Sie sind auf Seite 1von 6

Available online at www.sciencedirect.

com

ScienceDirect
Available online at www.sciencedirect.com
Procedia Engineering 00 (2017) 000–000
www.elsevier.com/locate/procedia
ScienceDirect
Procedia Engineering 199 (2017) 2808–2813

X International Conference on Structural Dynamics, EURODYN 2017

Vibration effects of loads due to groups crossing a lively footbridge


Paweł Hawryszkówa,*, Roberto Pimentelb, Felipe Silvac
a
aWrocław University of Science and Technology, Wybrzeże Wyspiańskiego 27 St., Wrocław 50-370, Poland
b
Universidade Federal da Paraíba, Departmento de Engenharia Civil – Campus Universitário sn, João Pessoa 58051-900, Brazil
Universidade
c
Universidade Federal da Bahia, Faculdade de Arquitetura, Rua Caetano Moura, 121, Salvador 40210-905, Brazil

Abstract

The suitability of the group load for vibration serviceability of footbridges in vertical direction is analysed. The rationale is that
the more pedestrians are on a footbridge the more the actual effect of the presence of pedestrians as dynamic systems is relevant
for the vibration response. Therefore, it is expected that, by increasing the number of pedestrians, representing pedestrians only
as loads would diverge from reality. To investigate this matter, measured responses from controlled tests of crossings of single
and of groups of pedestrians on a two span cable-stayed footbridge were compared to respective results from simulations using
a calibrated finite element (FE) model of the structure. The step frequency was controlled by a metronome, to reduce variability
such as getting out of rhythm or differences of phase among pedestrians walking in group. In the 3-D FE model, the loads were
represented by means of classical Fourier Series since the pacing rate was controlled in all tests. First, the results of the single
crossings of each pedestrian were employed to obtain the dynamic load factor for that pedestrian. In sequence, the same
comparison was made between the measured and simulated responses of the group. As the level of acceleration increased when
comparing the results from group with the individual crossings, the effect of the presence of pedestrians as dynamic systems were
more significant in the former, providing evidence that modeling the pedestrians as dynamic systems is needed for this load case.
© 2017 The Authors. Published by Elsevier Ltd.
© 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the organizing committee of EURODYN 2017.
Peer-review under responsibility of the organizing committee of EURODYN 2017.
Keywords: Footbridge; Vibrations; Experimental tests; Theoretical analyses.

1. Introduction

The historical development of vibration serviceability recommendations for footbridges in the vertical direction
focused first on a notional load produced by a single pedestrian crossing the structure in resonance condition with

* Corresponding author. Tel.: +48-71-320-45-62; fax: +48-71-320-35-45.


E-mail address: pawel.hawryszkow@pwr.edu.pl

1877-7058 © 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the organizing committee of EURODYN 2017.

1877-7058 © 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the organizing committee of EURODYN 2017.
10.1016/j.proeng.2017.09.565
Paweł Hawryszków et al. / Procedia Engineering 199 (2017) 2808–2813 2809
2 Author name / Procedia Engineering 00 (2017) 000–000

the fundamental mode of the structure. The British code BS 5400 [1] was one of the first that presented such
recommendations. Studies on the subject evolved and an upgrade of the codes took place about ten years ago,
in which the design case moved from a single load to loads due to several pedestrians.
The UK NA to BS EN 1991-2 [2] is one of the codes that specify a stream of pedestrians and a group
of pedestrians exciting the footbridge, as part of the vibration serviceability checks. Regarding the group load, its
size varies according to structure usage, ranging from 2 pedestrians for footbridges in rural areas to 16 pedestrians
in heavily used urban areas. The code also adopts a reduction factor to consider the unsynchronized combination
of actions in a pedestrian group when defining the total load to be applied. However, the pedestrians continued
to be considered only as loads and no account is taken of pedestrian-structure interaction.
Indeed, as has been shown from several studies, the action of pedestrians is not only as loads. A collection of such
studies can be found elsewhere [3]. A pedestrian is a full dynamic system, vibrating together with the structure.
Several models to represent the pedestrian as a dynamic system in vertical direction have been proposed, ranging
from adapted inverted pendulum [4], bipedal model [5] to single-degree of freedom models [6-15]. It should be
noted that the design condition of several pedestrians crossing a structure in resonance condition favors the
consideration of the pedestrians as dynamic systems, since the more pedestrians are on the structure at a given time
the more influential are their dynamic properties on the response of the vibrating system.
In this paper, an investigation is carried out regarding the potential effect of pedestrian-structure interaction
in vertical direction for the group load, taking as a case study a footbridge that presented its fundamental vertical
frequency within the usual frequency range of pacing rates of pedestrians. A calibrated 3-D finite element (FE)
model of the structure was developed and the dynamic load factor (DLF) of each test subject who took part in the
test campaign was obtained by adjusting the numerical to the experimental responses from individual crossings
of that test subject. In all numerical analyses, the pedestrian was modeled as a load, using the classical expression
of the Fourier series to describe it and taking as inputs the weight, velocity and pacing rate from the measurements.
In sequence, numerical and experimental results were compared for the crossings of the group. Since the level
of excitation is high for the group, it is expected that differences between numerical and experimental results arise,
due to the lack of human-structure interaction (HSI).

2. Methodology

2.1. Test structure

The investigated “Zlotnicka” footbridge (Fig. 1) is located in Wroclaw, Poland [16]. It is a sensitive structure
to human-induced vibrations due to the light, steel deck and the first frequency of vertical vibrations corresponding
to usual pacing rate of walking pedestrians.
Superstructure of the cable-stayed footbridge consists of a two span continuous steel deck supported by twenty
four stay cables type 2T15 connected to a steel pylon. The deck is made of two steel tube girders Ø323.9/12.5 mm,
connected by steel cross beams HEB 140 mm with spacing 2.00 m. Six longitudinal beams made of HEB 100 mm
with spacing varying from 0.40 to 0.50 m are placed on the cross-beams and a 12 mm thick steel plate is welded
to these beams. All steel elements are made of 18G2A steel (with strength of 280 MPa). The walk width is 3.00 m.
The length of each span is 34.00 m.

2.2. Research programme

The research programme consisted of a large number of Single Pedestrian Tests (SPT) and Group Pedestrian
Tests (GPT). Ten pedestrians took part in the test campaign. Crossings of each pedestrian and whole group
of pedestrians were repeated 10 times. 120 crossings were carried out – 100 SPT crossings with metronome,
10 GPT-F free crossings without metronome (not investigated here) and 10 GPT-R resonant crossings with
metronome. Pedestrians’ masses and crossing times were recorded.
Seven induction accelerometers by HBM GmbH (type B12/200), with SPIDER8 amplifier and CATMAN
program were used (Fig. 1). One of the accelerometers was attached to a pedestrian to help identifying start and
entire time of the crossing (see Fig. 2). The metronome was used to guarantee the proper pacing rate of a single
2810 Paweł Hawryszków et al. / Procedia Engineering 199 (2017) 2808–2813
Author name / Procedia Engineering 00 (2017) 000–000 3

pedestrian or a group of pedestrians. Special, “military” arrangement of pedestrians in group crossings was applied
to enhance synchronization (Fig. 2). The group leader, equipped with metronome, was the front pedestrian. The rest
of the group followed him and was organized in a matrix order (3 rows × 3 columns).

1.20
3.39 Accelerometers HBM B12/200
3.00
Deck plate Stay cables 2T15
1.10

a6 12 mm thick a2 in HDPE protection tubes

12.80
Tube 323.9/12.5
a4,a5 a3 a2,a6 a1
3.00

16.00 18.00 10.00 8.00 8.00 8.00


4.00 4.00 4.00 4.00 4.00 8.00 8.00 4.00 4.00 4.00 4.00 4.00

2.25
L = 34.00 m L = 34.00 m
68.00

Fig. 1. Test structure and location of measurement points.

Fig. 2. Examples of Single and Group Pedestrian Tests.

The first natural frequency was identified before the tests and was equal to f1 = 2.07 Hz. The damping ratio was
determined using the logarithmic decay procedure and its mean value was equal to ξ = 0.43%. It should be reported
that dynamical features of the footbridge had changed since the last test campaign [16] – the frequency value was
greater than before (probably due to the weather conditions or structural changes). It should be also noted, that the
damping level is significantly dependent on the number of pedestrians loading this structure – the damping ratio
ξ = 1.19% was determined in [17], when 10 people were located in antinode section of the first modal shape.

2.3. Computational model

The computational model was prepared in Ansys software (Fig. 3). It was very detailed and included secondary
elements such as handrails. This level of accuracy was required to achieve appropriate calibration between model
and the real footbridge. For all bars the 2-node linear beam188 Ansys bar element was adopted. This element
is based on the Timoshenko beam theory, having 6 degrees of freedom in each node. For the steel plate deck surface
the 4-node linear shell181 element was employed, with 6 degrees of freedom in each node. The shell element size
was 0.16 m wide by 0.66 m long. For all stayed-cables the 2-node linear link180 element was employed, with
3 degrees of freedom in each node.
The load, modeling the pedestrian, included the weight and the first harmonic in a Fourier series representation.
It was considered virtually moving with a constant velocity and having its position updated in every time step of the
dynamic analysis. Equivalent nodal forces were calculated in the element that the pedestrian is at a given time. Since
the weight and velocity were measured, the remaining load parameter to be determined for each pedestrian was the
DLF of the first harmonic.
Paweł Hawryszków et al. / Procedia Engineering 199 (2017) 2808–2813 2811
4 Author name / Procedia Engineering 00 (2017) 000–000

Fig. 3. Computational model and calculation results of the first and the second natural frequency of vertical vibrations.

3. Results

First, the DLF of each test subject was obtained from each of the ten single crossings each subject performed.
After observing some filtered experimental responses and the variation in shape they presented for the same
pedestrian, a decision was made of obtaining the DLF of the pedestrian from each crossing by taking the RMS
acceleration during the crossing as the metric, that is, the DLF was adjusted in order for the numerical RMS matches
the experimental counterpart. This was implemented in the following way: first, an initial numerical RMS
acceleration was obtained for a chosen DLF (e.g. from the literature); in sequence and since the model is linear, the
adjusted DLF was calculated by multiplying the chosen DLF by the ratio of the experimental and initial numerical
RMS acceleration. The obtained averaged DLFs for each pedestrian are shown in Table 1. A plot of experimental
and simulated crossings of a pedestrian is shown in Fig. 4a, after adjusting the DLF of the pedestrian.

a. b.

Fig. 4. Comparison of experimental and simulated crossings of a single pedestrian for different damping ratios
(scheme SPT-1-2: 1st pedestrian, 2nd crossing; accelerometer a2).

Table 1. DLFs and standard deviations for the pedestrians that participated in the test campaign.
Pedestrians Mean DLF Standard deviation Pedestrians Mean DLF Standard deviation
Pedestrian 1 0.10 0.009 Pedestrian 6 0.13 0.012
Pedestrian 2 0.12 0.010 Pedestrian 7 0.11 0.009
Pedestrian 3 0.10 0.008 Pedestrian 8 0.11 0.010
Pedestrian 4 0.10 0.009 Pedestrian 9 0.11 0.012
Pedestrian 5 0.10 0.010 Pedestrian 10 0.17 0.019

It should be noted that the values obtained for the DLFs are far below the values recommended in the literature
from walking in rigid surfaces and for the respective pacing rate. This difference can be understood together with
the comparison shown in Fig. 4a between the numerical and experimental response: the long walking path
associated to a central node and a vibrating surface may had caused the pedestrian to go out of phase with the
structure´s vibration. In addition, pedestrian-structure interaction may also be present to some extent and affected
the overall dynamics of the system. All of these factors were compensated in the numerical calculations by obtaining
2812 Paweł Hawryszków et al. / Procedia Engineering 199 (2017) 2808–2813
Author name / Procedia Engineering 00 (2017) 000–000 5

an equivalent DLF leading to the same RMS acceleration during the crossing. Values of the DLFs and standard
deviation shown in Table 1 presented some uniformity for the ten test subjects who took part in the test campaign
and evidenced that this was an adequate approach for the problem.
Furthermore, there was not a sort of a typical measured crossing. However, the experimental responses always
reached its first peak before the respective simulated response. This was also observed by Caprani et al. [18], which
explained this phenomenon by the influence of the pedestrian on the damping of the system. That is, this is a human-
structure interaction (HSI) effect which is not considered in the simulation. This effect can be confirmed
by observing Fig. 4b, in which the damping ratio of the mode was increased three times in the simulations, the
consequence being a significant reduction in the mismatch in time between measurements and simulations. The
reduction in the acceleration amplitudes of the simulation in Fig. 4b was expected since the DLF of the pedestrian
was kept the same in both plots.
In sequence and keeping the DLF of the pedestrians as originally obtained (see Table 1), as well as the damping
of the empty footbridge, a comparison was made between measured and simulated results for the group loading.
It should be noted that a metronome was employed to enhance synchronization, together with the military formation
previously discussed. A video camera with a resolution of 30 frames/s helped to follow the pedestrian movements,
particularly at about an extension of one quarter of the span closest to the abutment the camera was placed. Among
the ten crossings carried out, four were discarded due to significant occlusions that prevented phase differences
among the pedestrians to be identified by analyzing the images. It was possible to confirm that the pedestrians kept
the same (and unique) pacing rate and also to identify phase angles among the pedestrians. These phase angles were
introduced into the applied loads of the respective simulations. A plot is shown in Fig. 5 as a comparison
of measured and simulated results. In Table 2, all RMS values obtained for the six useful crossings are presented.

Fig. 5. Time response and respective spectrum of a crossing of the group (scheme GPT-R-1: 1st crossing; accelerometer a2).

Table 2. RMS values for the group crossings (accelerometer a2).


Crossing Simulated RMS (m/s2) Measured RMS (m/s2) Crossing Simulated RMS (m/s2) Measured RMS (m/s2)
1 0.8037 0.3059 4 0.6867 0.3565
2 0.6369 0.3487 5 0.7661 0.3530
3 0.6948 0.3436 6 0.6369 0.3497

It can be seen the experimental results presented good consistency/repeatability in each direction among the
crossings. This assures that the test subjects managed to walk together maintaining the same walking pattern. It can
also be seen (see Fig. 5) that the time shift between numerical and experimental results increased when compared
to the case of single crossings. This evidenced that an increase of damping (among other HSI effects) was more
significant for the group loading than for the individual crossings. In addition, simulations presented higher RMS
values than the respective experimental counterparts, without changing the DLF of each pedestrian previously
obtained. Overconservative estimations of the RMS acceleration resulted from the analysis of group loadings, even
for calibrated DLFs of individual pedestrians obtained from single crossings in the same test structure.
Paweł Hawryszków et al. / Procedia Engineering 199 (2017) 2808–2813 2813
6 Author name / Procedia Engineering 00 (2017) 000–000

4. Conclusions

Single crossings and crossings of a group of the same ten test subjects walking across a lively footbridge were
carried out, together with simulations, to investigate human-structure interaction effects in vertical direction under
resonance condition. The structure presented a fundamental vertical frequency within the range of usual pacing rates
of pedestrians and the help of a metronome made vibration levels to increase by about tenfold between single
crossings and crossing of the well synchronized group. Evidence of significant human-structure interaction effects
was observed both in the single and the group crossings. This was shown to cause a mismatch in time between
simulated and measured time responses due to an increase in damping during the crossing, when compared to the
damping of the empty footbridge. Interaction between pedestrians and structure was shown to be much significant
for the group load, leading to significant overestimation of vibration effects by employing the force model, even
when the dynamic load factor (force model) of the pedestrians were calibrated from the single crossings, by crossing
the same footbridge. Modeling the pedestrians as dynamic systems is recommended for this design case.

Acknowledgements

This paper is a result of Joint Research Work “Effects of modeling pedestrians as biodynamic systems”,
conducted in the international cooperation between researchers of universities in Poland and Brazil in years
2014-2017.

References

[1] British Standards Institute, BS5400 British Standards - Steel, concrete and composite bridges: specification for loads, Part 2, Appendix C,
London, 1978.
[2] British Standards Institute, UK National Annex to Eurocode 1: Actions on structures – Part 2: Traffic loads on bridges, NA to BS EN 1991-
2:2003, London, 2008.
[3] E. Shahabpoor, A. Pavic, V. Racic, Interaction between walking humans and structures in vertical direction: a literature review, Shock and
Vibration (2016) paper ID 3430285.
[4] M. Bocian, J.H.G. Macdonald, J.F. Burn, Biomechanically inspired modeling of pedestrian-induced vertical self-excited forces, Journal
of Bridge Engineering, 18(12) (2013) 1336-1346.
[5] J.W. Qin, S.S. Law, Q.S. Yang, N. Yang, Finite element analysis of pedestrian bridge dynamic interaction, Journal of Applied Mechanics
81(4) (2014) paper ID 041001.
[6] P. Fanning, P. Archbold, A. Pavic, A novel interactive pedestrian load model for flexible footbridges, in: SEM annual conference, 2005.
[7] F.T. Silva, R.L. Pimentel, Biodynamic walking model for vibration serviceability of footbridges in vertical direction, in: 8 th International
Conference on Structural Dynamics EURODYN, Leuven, Belgium, 2011, 1090-1096.
[8] C.C. Caprani, J. Keogh, P. Archbold, P. Fanning, Characteristic vertical response of a footbridge due to crowd loading, in: 8 th International
Conference on Structural Dynamics EURODYN, Leuven, Belgium, 2011, 978–985.
[9] F.T. Silva, H.M.B.F. Brito, R.L. Pimentel, Modeling of crowd load in vertical direction using biodynamic model for pedestrians crossing
footbridges, Canadian Journal of Civil Engineering 40(12) (2013) 1196–1204.
[10] M. Pfeil, N. Amador, R. Pimentel, R. Vasconcelos, Analytic-numerical model for walking person-footbridge structure interaction,
in: 9th International Conference on Structural Dynamics EURODYN, Porto, Portugal, 2014, 1079-1085.
[11] M.A. Toso, H.M. Gomes, F.T. Silva, R.L. Pimentel, Experimentally fitted biodynamic models for pedestrian-structure interaction in walking
situations, Mechanical Systems and Signal Processing 72-73 (2016) 590-606.
[12] C.C. Caprani, E. Ahmadi, Formulation of human–structure interaction system models for vertical vibration, Journal of Sound and Vibration
377 (2016) 346-367.
[13] E. Shahabpoor, A. Pavic, V. Racic, Identification of mass-spring-damper model of walking humans, Structures 5 (2016) 233-246.
[14] J.F. Jiménez-Alonso, A. Sáez, E. Caetano, F. Magalhães, Vertical crowd-structure interaction model to analyze the change of the modal
properties of a footbridge, Journal of Bridge Engineering 21(8) (2016) paper ID C4015004.
[15] M. Zhang, C.T. Georgakis, J. Chen, Biomechanically excited SMD model of a walking pedestrian, Journal of Bridge Engineering 21(8)
(2016) paper ID C4016003.
[16] J. Biliszczuk, W. Barcik, P. Hawryszków, C. Machelski, J. Tadla, Dynamic sensibility of cable stayed footbridges, Proceedings of the
Second International Conference Footbridge 2005, 6-8 December 2005, Venice, Italy.
[17] P. Hawryszków, Analysis of dynamical properties of footbridges, assessment of dynamical sensitivity and human comfort, PhD thesis,
Wrocław Universtity of Technology, 2009.
[18] C. Caprani, J.Qu, S. Zivanovic, N. Evans, E. Ahmadi, Quantification of human-structure interaction, in: MATEC Web of Conferences
24 EVACES’15, 6th International Conference on Experimental Vibration Analysis for Civil Engineering Structures, 2015, paper ID 07001.

Das könnte Ihnen auch gefallen