diagonal lines are used as reference
lines to globally position the arch in relation to the deck.

Figure 3.3: Left arch component Figure 3.4: Right arch component

After the arches have been placed in the structural system to accommodate the
global geometry, the lines are removed resulting in the final arch components, as

19
CHAPTER 3. METHOD

depicted in Figures 3.5 and 3.6.

Figure 3.5: Final part of the Figure 3.6: Final part of the
left side of the arch. right side of the arch.

Glued laminated beams

The glued laminated beams (GL-beams) were all modelled in the same part as beam
elements, in which there are five longitudinal beams with a fixed axial distance of
0.735 m and thirteen transverse beams with varying axial distances. To avoid over-
lapping elements the transverse beams are modelled with a small distance between
its end-points and the longitudinal beams. This distance corresponds to half the
thickness of the longitudinal beams as can be seen in Figure 3.7.

Z X

Figure 3.7: Modelling to avoid Figure 3.8: Results including


overlapping elements thickness

In design, the transverse beams are cutoff at their ends, see Figure 3.9. This has not
been taken in to consideration since it does not have any structural functionality
and its effect on the mass of the structure is extremely low. Instead, a simplification
was made in which the transverse beam was assumed to have a uniform section
throughout, as can be seen in Figure 3.10. The simplification was applied to the
parts coloured red in Figure 3.11.

20
3.1. FE-MODEL

Figure 3.9: Bridge section Figure 3.10: Abaqus model

Figure 3.11: Location of the simplified cross-section

Steel beams

The steel hangers and beams were created as 3D beam elements. Partitions in the
HEA beams were created in connections to the hangers and the longitudinal beams
for connecting purposes. Figures 3.12 and 3.13 illustrates the partitions/nodes for
beam 1 & 2 from Table 1.1.

Figure 3.12: Beam 1.

Figure 3.13: Beam 2.

Hangers and cross bracing

The hangers were modelled as 3D beam elements, defined by two nodes at each end.
Truss (bar) elements were used in the modelling of the cross bracing since they are
functionally limited to axial forces.

21
CHAPTER 3. METHOD

3.1.3 Material properties

Material inputs for the structural components are presented in Table 3.1. With the
exception of the deck, each section is assigned an orientation which specifies the
alignment of the local axis in relation to the global axis. For the deck plate, the
weighted density considering the mass of the asphalt is included.

Table 3.1: Initial material input in FE-model

Ex Ey Gxy Gxy Gyz ν ρ


Structural component Material
[GPa] [GPa] [GPa] [GPa] [GPa] [-] [kg/m3 ]
GL-beams & Arches GL30c 13 - - - - 0 440
Deck plate LVL 10.5 2.4 0.6 0.12 0.22 0 1940
Steel beams, Hangers
S355-J2 210 - - - - 0.3 7850
& Cross bracing

3.1.4 Connections

The FE-modelling approach described for the connections is the initial modelling
approach. Each type of connection is made up of a rigid link with variable properties
relating to translation in and rotation around the global axis x, y, z. The modifiable
properties have the optional alternative to assign a stiffness in a certain direction.
The connections in the initial model, which overwhelmingly consists of ties, are
represented by a restraint in every translation and rotation. Most of these ties
create an eccentricity at the connections, due to the fact that the real dimensions of
the existing bridge are used, which subsequently results in connections that are not
node to node. However, these connections are a subject of change in the sensitivity
analysis in Section 3.2.

Hanger to arch

The connections between the hangers and the arches are assumed to be rigid. The
hanger is pre-stressed and bolted to the glued laminated arch, see Figures 3.14a
and 3.14b. The connective point in the FE-model is presented by a tie-constraint
that ties all translations and rotations between each respective arch and hanger, see
Figure 3.14c.

22
3.1. FE-MODEL

(a) Elevation view: (b) Section view:


Drawing. Drawing. (c) FE-model.

Figure 3.14: (a) and (b) presents the connection from the drawing whereas (c)
presents the connection in the FE-model with a tie-constraint rigidly
connecting the two components.

Steel beam to hanger

The hangers are connected to the steel beam with six M27 nuts bolted with M27
screws. The connection is designed to be rigid in static design. In the initial FE-
model this connection was represented by a tie constraint between hanger and beam,
as illustrated in figure 3.16a. Since the beam elements are defined by their center
line, a small eccentricity which correspond to half the height of the beam, 0.135 m,
is set as the distance between the components.
However, it was noted that this had a unfavourable impact on the behavior of the
arch, resulting in mode-shapes that deviated from the actual behavior of the existing
bridge. It was noted that the stiffness of the arch increased, which led to the out
of plane rotation being affected. Taking into consideration that this is a dynamic
analysis, the connection between the hanger and the beam was changed to a hinge-
fix constraint with the aid of a rigid body. The rigid body was modelled with
the hinge at the connection to the steel beam, and a tie at its connection to the
hanger, resulting in a constraint that restrains all translations and rotations, with
the exception of out of plane rotation. The rigid bodies are the red lines illustrated
in figure 3.16b.

23
CHAPTER 3. METHOD

(a) Elevation view: (b) Section view:


Drawing. Drawing.

Figure 3.15: Design of the hanger to steel beam connection.

(a) 3D view: (b) 3D view:


FE-model. FE-model with rigid connection.

Figure 3.16: Hanger to beam connection.

24
3.1. FE-MODEL

Steel beam to arch and cross-bracing

As previously mentioned, one way that the arches connect to the deck plate is with
the use of steel beams attached at their ends to the arch. The connection comprises
of utilizing ten steel rods (type M24) fastened between the two components. These
steel rods are fastened through holes in a plate welded onto the ends of the steel
beams. The connection is further strengthened by utilizing toothed bulldog washers.
Thus, this connection is entirely restrained in terms of rotations and translations. In
the FE-model, the connection is modelled with a tie having the length 0.193 m. The
length of the tie represents the distance from the end of the steel beams to centre
line of the arch, the position at which the arches are defined. The cross-bracing is
connected to the steel beam through plates that are welded at its ends, depicted in
Figure 3.17a. Since the cross-bracing only is subjected to axial forces, and due to
the fact that they are modelled as truss elements, the connection between the steel
beam and cross-bracing is modelled as a tie. This tie is not a subject of change later
in the analysis, since there is an eccentricity in the existing structure.

(a) Section view: Drawing. (b) Sections view: FE-model.

Figure 3.17: Section views presenting the connection between the steel beams and
arch. Note that the red lines in (b) represents the tie constraint.

Longitudinal beam to transverse beam

The transverse beams do not have any major structural purpose for the dynamic
analysis, other than the weight contribution. Therefore the connection between
the longitudinal beam and the transverse beam does not carry any significance.
The connection between the structural components consists of screws that connect
the deck, the transverse beams and the longitudinal beams. This is depicted in
Figure 3.18a. Consequentially, a tie is set at the intersecting points between the two
structural components.

25
CHAPTER 3. METHOD

Longitudinal beam to deck

The deck is connected to the longitudinal glued laminated beams with screws, see
Figure 3.18 at distances ranging from 0.33-0.375 m, and glue. Since the longitu-
dinal beams are designed to carry the translations and rotations from the deck,
the connections are assumed to be rigid. The rigid connection is achieved by the
same means as above mentioned, with tie constraints. A line to line tie-constraint
is created, with a tie-length 0.3 m that accounts for the eccentricity.

(a) Bottom panel of the LVL-deck. (b) Top panel of the LVL-deck.

Figure 3.18: Design of the deck plate connecting to the glued laminated timberinated
beams.

Longitudinal beams to steel beams

Each longitudinal beam is resting on the HEA-beam, and placed between two steel
plates welded to the steel beam. The steel plates have elongated holes to account for
the difference in the thermal expansion coefficient between steel and timber. Going
through the elongated wholes are steel rods that connect the timber with steel.
In Abaqus this is initially modelled as a tie constraint, similar to above described
connections. The following Figures depict the connection, in both design and from
FEM.

(a) Section of the GL-beams attached to (b) Section of the GL-beams attached to
the steel beam. the steel beam from the FE-model.

Figure 3.19: Section views of the attachment between the longitudinal beams and
the steel beams

26
3.1. FE-MODEL

Crown of the arch

At the crown of the arch, a hinge is positioned to restrain all translations as well
as for allowing in-plane rotation. In the FE-model, the hinge is modeled with a
coupling constraint. This type of constraint enables the crown to act as a hinge.
Figure 3.20, illustrates the design of the hinge.

(a) Elevation view. (b) View from above.

Figure 3.20: The design of the crown of the arch.

3.1.5 Boundary conditions

The bridge is resting on concrete foundations. For each glued laminated timber
beam two steel plates are cast in to the concrete. The beams are then placed in
between the plates and screwed together. At the ends of the glued laminated timber
beams positioned at the upper level of its slope, the steel plates are designed with
holes such that no room is left for movement, see Figure 3.21a, consequentially all
translations are restrained. The boundary condition at the lower level of the slope,
however, has plates with room for movement in the longitudinal direction and is
therefore free to translate in that direction, see Figure 3.22a. The connection also
restrains rotation out of the plane, and rotation around its own axis. In the FE-
model, the support is modelled with respect to the eccentricity to account for the
height of the longitudinal beams at each end, see Figures 3.21b and 3.22b.

X
Y

(a) BC upper level slope,


wholes with no room for (b) BC upper level slope,
movement. FE-model.

Figure 3.21: Boundary condition for the support at the upper level of the slope from
both drawings and the FE-model.

27
CHAPTER 3. METHOD

(a) BC lower level slope,


wholes with room for (b) BC lower level slope
movement. FE-model.

Figure 3.22: Boundary condition for the support at the lower level of the slope from
both drawings and the FE-model.

Moreover, the arches are, as previously mentioned, three-hinged-arches, and are thus
modelled as such. At the support, as well as at the crown, the arches have a hinge
which restrains translations in all directions and frees in plane rotation. The hinges
at the support are cast in concrete abutments. Cast in the same abutments is the
cross bracing, that is modelled as fully fixed. In Figures 3.23 and 3.24 the design of
these supports are depicted, as well as its representative FE-model.

(a) BC left side: (b) BC left side:


Drawings. FE-model.

Figure 3.23: Boundary conditions supports at both the ends of the arches and cross
bracing (left side).

28
3.1. FE-MODEL

(a) BC right side: (b) BC right side:


Drawings. FE-m.

Figure 3.24: Boundary conditions supports at both the ends of the arches and cross
bracing (right side).

(a) BC left side, plane view: (b) BC right side, plane view:
Drawings. Drawings.

Figure 3.25: Plane view of boundary conditions at both the ends of the arches and
cross bracing.

29
CHAPTER 3. METHOD

3.2 Sensitivity analysis

A sensitivity analysis is conducted with the purpose of evaluating parameters that


effect the stiffness and the mass of the structure. By doing so, the most crucial influ-
encing factors with regards to its dynamic behaviour can be determined. However,
to conduct such an analysis, the initial FE-model, being modelled as explained in
the section above, is re-modelled in terms of connections. Instead of utilizing tie
constraints to connect the different parts of the bridge rigid bodies are created so
that the DOF:s at each end of the connecting cross-section easily can be changed at
see fit, similar to the connection described in the second paragraph of Section 3.1.4.
Prior to the initiation of the sensitivity analysis, the cross bracing was removed.
The reason for this was to evaluate the possibility to model the bridge without them
since the initial model greatly overestimated the stiffness of the bridge.
Moreover, as described in previous sections, the FE-model is created with the aid of
technical drawings and subsequently the design from a theoretical standpoint. By
conducting a sensitivity analysis and reviewing the effects on the modal parameters
and having the experimental results as a basis, a decision was made to account
for possible small differences due to real life construction practice. The analysis
also provides a deeper understanding of how the structure acts during dynamic
excitation. The procedure of the sensitivity analysis is thus to conduct one change
at a time, from which the effect on modal parameters are studied (the frequency and
mode shape). By studying one parameter at a time, its single-handed influence can
be determined, and thereby its effect on the dynamic properties of the structure.

30
3.2. SENSITIVITY ANALYSIS

3.2.1 Changes implemented

In the following Table, all changes conducted within the sensitivity analysis are
presented.

Table 3.2: Changes considered in the sensitivity analysis.

Model version Change


1 Restrained BC GL-beams
2 No eccentricity BC for GL-beams
3 Free My between main beams & steel beams
4 Only translations between main beams and deck
5 Fixed-hinge constraint between hanger and arch
6 10 % increase Ex , LVL
7 10 % decrease Ex , LVL
8 10 % decrease Ey , LVL
9 10 % increase Ey , LVL
10 10 % decrease Gxy , LVL
11 10 % increase Gxy , LVL
12 Springs between steel beams and main beams in the vertical direction
12-1 Springs between steel beams and main beams in the vertical direction.
Applies only for the steel beams attached to the hangers
12-2 Springs between steel beams and main beams in the vertical direction.
Applies only for the steel beams attached directly to the arch
14 10 % increase E, GL30c
15 10 % decrease E, GL30c
16 All rotations free at the hinge located at the crown of the arch
16-1 All rotations restrained at the hinge located at the crown of the arch
16-2 At the support of the arches, all rotations are free
16-3 At the support of the arches, all translations and rotations are restrained
17 Cross bracing included
18 Poisson’s ratio 0.1 for the LVL
19 10 % decrease Gxz , LVL
20 10 % increase Gxz , LVL

Boundary conditions at the GL-beams

Changes implemented in model version 1 & 2 entailed changing the boundary con-
dition at the glued laminated timber beams. These changes included removing the
initial eccentricity that the beams had. The basis for this approach was that the
mode shapes obtained from experiments deviated from the FE-model at the ends.
Changing the boundary condition was therefore a means to test whether or not the
model is sensitive in that regard.

31
CHAPTER 3. METHOD

Connections

Model version 3 entailed freeing the rotational DOF along the y-axis between the
main beams and steel beams. A change that was implemented since this connection
is highly dependant on the rods between the two components, which are described in
Section 3.1.4. The holes through which the rods are position are slightly elongated
and may therefore introduce some rotational freedom.
In the 4th model version, the rotations between the deck plate and the longitudinal
beams were released. This was deemed within reason since the longitudinal beams
should not be bending along with the deck plate.
The 5th model version consisted in modelling the connection between the arch and
the hanger as fixed-hinge (i.e fixed at the arch and a hinge at the hanger). The hinge
was only free to rotate in-plane. This change was implemented after consultation
with supervisors.

Material parameters

Model versions 6-11 and 19-20 entailed decreasing and increasing the stiffness of the
LVL-panel. These changes where conducted on all stiffness parameters, taking into
account that the LVL is modeled as an orthotrophic plate. In model version 18 the
Poisson’s ratio for the deck plate was overestimated. In model versions 14 and 15,
the stiffness for the glued laminated timber beams were increased and decreased to
consider their effect.

Implementing springs

After consulting supervisors, the conclusion was drawn that the assumption about
the vertical movement between the longitudinal beams and steel beams is entirely
restrained may be incorrect. A guarantee of a rigid connection can only be estab-
lished for static design. The reason being that once again the holes through which
the steel rods are positioned are elongated and allow for slight movement in the ver-
tical direction. To account for this, springs with a stiffness of 107 N/m is modelled.
Model version 12 is modelled with springs at the connective points between all steel
beams and main beams. The version labeled 12-1 has springs at the connections at
which the steel beams are directly attached to the arch, and model version 12-2 has
springs where the steel beams attach to the hangers.

Arches

All model versions labeled 16 consider changes with regards to the three-hinged arch.
Instead of considering the arch as a hinge allowing for in plane rotation, different
model versions that both frees and fully restrains the hinge at the crown is considered
in model versions 16 and 16-1. At the hinges making up the support of the arches,
models considering both a boundary condition with a hinge fully free from rotations

32
3.2. SENSITIVITY ANALYSIS

in all directions as well a fixed boundary condition was implemented in model version
16-2 and 16-3. The basis for this approach is the fact that theoretical design never
fully coincides with practice. A resulting consequence from this difference, is that
the hinge of the arch may have small gaps in the real mechanical hinge which could
allow for some local translation/rotation. The dynamic effect of this may be a local
vibration, that is minimal, but nevertheless existent.

Cross bracing

The effect of the cross bracing is studied in model version 17. The main purpose as
to why its effect is separately evaluated is mainly to gain deeper understanding of
how it affects the transverse mode.

3.2.2 Tuned model

After having conducted the sensitivity analysis, the modal parameters were, as men-
tioned extracted and compared to those obtained from the experiments. By evalu-
ating the results, several changes from the sensitivity analysis were combined into
a so called "tuned model" in order to obtain a FE-model that better reflects the
existing bridge and the experimental outcomes obtained from the experiments con-
ducted during warm weather. An element size of 0.03 m was used, which was only
limited by the maximum amount of nodes the student license for Abaqus allows.
Two model versions were created of this tuned model. The difference between these
models is that the connection between the steel beams and the hangers are made
up of ties with free rotation, while the other one is made up of the rigid bodies
described throughout this method. The version with the tie is denoted as "Tuned
v1", while the version with rigid body is denoted as "Tuned v2". Apart from their
difference at said connection, the tuned models included the following changes:

• Adjusting with springs between steel beams and main beams in the vertical
direction. Applies only for the steel beams attached to the arches

• Include cross bracing

• Use the characteristic value of the stiffness for the glued laminated timber
beams

• Have a fixed-hinge constraint between the hanger and arch

• Include the railing a non-structural mass

• Only allow translations between main beams and deck

33
CHAPTER 3. METHOD

Noted from these should be that the stiffness for GL30c was decreased by applying
the characteristic value instead of the mean value for the material. The reason
for this was that support was found with regards to the behaviour of the material
from tests having been conducted by SP Technical Research Institute of Sweden
on glued laminated timber beams in which their characteristic bending strength
and stiffness was measured, resulting in lower values than the values recommended
(Crocetti (2009)). Moreover, the railing was also included to increase the mass of
the structure and thus, decrease the natural frequencies. All other changes were
further explained above.

34
3.3. EXPERIMENTAL ANALYSIS

3.3 Experimental analysis

Experiments were conducted in order to verify the natural frequencies and mode
shapes obtained from the FE-model as well as for calculating the damping ratio. To
consider the seasonal effects, two experiments were carried out during two separate
occasions, one during the winter season and another during the summer/spring
season. Accelerometers were used as measurement instruments. Excitation was
carried out by applying three different methods
1. Jump tests (output-only), excitation occurs in the vertical direction. Meant
to identify the vertical and torsional modes.
2. Hammer tests (output-input), excitation occurs in both vertical and horizon-
tal direction. Meant to identify all modes. These tests yield the frequency
response function.
3. Ambient vibration tests (AVT) (output-only). Meant to excite all modes.
Furthermore, the position of the accelerometers and the hammer strikes/jumps were
determined by evaluating the maximum deflection from the mode shapes in the FE-
model. These were limited to capture the
• 1st vertical mode,
• 2nd vertical mode,
• 3rd vertical mode,
• 1st transverse mode and
• 1st torsional mode.
Finally, the entire experimental process is performed according to the following
steps,
1. Determine the placement of the accelerometers.
2. Determine the positions in which the hammer/jump tests will occur.
3. Connect all measurement instruments.
4. Conduct 3 hammer tests in one position.
5. Conduct 3 jump tests in the same position as step 4.
6. Repeat step 3 and 4 for all positions determined in step 2.
7. Conduct an ambient vibrations test.

35
CHAPTER 3. METHOD

3.3.1 Measuring process

The measuring process entailed connecting 16 or 17 Clibrys MEMS accelerometers


(depending on which seasonal experiment) and one Dytran 5803A impulse hammer
from Acoutronic to a MGCplus measurement amplifier from HBM connected to a
computer. The amplifier both amplifies and samples the data, which allows for data
to be stored for further software analysis. During the experiments, data is sampled
from the accelerometers as well as from the force cell within the impulse hammer.
Data is then stored and visualized in the computer using the data acquisition soft-
ware (DAQ) Catman Prof from HBM. Figure 3.26 illustrates the process of acquiring
the measurement process.

Computer with
Connection Catman
Retrieved signal
A1-A17 Accelerometer

Measurement Impulse
amplifier hammer

A1 A2 A3 A4 A5 A6 A7 A8 A9 A10 A11 A12 A13 A14 A15 A16 A17

Figure 3.26: Process of acquiring data from experiments.

In order to have two accelerometers measuring in more than one direction on the
same location of the bridge, a heavy block is utilized for mounting purposes, see
Figure 3.27.

Figure 3.27: Illustrative figure of the accelerometers mounted on a block. The picture
is taken during the winter experiment.

36
3.3. EXPERIMENTAL ANALYSIS

3.3.2 Setup: Experiment winter season

The experimental setup is illustrated in Figure 3.28. Moreover, Table 3.3 depicts
the position, placement and the direction each accelerometer measures in and Table
3.4 presents the placement of each hammer and jump test conducted. These tables
compliment the Figure of the setup.

Figure 3.28: Plan view showcasing the setup for the winter experiment.

Table 3.3: Position, direction and placement of each accelerometer.

Accelerometer Measured direction Placement


A1-8 Vertical (z) Deck
A9-13 Transverse (y) Deck
A14 Longitudinal (x) Deck
A15-16 Transverse (x) Hanger

Table 3.4: Position and placement of hammer and jump tests.

Position Placement
Meant to excite
Hammer/Jump Jump/Hammer
V1 1st & 2nd vertical mode Deck
V2 3rd vertical modes Deck
T1 1st torsional mode Deck
H1 1 & 2nd horizontal mode
st
Railing
H2 1st & 2nd horizontal mode Hanger
H3 1st & 2nd horizontal mode Deck

37
CHAPTER 3. METHOD

3.3.3 Setup: Experiment summer/spring season

The summer/spring setup has the additional accelerometers A1, A3 and A8 to cap-
ture the mode shapes in more detail. Apart from that, A14 and A15 were removed
from the winter setup. The reason for the removal was that the conclusion was
drawn that there was no benefit in measuring at those positions and direction. Note
that the numbering for the accelerometers between the two setups differ. The sum-
mer/spring setup is seen in Figure 3.29. Specifics regarding the accelerometers and
the tests carried out compliment the Figure of the setup and is found in Tables 3.5
and 3.6.

Figure 3.29: Plan view showcasing the setup for the summer/spring experiment.

Table 3.5: Postion, direction and placement of each accelerometer.

Accelerometer Measured direction Placement


A1-11 Vertical (z) Deck
A12-17 Transverse (y) Deck
A18 Transverse (x) Hanger

Table 3.6: Postion and placement of hammer and jump tests.

Position Placement
Meant to excite
Hammer/Jump Jump/Hammer
V1 1st & 2nd vertical modes Deck
V2 3rd vertical mode Deck
T1 1st torsional mode Deck
T2 1st torsional mode Deck
H1 1st transverse mode Deck

38
3.3. EXPERIMENTAL ANALYSIS

3.3.4 Data processing

Experimental results were processed in Matlab utilizing the Catman file importer
from Mathworks. Data is transformed from time domain to frequency domain by
applying either the FFT- or Pwelch (PSD) function. Deploying the "peak-picking"
method obtained natural frequencies, a method that assesses the natural frequencies
as the peaks in the frequency domain. Evaluating natural frequencies through both
functions reassures the validity in the extracted frequencies. However, once the
frequencies are verified, further analysis is preformed through exclusively the FFT-
function.
As noise from passing cars disrupts the signal, a 4th order low-pass filter with a cutoff
frequency of 20 Hz is applied by adopting the butter-function in Matlab. Figure
3.30 is an example of the unfiltered and filtered signal in the time domain. The
illustrated example is of a hammer test meant to excite the 1st vertical mode. Note
the difference in amplitude with and without a filter.

A1-time domain A2-time domain A3-time domain A4-time domain


6 2
2 5
A1-time domain A2-time domain A3-time domain A4-time domain
4
acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]

0.04 0.1 0.04 0.04


1
2
acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]
0 0.02 0.02
0.05 0.02
0 0
0
0 0
-2 -2 0 0
-1 -0.02 -0.02
-4
-0.05 -0.02
-4 -5 -2 -0.04 -0.04
0 10 20 0 10 20 0 10 20 0 10 20
t [s] t [s] t [s] t [s] -0.06 -0.1 -0.06 -0.04
0 10 20 0 10 20 0 10 20 0 10 20
A5-time domain A6-time domain A7-time domain A8-time domain t [s] t [s] t [s] t [s]
15 10 5 1
A5-time domain A6-time domain A7-time domain A8-time domain
acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]

10 5 0.3 0.05 0.04


0.5 0.05
acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]

acceleration [m/s2]
5 0
0.2
0 0 0.02
0
0 -5
0.1 0
-0.5
-5 -10 0
-0.05
0
-10 -15 -5 -1
0 10 20 0 10 20 0 10 20 0 10 20
-0.02
t [s] t [s] t [s] t [s] -0.1 -0.1 -0.05
0 10 20 0 10 20 0 10 20 0 10 20
t [s] t [s] t [s] t [s]

(a) Unfiltered signal for accelerometer A1 -


8. (b) Filtered signal for accelerometer A1 - 8.

Figure 3.30: Time domain with and without a low-pass filter.

Consider Figure 3.30b in which the signal evidently still contains noise from external
sources. By zeroing out the accelerations prior to the hammer strike/jump and after
free vibration, the signal is further cleared from unwarranted noise. However, as can
be seen from Figure 3.30b, the time at which free vibrations ends is difficult to
distinguish. Therefore, an iterative process is utilized in which the peaks in the
frequency domain are ensured to remain at the same position as it was prior to
altering the signal.
Another difficulty in the signal can be seen in Figure 3.31, which illustrates the
time and frequency domain for the 6th accelerometer from Figure 3.30b. The time
domain depicts a large peak at the time of the hammer strike which causes a shift
in the frequency domain, a phenomenon that arises for all accelerometers closest to
the hammer strike/jump.

39
CHAPTER 3. METHOD

Time domain A6 Frequency domain A6


0.3 80

0.25 Large peak 70

0.2 60

0.15 50

amp (m/s2)
acc (m/s2)

0.1 40

0.05 30

0 20

-0.05 10

Shift
-0.1 0
0 5 10 15 20 0 2 4 6 8 10

t (s) f (Hz)

Figure 3.31: Shifted frequency domain caused by a large peak in the time domain.

However, zeroing out the peak shifts back the frequency domain, as can be seen in
Figure 3.32.
Time domain A6 Frequency domain A6
0.06 60

0.04
50

0.02
40

0
amp (m/s2)
acc (m/s2)

30

-0.02

20
-0.04

10
-0.06

-0.08 0
0 5 10 15 20 0 2 4 6 8 10

t (s) f (Hz)

Figure 3.32: Response from a cleared signal.

Once the signals have been cleared for the jump and ambient vibration tests, an
FFT is performed and the mode shapes, natural frequencies and damping ratios
were extracted/calculated. For the hammer tests, the frequency response function
is determined and therefrom the mode shape, natural frequency and damping ratio
is calculated/determined. Finally, Figure 3.33 below represents a flowchart that
summarizes the entire process of how data was processed.

40
3.3. EXPERIMENTAL ANALYSIS

Extract hammer Determine the


FFT
Extract natural input FRF
frequencies Hammer
test

Extract natural
FFT + Clear signal
frequencies,
lowpass in time
Matlab domain mode shapes and
Data from Evaluate time filter Evaluate time and Type
damping ratio
experiments domain frequency domain of test

PSD
(Pwelch) Jump test /
Extract natural Ambient
frequencies vibrations Extract natural
frequencies,
mode shapes and
damping ratio

Figure 3.33: Flowchart depicting the data processing.

Determining the mode shape & Damping ratio

The mode shape is the deflected shape of the structure for a particular governing
natural mode/frequency. Applying this in the experimental analysis requires con-
sidering, for instance, the bridge being excited in the vertical direction. To extract
the mode shape, the amplitude and its direction (upwards or downwards) is deter-
mined for the governing mode. The movement of each accelerometer is extracted
and an interpolation between the position of each accelerometer is performed from
which the mode shape is constructed. Furthermore, the damping ratio is extracted
by applying the half-power bandwidth (HPB) method.

Verifying the mode shape

The concept of extracting the mode shape from experiments is dependant on the
accelerometers moving in phase. This is verified by isolating the natural frequency of
interest with a 1st order band-pass filter. Therefrom, the time domain is considered
from which the movement of each accelerometer in relation to one another is seen.

Hammer input in FEM

To reconstruct an experimental test in FEM, the force from a hammer test is im-
ported as a tabular load at the same position along the bridge deck as it was during
the experiments. By applying the step "Modal Dynamics" in Abaqus, the time his-
tory response can be reconstructed for loads that are time-dependant with the use
of modal superposition. The response is evaluated by using the damping ratio calcu-
lated from the experiments for the first 5 modes. In Abaqus, 6 modes are requested
with frequencies ranging from 4.3 Hz to 8 Hz, in order to capture the desired modes.
This method is a way of replicating the hammer input in the numerical model, from
which the accelerations at the same position of a particular accelerometer can be
extracted and then transferred into the frequency domain with the use of an FFT.

41
CHAPTER 3. METHOD

Comparing the FRF:s between the accelerometer measuring the impact from the
experiments and the numerical model is a way of validating the numerical model.

42
Chapter 4

Results

4.1 Experimental results

The experiential results are presented by providing the frequency domains/FRF


(depending on the type of test) at all positions that the tests were conducted (see
Sections 3.3.2 and 3.3.3). This is considered for both the cold weather and warm
weather experiments, respectively. However, since several tests were taken at each
particular position, not all responses are covered in the results. To account for
this, a representative amount of responses are presented to reflect that the outcome
from the frequency domain is repetitive for all accelerometers meant to capture the
particular mode considered. Consequentially, the results serve as a basis that the
considered peaks representing each mode is valid.
Moreover, the extracted natural frequencies and calculated damping ratios are then
presented in Section 4.1.3, covering the comparison between the two experiments. In
the extraction and calculation of these modal parameters, all tests that were deemed
within reason were part of determining the magnitude of each parameter. Subse-
quently, noisy frequency responses were excluded in the determination of damping
ratio since it affects the calculated value. However, the outcome of this is that the
extracted values deviate and consequentially that deviation is employed within the
results. It should be noted that results from the ambient vibration tests are not
included, since the response was poor in comparison to the two other types of tests.
Frequency responses covering the difference between the two tests are then presented.
Note that the 2D and 3D mode shapes from the experiments are presented and
included in the results from the sensitivity analysis in Sections 4.2.2 and 4.2.3,
respectively. As mentioned in Section 3.3.4, the phase angle were distinguished to
verify that the accelerometers measured data simultaneously, these are presented in
Appendix A.

43
CHAPTER 4. RESULTS

4.1.1 Winter experiment: March 5, 2019, -10◦ C

Position V1: FRF - Hammer tests labeled 5,6,7

0.8 2.5 2.5 2


Test 5 Test 5 Test 5 Test 5
Test 6 Test 6 Test 6 Test 6
2 2
0.6 Test 7 Test 7 Test 7 1.5 Test 7

1.5 1.5
0.4 1
1 1
0.2 0.5
0.5 0.5

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

2 2 2.5 2
Test 5 Test 5 Test 5 Test 5
Test 6 Test 6 Test 6 Test 6
2
1.5 Test 7 1.5 Test 7 Test 7 1.5 Test 7

1.5
1 1 1
1
0.5 0.5 0.5
0.5

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.1: FRF vertical accelerometers.

Position V1: Frequency domain - Jump tests labeled 8,9,10

0.6 1 1 1
Test 8 Test 8 Test 8 Test 8
Test 9 Test 9 Test 9 Test 9
Test 10
0.8 Test 10
0.8 Test 10
0.8 Test 10
0.4
0.6 0.6 0.6

0.4 0.4 0.4


0.2
0.2 0.2 0.2

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

1 1 1 1
Test 8 Test 8 Test 8 Test 8
Test 9 Test 9 Test 9 Test 9
0.8 Test 10
0.8 Test 10
0.8 Test 10
0.8 Test 10

0.6 0.6 0.6 0.6

0.4 0.4 0.4 0.4

0.2 0.2 0.2 0.2

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.2: Frequency domain vertical accelerometers.

44
4.1. EXPERIMENTAL RESULTS

Position V2: FRF - Hammer tests labeled 11,12,13,14

1 1 4 2
Test 11 Test 11 Test 11 Test 11
Test 12 Test 12 Test 12 Test 12
0.8 0.8
Test 13 Test 13 3 Test 13 1.5 Test 13
Test 14 Test 14 Test 14 Test 14
0.6 0.6
2 1
0.4 0.4
1 0.5
0.2 0.2

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

2.5 2.5 4 2.5


Test 11 Test 11 Test 11 Test 11
Test 12 Test 12 Test 12 Test 12
2 2 2
Test 13 Test 13 3 Test 13 Test 13
Test 14 Test 14 Test 14 Test 14
1.5 1.5 1.5
2
1 1 1
1
0.5 0.5 0.5

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

Figure 4.3: FRF vertical accelerometers.

Position V2: Frequency domain - Jump tests labeled 15,16,17

80 60 300 150
Test 15 Test 15 Test 15 Test 15
Test 16 Test 16 Test 16 Test 16
60 Test 17 Test 17 Test 17 Test 17
40 200 100

40

20 100 50
20

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

200 200 300 200


Test 15 Test 15 Test 15 Test 15
Test 16 Test 16 Test 16 Test 16
150 Test 17 150 Test 17 Test 17 150 Test 17
200

100 100 100

100
50 50 50

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

Figure 4.4: Frequency domain vertical accelerometers.

45
CHAPTER 4. RESULTS

Position T1: FRF - Hammer tests labeled 29,30,31

0.8 2 1.5 2
Test 29 Test 29 Test 29 Test 29
Test 30 Test 30 Test 30 Test 30
0.6 Test 31 1.5 Test 31 Test 31 1.5 Test 31
1

0.4 1 1

0.5
0.2 0.5 0.5

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

2 2 1.5 2
Test 29 Test 29 Test 29 Test 29
Test 30 Test 30 Test 30 Test 30
1.5 Test 31 1.5 Test 31 Test 31 1.5 Test 31
1

1 1 1

0.5
0.5 0.5 0.5

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.5: FRF vertical accelerometers.

Position T1: Frequency domain - Jump tests labeled 32,33,34

150 250 200 300


Test 32 Test 32 Test 32 Test 32
Test 33 Test 33 Test 33 Test 33
200
Test 34 Test 34 150 Test 34 Test 34
100 200
150
100
100
50 100
50
50

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

250 300 150 250


Test 32 Test 32 Test 32 Test 32
Test 33 Test 33 Test 33 Test 33
200 Test 34 Test 34 Test 34
200 Test 34
200 100
150 150

100 100
100 50
50 50

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.6: Frequency domain vertical accelerometers.

46
4.1. EXPERIMENTAL RESULTS

Position H1: FRF - Hammer tests labeled 11,12,13,14

0.4 0.8 0.8


Test 28 Test 28 Test 28
Test 24 Test 24 Test 24
0.3 Test 25 0.6 Test 25 0.6 Test 25

0.2 0.4 0.4

0.1 0.2 0.2

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

0.6 10 4
Test 28 Test 28 Test 28
Test 24 Test 24 Test 24
8
Test 25 Test 25 3 Test 25
0.4
6
2
4
0.2
1
2

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.7: FRF transverse accelerometers.

47
CHAPTER 4. RESULTS

4.1.2 Spring Experiments: March 17, 2019, 17◦ C

Position V1: FRF - Hammer tests labeled 1,2,3

0.4 1 1 2
Test 1 Test 1 Test 1 Test 1
Test 2 Test 2 Test 2 Test 2
0.2 0.5 0.5 1
Test 3 Test 3 Test 3 Test 3

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

2 4 2 1
Test 1 Test 1 Test 1 Test 1
Test 2 Test 2 Test 2 Test 2
1 2 1 0.5
Test 3 Test 3 Test 3 Test 3

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

2 2 2
Test 1 Test 1 Test 1
Test 2 Test 2 Test 2
1 1 1
Test 3 Test 3 Test 3

0 0 0
0 5 10 0 5 10 0 5 10

Figure 4.8: FRF vertical accelerometers.

Position V1: Frequency domain - Jump tests labeled 19,20,21

100 150 150 300


Test 19 Test 19 Test 19 Test 19
Test 20 100 Test 20 100 Test 20 200 Test 20
50 Test 21 Test 21 Test 21 Test 21
50 50 100

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

300 600 200 100


Test 19 Test 19 Test 19 Test 19
200 Test 20 400 Test 20 Test 20 Test 20
Test 21 Test 21 100 Test 21 50 Test 21
100 200

0 0 0 0
0 5 10 0 5 10 0 5 10 0 5 10

300 600 300


Test 19 Test 19 Test 19
200 Test 20 400 Test 20 200 Test 20
Test 21 Test 21 Test 21
100 200 100

0 0 0
0 5 10 0 5 10 0 5 10

Figure 4.9: Frequency domain vertical accelerometers.

48
4.1. EXPERIMENTAL RESULTS

Position V2: FRF - Hammer tests labeled 4,5,6

1 2 1 1
Test 4 Test 4 Test 4 Test 4
Test 5 Test 5 Test 5 Test 5
0.5 1 0.5 0.5
Test 6 Test 6 Test 6 Test 6

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

2 4 2 0.2
Test 4 Test 4 Test 4 Test 4
Test 5 Test 5 Test 5 Test 5
1 2 1 0.1
Test 6 Test 6 Test 6 Test 6

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

2 4 2
Test 4 Test 4 Test 4
Test 5 Test 5 Test 5
1 2 1
Test 6 Test 6 Test 6

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.10: FRF vertical accelerometers.

Position V2: Frequency domain - Jump tests labeled 22,23,24

100 200 200 100


Test 22 Test 22 Test 22 Test 22
Test 23 Test 23 Test 23 Test 23
50 Test 24 100 Test 24 100 Test 24 50 Test 24

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

400 500 200 40


Test 22 Test 22 Test 22 Test 22
Test 23 Test 23 Test 23 Test 23
200 Test 24 Test 24 100 Test 24 20 Test 24

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

400 500 400


Test 22 Test 22 Test 22
Test 23 Test 23 Test 23
200 Test 24 Test 24 200 Test 24

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.11: Frequency domain vertical accelerometers.

49
CHAPTER 4. RESULTS

Position T1: FRF - Hammer tests labeled 7,8,9

0.4 1 1 2
Test 7 Test 7 Test 7 Test 7
Test 8 Test 8 Test 8 Test 8
0.2 Test 9 0.5 Test 9 0.5 Test 9 1 Test 9

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

2 2 2 1
Test 7 Test 7 Test 7 Test 7
Test 8 Test 8 Test 8 Test 8
1 Test 9 1 Test 9 1 Test 9 0.5 Test 9

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

2 2 2
Test 7 Test 7 Test 7
Test 8 Test 8 Test 8
1 Test 9 1 Test 9 1 Test 9

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.12: FRF vertical accelerometers.

Position T1: Frequency domain - Jump tests labeled 26,27,28

100 200 200 400


Test 26 Test 26 Test 26 Test 26
Test 27 Test 27 Test 27 Test 27
50 Test 28 100 Test 28 100 Test 28 200 Test 28

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

400 400 400 100


Test 26 Test 26 Test 26 Test 26
Test 27 Test 27 Test 27 Test 27
200 Test 28 200 Test 28 200 Test 28 50 Test 28

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

400 400 400


Test 26 Test 26 Test 26
Test 27 Test 27 Test 27
200 Test 28 200 Test 28 200 Test 28

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.13: Frequency domain vertical accelerometers.

50
4.1. EXPERIMENTAL RESULTS

Position H1: FRF - Hammer tests labeled 13,14,15

0.5 1.5 1.5 1.5


Test 13 Test 13 Test 13 Test 13
0.4 Test 14 Test 14 Test 14 Test 14
Test 15 1 Test 15 1 Test 15 1 Test 15
0.3

0.2
0.5 0.5 0.5
0.1

0 0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

0.8 0.3 2
Test 13 Test 13 Test 13
Test 14 Test 14 Test 14
0.6 1.5
Test 15 0.2 Test 15 Test 15

0.4 1

0.1
0.2 0.5

0 0 0
2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

Figure 4.14: FRF transverse accelerometers.

51
CHAPTER 4. RESULTS

4.1.3 Comparison experiments between seasons

Frequency domain

The following Figures, illustrate an FRF from the same accelerometer during winter
and spring. The type of test presented in Figure 4.15a is a hammer test at T1,
depicting all 3 vertical modes and the torsional mode, during both seasons. Figure
4.15b depicts a hammer test at H1 during both seasons and therefore showcases the
transverse mode.

2 0.8

1.8 -10°C 0.7


+17°C
1.6 -10°C
0.6 +17°C
1.4

0.5
1.2

1 0.4

0.8
0.3

0.6
0.2
0.4

0.1
0.2

0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10

(a) Response torsional test. (b) Response transverse test.

Figure 4.15: Frequency response during both seasons.

Frequencies and damping ratio

Table 4.1 depicts the values for the natural frequencies and their respective differ-
ence, whereas Table 4.2 presents calculated damping ratio and the corresponding
difference. Note that the difference is calculated as going from cold to warm tem-
perature.

Table 4.1: Resulting natural frequencies.

Mode Frequency, f , [Hz] Difference [%]


-10C +17C
1st transverse 5.50 ± 0.06 4.36 ± 0.01 -20.7
1st vertical 5.37 ± 0.06 5.13 ± 0.03 -4.5
2nd vertical 6.14 ± 0.02 5.71 ± 0.03 -7
3rd vertical 6.56 ± 0.04 6.18 ± 0.03 -5.8
1st torsional 7.18 ± 0.03 6.71 ± 0.02 -6.5

52
4.1. EXPERIMENTAL RESULTS

Table 4.2: Resulting damping ratios.

Mode Damping ratio, ζ, [%] Difference [%]


-10C +17C
1st transverse 0.83 ± 0.07 2.04 ± 0.21 146
1st vertical 1.02 ± 0.28 1.03 ± 0.13 1.04
2nd vertical 1.23 ± 0.30 1.19 ± 0.37 -3.25
3rd vertical 1.37 ± 0.38 1.08 ± 0.20 -21.2
1st torsional 0.99 ± 0.21 1.17 ± 0.23 18.2

4.1.4 Hammer input in FEM

The hammer input from the same torsional test is showed for two different models
from the sensitivity analyses. The peaks represent, in order, the first vertical mode,
second vertical mode, third vertical mode and the first torsional mode.

3 1.4

Experiments 1.2 Experiments


2.5 FEM: Model version Tuned v1 FEM: Model version 16-1

1
2

0.8
1.5
0.6

1
0.4

0.5
0.2

0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10

(a) Response compared to the model version (b) Response compared to the model version
"Tuned v1". "16-1".

Figure 4.16: Frequency domain from a hammer input from two different models from
the sensitivity analysis.

53
CHAPTER 4. RESULTS

4.2 Sensitivity analysis

The results from the sensitivity analysis only considers the changes that had an
impact on the mode shapes or the natural frequencies. Thus, the results presented
are denoted in accordance to the model versions labeled in Table 3.2. Note that the
label "Initial", refers to the model with tie constraints that was modeled prior to
the sensitivity analysis. The sensitivity analysis only considers the 2D mode shapes.

4.2.1 Frequencies

The following Table presents the difference in percentage between the different mod-
els in the sensitivity analysis compared to the values obtained from the experimental
results from spring season.

Table 4.3: Frequencies compared to the warm weather experiments.

Experiments Initial Base 1 4 7 15 16-1 Tuned Tuned


Mode
17◦ C [Hz] [%] [%] [%] [%] [%] [%] [%] v1 [%] v2 [%]
1st transverse 4.36 21.6 3.1 3.1 1.4 0.0 0.7 3.3 -0.1 2.9
1st vertical 5.13 29.9 23.3 23.3 22.4 17.8 18.1 24.2 1.4 4.6
2nd vertical 5.71 32.3 21.0 21.0 19.8 15.5 16.4 28.8 -0.2 6.1
3rd vertical 6.18 33.9 30.1 30.1 27.3 24.9 25.3 32.1 6.4 8.3
1st torsional 6.71 28.5 28.2 28.2 26.1 22.5 23.0 37.4 4.9 7.5

4.2.2 2D Mode shapes

1 Base model
1
4
0.8
7
15
0.6 16-1
Inital model
0.4 Tuned v1
Tuned v2
0.2 Experiments

-0.2

-0.4

-0.6

-0.8

-1

0 5 10 15 20 25 30 35 40

Figure 4.17: 2D mode shapes depicting the 1st vertical mode.

54
4.2. SENSITIVITY ANALYSIS

1 Base model
1
4
0.8
7
15
0.6 16-1
Inital model
0.4 Tuned v1
Tuned v2
0.2 Experiments

-0.2

-0.4

-0.6

-0.8

-1

0 5 10 15 20 25 30 35 40

Figure 4.18: 2D mode shapes depicting the 2nd vertical mode.

0.8

0.6

0.4

0.2

0
Base model
-0.2 1
4
-0.4 7
15
-0.6 16-1
Inital model
-0.8 Tuned v1
Tuned v2
Experiments
-1

0 5 10 15 20 25 30 35 40

Figure 4.19: 2D mode shapes depicting the 3rd vertical mode.

55
CHAPTER 4. RESULTS

Base model
1
1
4
7
15
16-1
0.8 Inital model
Tuned v1
Tuned v2
Experiments
0.6

0.4

0.2

0
0 5 10 15 20 25 30 35 40

Figure 4.20: 2D mode shapes depicting the 1st transverse mode.

1 Base model
1
4
0.8
7
15
0.6 16-1
Inital model
0.4 Tuned v1
Tuned v2
0.2 Experiments

-0.2

-0.4

-0.6

-0.8

-1

0 5 10 15 20 25 30 35 40

Figure 4.21: 2D mode shapes depicting the 1st torsional mode.

56
4.2. SENSITIVITY ANALYSIS

4.2.3 3D Mode shapes

The 3D mode shapes are presented from the FE-model as well as from the experi-
ments. Note that the black dots in the experimental figures indicate the position of
the accelerometers. The considered mode shapes are from the spring experiments
and from the model version 16-1, in which the rotations at the hinge at the crown
of the arch is restrained.

(a) FE-model. (b) Experiments.

Figure 4.22: 1st vertical mode.

(a) FE-model. (b) Experiments.

Figure 4.23: 2nd vertical mode.

(a) FE-model. (b) Experiments.

Figure 4.24: 3rd vertical mode.

57
CHAPTER 4. RESULTS

(a) FE-model. (b) Experiments.

Figure 4.25: 1st transverse mode.

(a) FE-model. (b) Experiments.

Figure 4.26: 1st torsional mode.

58
Chapter 5

Discussion

5.1 Experiments

After reviewing the frequency domains from the two experiments, it is evident that
results from the winter experiment was noisier than that from the spring experi-
ments, and thus more difficult to analyze. The reason for this was both a difference
in traffic volume, but also a great difference in wind speed. When regarding Figures
4.1 and 4.2 it can clearly be seen that the peak representing the 1st vertical mode
(at about 5.50 Hz) was extremely small in amplitude and was from that experiment
not concluded to be a vertical mode. However, by reviewing Figures 4.8 and 4.9,
3 peaks corresponding to the 3 first vertical modes are clearly identified, the first
being at about 4.4 Hz. Thus, concluding that the bridge had 3 vertical modes below
10 Hz was discovered after the spring experiments.
Moreover, the frequency domains corresponding to the response from the accelerom-
eters measuring in the transverse direction clearly indicates a distinguished peak at
about 2.5 Hz (see Figures 4.7 and 4.14). However, regarding the response from
the accelerometers capturing the behaviour of the arch, the same peak reappears
but with amplitudes greater than those from the accelerometers on the deck. After
thorough investigations, along with reviewing the same mode from the FE-model,
conclusions were drawn that that particular mode is a local mode for the arch,
and thus the behaviour is reflected as the arches moving the deck, rather than the
opposite. Therefore, that particular mode was not considered in the results.
Furthermore, studying the results from the experimental analysis clearly indicates
a correlation between temperature and the natural frequencies, in which increased
temperature causes decreased natural frequencies. This outcome reflects those from
the literature review. The results also indicate that the mode having the biggest
seasonal effect is the 1st transverse mode, decreasing by 21%, as opposed to the
1st vertical mode that only decreased by mere 5 %. Concluded from the literature
review can be that the seasonal effect due to the asphalt layer most probably exists,
and that its effect mainly occurs in the transverse direction. This outcome is fairly
logical since the asphalt mainly interacts with the deck plate through shear transfer.

59
CHAPTER 5. DISCUSSION

However, by looking at the damping ratio, determining any type of correlation due
to change in temperature was not possible. Instead, the results differed depending
on which mode was considered since some modes decreased and others increased in
magnitude. It is however evident that the greatest impact the seasonal effect had
on the damping ratio was once again on the 1st transverse mode, that more than
doubled going from cold to warm temperature. A result that once again verifies that
the effect from the asphalt mainly occurs through shear transfer.
When considering recommended values regarding the damping ratio, Hägernäs bridges
has slightly lower values for one particular mode, in comparison to the recommended
values of 1-1.5 %, depending on the type of joints in the structure. Since the struc-
ture is glued between the deck plate and the grid made up from glued laminated
beams, but consists of various mechanical joints at other connecting parts of the
structure, concluding which value from the Eurocode to consider is difficult. How-
ever, the 1st transverse mode, with a damping ratio of 0.83 % during winter, fails
when regarding both values. On the other hand, its seasonal effect was large and
during spring the values were above recommendation. Since the bridge is meant
solely for pedestrian use, the loading condition during winter is most probably lower
than that during spring, which in this case can be considered favourable. Moreover,
the 1st design criteria, which regulates the natural frequencies, is clearly fulfilled,
even when regarding the results during the warm weather conditions.
During the thesis, a similar project numerically and experimentally dynamically as-
sessing a similar structure (namely the Vega bridge) was ongoing. There, the author
also concludes that natural frequencies decrease with increased temperature (Neil-
son (2019)). As for the Hägernäs bridge, Vega bridge concluded that the seasonal
effect is largest for the 1st transverse mode. The results from calculating the damp-
ing ratio varied depending on whether the results were determined using the jump
tests or hammer tests. However, the results from the Vega bridge follows those for
the Hägernäs bridge in which the damping ratio increases or decreases depending on
which mode to consider. However, the values in terms of damping ratio were a lot
greater for the Vega bridge, and thus it is concluded that the damping is case spe-
cific. Thus, from a designers point of view, the recommendation set in the Eurocode
should be followed.

5.2 Sensitivity analysis

The results from the FE-model suggests that the bridge is stiffer than the exper-
imental results, which is to be expected, since Finite Element Modelling tend to
overestimate the capacity.
The transverse mode was largely unaffected by the sensitivity analysis, as can be
seen in Figure 4.20. The main reason for this is that the deck has such a high modal
mass, in relation to the arches, which results in the arches having a small influence
on the deck. An example of how this looks is illustrated in Figure 4.25a.

60
5.2. SENSITIVITY ANALYSIS

By modifying its parameters, the bridge behavior and the cause of this behavior
was identified. It was noted that the vertical modes are mostly affected by the
behavior of the arch. The parameters that affected the out of plane behavior of
the arch were the most significant, since the arch moves transversely during the
vertical modes of the deck. This is illustrated in Figures 4.22-4.24. Note that since
the bridge is asymmetrical, regular sinus shaped modes are not to be expected in
the vertical direction. Moreover, changing the boundary conditions of the glued
laminated timber beams, in model version 1, did not have an effect on the mode
shapes, or the frequencies. This supports the above mentioned arch dependant
behavior.
As expected, modifying the stiffness did have slight impact on the frequencies, but
the mode shape remained the same (model versions 7 & 15). However, changing
the material property of GL30c, which is the material used in the beams and the
arch, had a great impact on the eigenfrequencies. Since the beams are split, a
reduced characteristic value of the stiffness were implemented in ”Tuned model”.
The result from this was a mode shape somewhat accurate to real life experiments,
and frequency values coinciding within reason. Regarding the tuned model, it was
also noticed that the mode shaped varied with type of hinge used between the
hangers and the main beams (the only difference between "Tuned model v1" and
"Tuned model v2"). Using a hinge with a rigid body opened up the possibility to
put the hinge directly at the steel beams, as described in Section 3.1.4. Another
approach, which resulted in a more accurate mode shape, was to use a tie in which
all rotation was allowed (model version denoted Tuned v1). The reason for this is
that the hinge created with the tie automatically assigns the center of rotation as
the mid span of the distance between the nodes. It should be noted that "Tuning" a
model is most often needed as is evident from the publication by Cruz et al. (2009).
The model version 16-1, restraining the rotations at the crown of the arches resulted
in the best mode shapes by far, which can be seen in Section 4.2.2. Consequentially,
this would imply that the hinge at the crown is not behaving like a hinge, which could
be a result of friction forces. It should be noted that there is a great uncertainty
in this, since there is no documented evidence of such. Restraining the rotations
at the crown also resulted in a much stiffer model, as expected, which results in
frequencies well above the experimentally obtained values (at most by 37.4% which
was for the torsional mode). Taking into regard the report about the split beams,
and implementing underestimated values of stiffness into the model, resulted in
frequencies that still slightly overestimates the frequencies. There is however no
indication that the split beams would in reality have a elastic modulus of 8 GPa.
The elastic modulus would instead be closer to 11 GPa, as stated by the report.
Recalling that the FE-model overestimate the frequencies, and actions taken to lower
these values included both changing material properties, connections and adding
mass in terms of railing resulted in a model which still overestimated the frequen-
cies. Thus, it can be concluded that even by considering all these changes from a
designers point of view, modelling this structure and evaluating its resulting frequen-
cies is slightly dangerous. It can also be concluded that since several parameters
were modified to reach similar values in terms of frequency, having a model that

61
CHAPTER 5. DISCUSSION

reflects the real dynamic bridge behaviour requires in dept evaluations, in partic-
ular concerning the connections to the arch, especially those between timber and
steel. Having this in mind, the results from the literature review, concerning timber
bridges that had been similarly evaluated and modeled, similar issues arise. For
instance, in (Rønnquist and Wollebaek (2019)) the authors had to modify material
properties (mainly stiffness) to conclude values that are similar to those from their
experiments. Thus, the conclusion that the material properties are lower than those
from the manufacturing catalogues is strengthened.

5.2.1 Hammer input in FEM

As mentioned in Section 3.3.4, modelling the hammer impact in to the FE-model is a


great way to reinforce the validity of the model. Restraining the hinge at the crown
of the arch is clearly not a viable approach, as can be seen from the frequency domain
in Figure 4.16b. The figure to the left shows that the tuned model is fairly accurate
for the first two vertical modes. However for the third vertical and the first torsional
mode, the amplitude is greatly deviating from the experimental results. Part of the
difference for this error is the fact that FE-models tend to be more inaccurate with
higher modes. The main reason is that the complexity of the connections could not
be fully recreated in the FE-model which would suggest that further knowledge of
the dynamic behavior should be obtained.

5.2.2 Modelling the asphalt

Considering the asphalt and consequentially its seasonal effect in the FE-model re-
quires having a model that yields lower values than those measured during spring
(when the asphalt modelled as an added mass). This was, however, not the case for
any of the models evaluated and thus modelling the asphalt and thereby introducing
even more stiffness within the model is futile. When also regarding the connections
and their influence on the dynamic behaviour (in particular those concerning the
arch), it is evident that the structure is far too complex to conclude that the as-
phalt is the only source causing seasonal changes. This can be concluded without
even considering temperature as a factor in the FE-model. Another prerequisite for
modelling the asphalt would be measuring the temperature of the asphalt during
both experiments, as was included in the studies considered in the literature review,
and not conducted during the experiments presented in this thesis.

62
Chapter 6

Conclusions

In the following sections, all conclusions are presented along with certain aspects to
be considered with regards to future work. It should be noted that the conclusions
are case-specific, and may only be considered somewhat general if it complies with
the literature review.

6.1 Experiments
• The magnitudes of the natural frequencies and damping ratios were affected
by temperature.

• The natural frequencies decreased with increasing temperature.

• It was impossible to determine any correlation between the change in temper-


ature with the damping ratio.

• Both in terms of frequency and damping, the temperature effect was greatest
for the 1st transverse mode - suggesting that the asphalt effects that mode in
particular.

• The Eigenfrequency condition is fulfilled (CEN (2001)).

• The 1st transverse mode has slightly lower damping ratio than the recom-
mended values for wooden bridges in which the damping ratio has yet to be
verified. However, all other modes are above recommendation.

63
CHAPTER 6. CONCLUSIONS

6.2 FE-model
• The connections proved to be the biggest influence factor on the bridge.

• Uncertainties in the connections between steel and timber, and steel to steel
need to be dynamically assessed, before an accurate FE-description can be
made.

• Arches are the main influencing structural component, and dependent of the
connections.

• Modelling the asphalt layer is difficult, since the extent of shear transfer be-
tween the asphalt and the LVL plate is unknown.

• Laboratory tests are preferable in order to quantify the shear transfer between
the asphalt and LVL plate.

6.3 Future work

6.3.1 Connections

Since the connections were vital in the dynamic behaviour of the structure, in par-
ticular the mode shapes, further works should entail in depth analyses of those.
The main connections to consider are those considering the arch and subsequently
those affecting the arch behaviour. For instance, locally assessing the behaviour of
these connections during dynamic excitation may be of interest. In terms of the
modelling aspect, the connections could be modelled as separate parts (steel plates,
washers, screws etc.) in which their respective contribution on stiffness can be con-
sidered. This approach would be extremely detailed but would remove all ties and
rigid bodies from the model. By doing so, the stiffness contribution each tie has,
depending on whether it is a node-to-node constraint with a certain distance or a
surface-to-surface connection, could be removed. After all, this contribution is not
with regards to the stiffness caused from the connection but rather an outcome of
using the most simple way of modelling. Finally, it should be mentioned that the
steel and timber connections, as strengthened by the literature review, should be
further studied for a deeper understanding.

6.3.2 Modelling the asphalt layer

As mentioned in the discussion, as well as being a part of the literature review,


the asphalt layer is most likely to be the main contributing factor as to why the
seasonal effect occurs. However, after conducting the sensitivity analysis, and con-
cluding that the connections were vital in design, without even considering the
temperature, modelling the asphalt should be a part of a more extensive research

64
6.3. FUTURE WORK

process when considered in future work. To capture the effect of the asphalt, fur-
ther experimentation should be conducted in which the asphalt temperature should
be measured, and if possible, experiments in which the asphalt behaviour can be
regarded locally should be assessed. To get a full understanding of the effect the
asphalt has, modelling it should entail modelling its effect during several occasions
for which experiments have been conducted as well. By applying this process, the
effect of the asphalt could be regarded from for a set of temperatures ranging over
a large temperature span. This effect could then be modelled and the evaluated
results compared with the FE-model in which the same temperature conditions are
present.
Furthermore, it should be mentioned that the effect of the asphalt is more of an inter-
esting subject from an academic perspective since the designers aims to capture the
lower range of the frequency spectra and still fulfill design criteria. Thus, considering
the asphalt and its effect is more of an interest from an academic perspective.

65
Bibliography

CEN, 2001. Eurocode - Basis of structural design. Brussels.

CEN, 2004. Eurocode 5: Design of timber structures part 2: bridges. Brussels.

Crocetti, R., 08 2009. Strength of split glulam beams. p. 10.

Cruz, P. J. S., Salgado, R., Branco, J., 2009. Dynamic analysis and structural eval-
uation of gois footbridge. Mecânica Experimental 17, 129–137.

Cunha, A., Caetano, E., 06 2006. Experimental modal analysis of civil engineering
structures. Sound and Vibration 40.

EEA, 2015. Intensified global competition for resources (gmt 7).

Feltrin, G., Schubert, S., Steiger, R., 10 2011. Temperature effects on the natural
frequencies and modal damping of timber footbridges with asphalt pavement.

Gsell, D., Schubert, S., Feltrin, G., 2010. Influence of asphalt pavement on damp-
ing ratio and resonance frequencies of timber bridges. Engineering Structures 32,
3122–3129.

Gülzow, A., Steiger, R., Gsell, D., Wilson, W., Feltrin, G., 2015. Dynamic field
performance of a wooden trough bridge. EMPA - Swiss Federal Laboratories for
Materials Science and Technology, 441–450.

Hamm, P., 2007. Ein beitrag zum schwingungs- und dämpfungsverhalten von fuss-
gängerbrücken aus holz. Report 2007, Civil Engineering, TU München, Germany.

He, J., Fu, Z.-F., 2001. Modal Analysis. Reed Educational and Professional Pub-
lishing Ltd, 225 Wildwood Avenue, Woburn.

Neilson, J., 2019. Numerical and experimental dynamic analyses of the vega pedes-
trian bridge including seasonal effects. M.Sc. Thesis, Royal Institute of Technology,
Stockholm, Sweden.

Pousette, A., Fjellström, P.-A., 03 2009. Broar av trä.

Rønnquist, A., Wollebaek, L., 05 2019. Dynamic behavior and analysis of a slender
timber footbridge.

67
BIBLIOGRAPHY

Sétra, 2006. Technical guide Footbridges - Assessment of vibrational behaviour of


footbridges under pedestrian loading. 28 rue des Saints-Pères - 75007 Paris -
France.

Zivanovic, S., Pavic, A., Reynolds, P., 03 2007. Finite element modelling and updat-
ing of a lively footbridge: The complete process. Journal of Sound and Vibration
301, 126–145.

Čecháková, V., Rosmanit, M., Fojtik, R., 12 2012. Fem modeling and experimental
tests of timber bridge structure. Procedia Engineering 40, 79–84.

68
69
APPENDIX A. PHASE ANGLES

Appendix A

Phase angles

10-3
Time domain
A1
3 A2
A3
A4
2 A5
A6
A7
1
A8
acceleration [m/s2]

-1

-2

-3

11.7 11.8 11.9 12 12.1 12.2

t [s]

(a) From cold weather experiments.


10-3
Time domain
3
A1
A2
A3
A4
2
A5
A6
A7
1
A8
acceleration [m/s2]

A9
A10
A11
0

-1

-2

9.45 9.5 9.55 9.6 9.65 9.7

t [s]

(b) From warm weather experiments.

Figure A.1: Vertical accelerometers.

70
10-4
Time domain
3 A9
A10
A11
A12
2
A13
A14
A15
1 A16
acceleration [m/s2]

-1

-2

-3

11.9 11.95 12 12.05 12.1 12.15 12.2 12.25 12.3 12.35

t [s]

(a) From cold weather experiments.


10-3
Time domain
1.5
A9
A10
A11
1
A12
A13
A14
A15
0.5 A16
acceleration [m/s2]

-0.5

-1

-1.5

10.5 10.55 10.6 10.65 10.7 10.75 10.8 10.85 10.9

t [s]

(b) From warm weather experiments.

Figure A.2: Transverse accelerometers. Note that A14 measures in the longitudinal
direction and that A11 was turned off.

71
Appendix B

Technical drawings

73
TRITA TRITA-ABE-MBT-19473

www.kth.se

You might also like