Sie sind auf Seite 1von 14

Numerical and Experimental Evaluation of the Dynamic

Performance of a Footbridge with Tuned Mass Dampers


Katrien Van Nimmen1; Pieter Verbeke2; Geert Lombaert3; Guido De Roeck4; and Peter Van den Broeck5

Abstract: This article presents an evaluation of the dynamic behavior of a slender steel footbridge before and after the installation of two
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

tuned mass dampers (TMDs). The results of the experimental study show that the damping devices lead to an increase of the effective damping
ratio of the critical mode. Additional tests involving the structural vibrations induced by a limited number of persons revealed that the TMD
units are effective in reducing the structural response. However, the obtained reduction highly depends on the type of human excitation. To
verify the response of the footbridge for large groups and crowds, a comprehensive numerical analysis is performed. The results are compared
to the response predicted by the procedures of the Setra and HiVoSS design guides. For the bridge without TMD units, a significantly higher
structural response is predicted by the design guides; the bridge has a short span and is lightly damped, so the steady-state resonant conditions
assumed in the design guides are not reached for pedestrians who cross the bridge at a normal walking speed. For the bridge with TMD units,
the design guides predict a significantly lower structural response than do the more detailed numerical simulations because the critical mode is
highly damped and no longer dominates the structural response. In addition, the effectiveness of the damping devices is considerably less than
that expected on the basis of the corresponding increase in effective damping ratio because crowd-induced loading is near-harmonic and tran-
sient, not harmonic. DOI: 10.1061/(ASCE)BE.1943-5592.0000815. © 2016 American Society of Civil Engineers.
Author keywords: Human-induced vibrations; Footbridge; Tuned mass damper (TMD); TMD effectiveness; Random pedestrian traffic;
Full-scale testing.

Introduction variabilities of pedestrian excitation. In design procedures, how-


ever, crowd-induced loading is simplified to an equivalent deter-
Because of their great slenderness, contemporary footbridges are often ministic load [Association Française de Genie Civil 2006 (referred
highly susceptible to human-induced vibrations. Hence, predicting the to here as the Setra design guide); Heinemeyer et al. 2009 (referred
dynamic response of these structures under crowd-induced loading to here as the HiVoSS design guide)].
has become a critical aspect of the vibration serviceability evaluation. When vibration serviceability problems are encountered, the
Footbridges are subjected to the simultaneous actions of groups dynamic performance of the structure can be improved by vibration
of pedestrians or crowds. To represent pedestrian-induced loading countermeasures implemented either before construction or in ret-
accurately, the corresponding load model has to account for inter- rofit (Brownjohn et al. 2004; Butz et al. 2011). Traditional remedia-
subject and intra-subject variability (Caprani et al. 2012). In addi- tion (i.e., changes in mass or stiffness) may require important and
tion, the pedestrians may also synchronize their motion with other undesired modifications of the structural design. Conversely, tuned
pedestrians or the bridge itself (Bocian et al. 2012; Carroll et al. mass dampers (TMDs) (Den Hartog 1985) can be adapted to aes-
2012). A statistical analysis, therefore, is required to account for the thetic requirements (e.g., installed in hidden compartments), leav-
ing the original structural design untouched (Ainsworth et al. 2011;
Meinhardt 2012). This vibration-mitigation technique is making
1
KU Leuven, Dept. of Civil Engineering, Structural Mechanics Section, strong gains in popularity for use in footbridges (see Fig. 1).
B-3001 Leuven, Belgium; KU Leuven, Dept. of Civil Engineering, A TMD is effective in reducing the structural response for exci-
Technology Cluster Construction, Structural Mechanics and Building tation at or near the targeted natural frequency of the footbridge
Materials Section, Technology Campus Ghent, B-9000 Ghent, Belgium (Weber et al. 2006). The corresponding reduction in the structural
(corresponding author). E-mail: katrien.vannimmen@kuleuven.be response is generally assumed to be proportional to the obtained
2
KU Leuven, Dept. of Civil Engineering, Technology Cluster Construction, increase in the effective damping ratio (Weber and Feltrin 2010).
Structural Mechanics and Building Materials Section, Technology Campus
Ghent, B-9000 Belgium.
However, the effectiveness of a TMD depends on accurate tuning
3
KU Leuven, Dept. of Civil Engineering, Structural Mechanics Section, of the damping device but also on the characteristics of the excita-
B-3001 Leuven, Belgium. tion (Caetano et al. 2010). The effectiveness is maximal for har-
4
KU Leuven, Dept. of Civil Engineering, Structural Mechanics Section, monic excitation near the original resonance frequency of the foot-
B-3001 Leuven, Belgium. bridge and is expected to be smaller as a result of the transient and
5
KU Leuven, Dept. of Civil Engineering, Structural Mechanics Section, random nature of pedestrian excitation (Occhiuzzi et al. 2008).
B-3001 Leuven, Belgium; KU Leuven, Dept. of Civil Engineering, In this article, an extensive numerical and experimental study of
Technology Cluster Construction, Structural Mechanics and Building the vibration serviceability of a slender steel footbridge in Charleroi,
Materials Section, Technology Campus Ghent, B-9000 Ghent, Belgium.
Belgium, is presented. To assess the effectiveness of the installed
Note. This manuscript was submitted on December 1, 2014; approved
on May 29, 2015; published online on January 22, 2016. Discussion period damping devices, the dynamic performance under pedestrian-induced
open until June 22, 2016; separate discussions must be submitted for indi- excitation is evaluated with and without TMDs.
vidual papers. This paper is part of the Journal of Bridge Engineering, © The outline of this paper is as follows. First, the Charleroi foot-
ASCE, ISSN 1084-0702. bridge and the results from the experimental study are presented.

© ASCE C4016001-1 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


Second, the vibration serviceability assessment, performed accord- positioned every 4.25 m on both sides of the main girder. The struc-
ing to the current codes of practice, the Setra (Association Française ture is supported by 2  2 neoprene bearings. In a single bearing, the
de Genie Civil 2006) and HiVoSS (Heinemeyer et al. 2009; Butz et horizontal translations are fixed. On the other side of the span, the lat-
al. 2008) design guide, is discussed. Then, the TMD design strategy eral translations are fixed at the center of the cross section.
is addressed briefly, as are the estimates of the corresponding reduc- The natural frequencies and corresponding mode shapes are pre-
tion in the structural response, which are made on the basis of the dicted by using a detailed FE model of the structure. The model con-
identified effective damping ratio. Third, the Monte Carlo simulation sisted of Timoshenko beam elements for the stringers and cross
of pedestrian traffic is discussed. The predicted acceleration levels girders and a regular mesh of 4-node shell elements (Mindlin-
are compared to those obtained from the procedures in the design Reissner theory) for the steel plates composing the steel box girder.
guides Setra (2006) and HiVoSS (2009) and the in situ observations. The boundary conditions are applied as indicated on the architec-
tural drawings, assuming that the neoprene bearings restrict only
the vertical translations. The resulting FE model has a total of
Charleroi Footbridge 27,866 degrees of freedom (DOF). The first three calculated modes
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

are presented in Fig. 3.


The Charleroi footbridge is a slender steel structure with a single
span of 38.25 m and a width of 13.35 m, resulting in an effective
Operational Modal Characteristics of the Footbridge
bridge deck area of 510 m2 (Fig. 2). The structure is composed of a
rectangular steel box girder with a width of 6.5 m and a height of To verify the numerically predicted natural frequencies and mode
0.8 m at midspan, which decreases to 0.5 m at the sides [Figs. 2(b and shapes, a measurement campaign was carried out before installation
c)]. Tapered I-profiles, interconnected by five stringers, are of the TMD units. In six setups, triaxial measurements were made at
a total of 50 locations on the bridge deck (regular 5  10 grid).
30 Output-only system identification is performed on the basis of ambi-
ent vibrations, mainly caused by wind (Peeters and De Roeck 1999;
Reynders and De Roeck 2008; Reynders et al. 2008). Table 1 lists the
# Footbridges

20 natural frequencies and modal damping ratios of the three modes


identified with a natural frequency below 6 Hz. These results show
10 that the identified modal damping ratios are very low and close to the
values recommended for steel structures in the design guides
(Association Française de Genie Civil 2006; Heinemeyer et al. 2009)
0
1982 1990 1998 2006 2014 (i.e., 0.4%). The natural frequency and modal damping ratio of the
Year fundamental mode identified from ambient excitation agree with the
results from a free decay analysis that involved larger vibration
Fig. 1. The number of footbridges with TMDs, installed by GERB amplitudes, which was performed by the manufacturer of the damp-
Engineering (2014) and Maurer Söhne (2013) ing devices (Colombel 2014). Table 1 presents also the comparison
with the numerically predicted natural frequencies and mode shapes.

Fig. 2. (a) The Charleroi footbridge (image by authors); location of the TMDs at (b) midspan (centerline) and (c) cross section

Fig. 3. Top and side views of the first three calculated modes of the Charleroi footbridge

© ASCE C4016001-2 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


The translational stiffnesses of the neoprene bearings are calibrated jumping on the bridge deck. The pacing rate of the activity was
to further improve the agreement between the three identified and synchronized by using a metronome signal and chosen to match the
corresponding calculated modes (see Table 1). The low relative error fundamental natural frequency of the footbridge (1.66 Hz). In addition
on the natural frequencies and the high modal assurance criterion to these synchronized activities, an additional experiment considered
(MAC) values indicate good agreement between the predicted and the persons walking at a self-selected speed. For the stationary activ-
calculated modal characteristics. ities (jumping and bobbing), the persons were gathered at midspan.
In case of walking and running, the persons were requested to travel
Tuned Mass Dampers to the other side of the span and back. The corresponding acceleration
levels of the bridge deck were registered at midspan next to the para-
Two identical TMD units (mTMD ¼ 2; 480 kg, fTMD  1:6 Hz, and pet using a sampling rate of 200 Hz. The digital signals are resampled
j TMD  12%) were installed at midspan (see Fig. 2) (i.e., at the at 50 Hz by means of a digital low-pass filter to avoid aliasing.
antinode of the fundamental mode). On the basis of the modal mass Fig. 4 presents the measured maximum vertical acceleration lev-
as calculated by the FE model, this TMD configuration corresponds els for the different load cases. To exclude potential nonrelevant in-
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

to a mass ratio of 3.8%. Note that these dampers will not affect the stantaneous peak levels, the maximum RMS values of the vertical
second or third mode of the bridge, because they are installed at acceleration levels, calculated with a time window of 1=~f 1 ðin sÞ,
nodes of the corresponding mode shapes. Here, the effectiveness of are presented as well. Fig. 4 shows that the acceleration levels
the damping devices is evaluated on the basis of the effective damp- increase with the number of persons on the bridge deck, as
ing ratio as identified under ambient excitation. expected. The increase is less than proportional to the number of
The ambient vibrations are registered with triaxial sensors in a persons, even when a metronome is used for synchronization, which
single setup at a total of 12 locations on the bridge deck. Output- indicates that full synchronization was not attained or that the struc-
only system identification (Peeters and De Roeck 1999; Reynders tural damping slightly increased as a result of high vibration ampli-
and De Roeck 2008) again identifies three modes with a natural fre- tudes or human–structure interaction effects (Wang et al. 2011). As
quency below 6 Hz. As expected, the identified modal characteris- expected, the highest acceleration levels are reached for the load
tics of the second and third mode are quasi-identical to the corre- case jumping. The large displacements corresponding to the low-
sponding results identified before the installation of the TMD units frequency oscillations at the fundamental natural frequency of the
(see Table 1). The identified fundamental mode is again the funda- footbridge (1.66 Hz) were perceived as particularly highly uncom-
mental vertical bending mode, but the corresponding natural fre- fortable by the participants. The high acceleration levels presented
quency decreased to 1.6 Hz, and the modal damping ratio increased in Fig. 4 highlight the strong sensitivity of the Charleroi footbridge
to 6%. These results are again in excellent agreement with those esti- to human-induced vibrations.
mated from the free decay analyses performed by the manufacturer After installation of the TMD units, a number of experiments
(Colombel 2014). These analyses show that the effective damping were repeated to assess the effectiveness of the damping devices.
ratio of the fundamental mode has increased by a factor of 15. Fig. 5 presents the peak and RMS values of the vertical accelera-
tions registered during free walking and during synchronized walk-
ing and jumping for 1, 8, or 18 persons on the bridge deck. Fig. 5
Human-Induced Vibrations: Full-Scale Experiments
shows that the acceleration levels are significantly lower after in-
Prior to the installation of the TMD units, a number of tests involving stallation of the TMD units.
different types of human excitation were performed. These experi- The maximum RMS values of the acceleration levels induced by
ments involved 1, 8, or 18 persons walking, running, bobbing, or 8 and 18 persons are reduced by factors of 1.5, 5.1, and 3.5 for the

~ Natural Frequency ~f , and Damping Ratio j~ ) of the First Three Modes and Corresponding
Table 1 Identified Modal Characteristics (Mode Number N,
Mode Number N, Natural Frequency f, the Modal Assurance Criterion (MAC), and the Relative Error in Natural Frequency « for the Initial and Calibrated
Numerical Models

Measured Initial FE Model Calibrated FE Model


~
N ~f (Hz) j~ (%) N f (Hz) ɛ (%) MAC (–) N f (Hz) ɛ (%) MAC (–)
1 1.66 0.41 1 1.65 –0.6 0.99 1 1.67 –0.7 0.99
2 5.16 0.29 2 5.22 –1.2 0.90 3 5.12 0.8 0.90
3 5.56 0.48 3 6.15 10.6 0.91 4 5.59 –0.6 0.95

Fig. 4. The maximum vertical acceleration and RMS values resulting from (a) 1 person, (b) 8 persons, and (c) 18 persons for free walking (I),
synchronized walking (II), free running (III), synchronized jumping (IV), and bobbing (V) at the fundamental natural frequency of the footbridge (~f 1 )

© ASCE C4016001-3 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


load cases of free walking, synchronized walking, and jumping, Design Guides
respectively. The reduction observed for a single person is smaller
The Setra (Association Française de Genie Civil 2006) and HiVoSS
and considered not relevant because of the nonnegligible contribu-
(Heinemeyer et al. 2009) design guides assess the vibration service-
tion of ambient excitation. The obtained reduction is maximal for ability of a footbridge on the basis of the maximum acceleration lev-
synchronized walking, because the contribution of the fundamental els for different pedestrian densities. The response of the structure
mode is predominant in this case. This reduction can also be is predicted assuming that the first or the second harmonic of the
observed clearly in the time history and amplitude spectrum of the walking load coincides with one of the natural frequencies of the
acceleration levels shown in Fig. 6, induced by 18 persons walking structure. It is furthermore assumed that the structure is lowly
in synchronization at ~f 1 . The somewhat lower reduction obtained damped and that the resonant mode dominates the structural
for jumping is explained by the larger contribution of the higher response. A distinction is made between sparse and dense crowd
modes, excited by the higher harmonics of the jumping load (Ellis conditions to account for the higher level of synchronization that is
and Ji 2004). expected for high pedestrian densities.
These practical tests reveal the sensitivity of the footbridge to
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

The key element of the design guides is that the pedestrian exci-
human-induced vibrations and enable the identification of critical tation that results from N pedestrians (in persons/s) is represented
modes. The results show that TMD units are effective in reducing by an equivalent deterministic load that is distributed uniformly on
the structural response. The obtained reduction depends on the type the bridge deck. This simplified load model consists of an equiva-
of human excitation. lent number ðNeq Þ of perfectly synchronized pedestrians, which is
defined such that the acceleration level obtained is the same as the
95th percentile value of the peak accelerations caused by N pedes-
Vibration Serviceability Assessment trians predicted by the considered stochastic load model. The equiv-
alent number of pedestrians, Neq , is defined as follows:
The French design guide Setra (Association Française de Genie qffiffiffiffiffiffiffiffi
Civil 2006) and the European design guide HiVoSS (Heinemeyer Neq ¼ 10:8 j j N for d < 1 person=m2 (1)
et al. 2009; Butz et al. 2008) enable designers to check the
vibration serviceability of a footbridge for pedestrian den- pffiffiffiffi
sities up to 1.5 persons/m2. This section provides a brief sum- Neq ¼ 1:85 N for d  1 person=m2 (2)
mary of the corresponding design procedures. For a more
comprehensive discussion, the reader is referred to the work where j j denotes the modal damping ratio; and d ðin persons=m2 Þ
of Van Nimmen et al. (2014a). The next section discusses represents the considered pedestrian density. The pedestrian den-
how a TMD can be accounted for in the methodology of the sities that must be considered in the analysis depends on the
design guides and presents the results of the vibration service- expected pedestrian traffic. The traffic classes considered in the
ability assessment of the Charleroi footbridge, with and with- design guides vary from 15 persons distributed over the entire
out TMDs. bridge deck up to a density of 1.5 persons/m2. The corresponding

Fig. 5. The maximum vertical acceleration and RMS values before (dark) and after (light) installation of the TMDs for (a) 1, (b) 8, and (c) 18 persons
for free walking (I), synchronized walking (II), and synchronized jumping (IV) at the fundamental natural frequency of the footbridge (~f 1 ) and the
maximum (dashed) and maximum (solid) RMS values of the ambient accelerations

Fig. 6. (a) Time history and (b) amplitude spectrum of the vertical acceleration levels registered at midspan next to the parapet for 18 persons walking
in a synchronized manner at the fundamental natural frequency of the footbridge (~f 1 ) before (light) and after (dark) installation of the TMDs

© ASCE C4016001-4 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


Fig. 7. The load amplitude Qeh ðfj Þ for the vertical component of the Fig. 8. The maximum effective damping ratio j j;eff ðÞ in terms of the
walking load according to the Setra (Association Française de Genie mass ratio m for different values of the inherent structural modal damping
ratio j j ðÞ: 0.00 (solid line), 0.01 (dashed line), and 0.02 (dotted line)
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Civil 2006) (thick dark line) and HiVoSS (Heinemeyer et al. 2009;
Butz et al. 2008) (thick light line) design guides and the value recom-
mended by Tricon (Van Nimmen et al. 2014a; Van den Broeck and De
Roeck 2012) (thin black line) for application in the design stage (dashed Note that the predicted maximum acceleration levels depend
line) and after experimental verification of the natural frequencies only on the natural frequency of the considered mode j through the
(solid) load amplitude Qeh ðfj Þ [see Eq. (4) and Fig. 7]. Within the frame of
the Tricon project (Van den Broeck and De Roeck 2012), Van
Nimmen et al. (2014a) showed that in the design stage, uncertainty
amplitude of the equivalent load qeq;e (in N=m2 ) in direction e (ver- with respect to the predicted dynamic properties of the footbridge is
tical, lateral, or longitudinal) is defined as follows: inevitable. Even with a detailed finite element model, deviations up
to 10% in terms of natural frequencies are expected (Van Nimmen
Neq
qeq;e ¼ Qeh ðfj Þ (3) et al. 2014a). For these reasons, the authors recommended the modi-
Aeff fied load factor shown in Fig. 7 for the vertical component of the
walking load.
with The dynamic behavior of the footbridge near resonance strongly
Qeh ðfj Þ ¼ aeh G c eh ðfj Þ and 0  c eh ðfj Þ  1 (4) depends on the damping ratio, a parameter which is difficult to esti-
mate in design stage. On the basis of experience obtained from simi-
lar structures, minimum and mean values for the damping ratio are
where fj is the natural frequency of mode j under considera-
suggested by the design guides for different construction types. For
tion; Aeff ðm2 Þ is the effective bridge deck area; and Qeh ðfj Þ (in the case of steel structures, a mean damping ratio of 0.4% is
N) is the load amplitude of the hth harmonic of the walking suggested.
load in direction e generated by a single pedestrian, defined as To assess the vibration serviceability, the predicted acceleration
the product of the dynamic load factor aeh ðÞ; the weight levels ½€u max ðin m=s2 Þ are compared to the limit values of the com-
G ¼ 700 ðin NÞ, and the reduction coefficient c eh ðfj ÞðÞ: The fort classes, corresponding to the comfort requirements specified by
latter accounts for the probability that the step frequency (or its the bridge owner ½€ u lim ðin m=s2 Þ:
second harmonic) coincides with the natural frequency of mode
j under consideration and, thus, the occurrence of resonance.
The load factor Qeh for the vertical component of the walking TMD Design Strategy
load, as defined by the design guides, is presented in Fig. 7.
Because resonant conditions are considered, the design guides When the adopted comfort requirements are not satisfied ð€ u j;max >
assume that the response of the structure is dominated by the contri- €
u lim Þ; modifications to the design or vibration countermeasures are
bution of the considered mode. The maximum acceleration levels needed (Weber et al. 2006; Varela and Battista 2011). The high

u jemax (in m/s2) in direction e, observed in the antinode of that vibration amplitudes that occur in near-resonant conditions can be
mode, are therefore calculated as follows: reduced by increasing the effective damping of the structure
through the implementation of TMDs (Weber and Feltrin 2010;
Fje Viguie 2010; Setareh et al. 2006).

u je max ¼ maxj f je j (5)
2j j First, a choice of the mass of the TMD ðmTMD Þ is made, usu-
ally in the range of 1%–4% of the modal mass mj associated with
with the mode of vibration under consideration. The optimal stiffness
X ðkTMD Þ of the spring and damping constant ðcTMD Þ of the linear
Fje ¼ qeq;e aeff;k j f je;k j (6) viscous damper depend on the desired reduction (Weber et al.
k 2006). Because the vibration criteria of the design guides are
expressed in terms of accelerations, the TMD for this study is
where f je is the vector that collects the mass-normalized modal dis- designed such that the structural accelerations are minimized.
placements of mode in direction e for all neff nodes of the effective The corresponding optimal TMD characteristics are derived by
bridge deck area Aeff ; Fje is the modal load in direction e; and aeff is Asami et al. (2002) for the case in which the structure, vibrating
the vector that collects
 the bridge deck area  allocated to the corre- in the mode considered, is simplified to an undamped single-DOF
X 
sponding nodes Aeff ¼ aeff;k ðin m Þ . By taking the absolute
2 (SDOF) system. The optimal frequency ðfTMD Þ and damping ratio

k ð j TMD Þ of the TMD are given as a function of the ratio m of the
value of the modal displacements in Eq. (5), the equivalent load is mass of the TMD and the relevant modal mass ð m ¼ mTMD =mj Þas
chosen such that the modal load is maximized. follows:

© ASCE C4016001-5 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


than for dense pedestrian traffic, which can be understood consider-
ing that the design guides are developed for footbridges with low in-
herent damping and not for highly damped structures, as are consid-
ered here.
Following Eq. (10), the minimum effective damping ratio reads
as follows:

For d < 1 ðperson=m2 Þ For d  1 ðperson=m2 Þ


   
u j;max 2
€ €
u j;max (11)
j j;eff ¼ j j;max j j;eff ¼ jj

u lim €
u lim

where j j is the original structural damping ratio of the footbridge.


Fig. 9. The equivalent number of pedestrians (dark lines) and maximum
The minimum mass ratio ð m min Þ of the TMD is calculated from the
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

acceleration levels (light lines) in terms of the inherent structural damping


minimum effective damping ratio ð j j;eff Þ by means of Eq. (9). The
ratio for sparse (solid lines) and dense (dashed lines) pedestrian densities
actual mass ratio is often chosen to be (slightly) higher than m min
for reasons of robustness and safety (Papadimitriou et al. 1997;
sffiffiffiffiffiffiffiffiffiffiffiffiffi Marano et al. 2010) but also to limit the relative displacement am-
 1
fTMD ¼ fj (7) plitude of the TMD with respect to the primary structure and, thus,
1þm the required space.

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffirffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Results
3m 27
j TMD ¼ 1þ m (8) Only the fundamental mode of the Charleroi footbridge has a natu-
8ð1 þ m Þ 32
ral frequency that falls in the range between 0.5 and 5.0 Hz, identi-
fied as critical for human-induced vibrations by the design guides.
These values are close to optimal for lightly damped structures. Given the low value of the fundamental natural frequency, reso-
The maximum steady-state response of the structure with a TMD is nance at the first harmonic of the vertical walking load may occur.
expressed as a function of the effective damping ratio ð j j;eff Þ of a A modal damping ratio of 0.4% is assumed, as recommended by
SDOF as proposed by Weber and Feltrin (2010). This effective the design guides for steel structures. Fig. 10(a) presents the maxi-
damping ratio is not a modal damping ratio as v j is no longer a natu- mum vertical acceleration levels predicted on the basis of Eq. (5).
ral frequency of the system when the TMD is added to the foot- Fig. 10(a) shows that a minimum level of comfort is expected for
bridge. For structures with low inherent damping ½ j j < 0:02 ðÞ; low pedestrian densities, whereas unacceptably high vibration lev-
the effective damping ratio j j;eff is approximated by (Weber and els are found for high pedestrian densities. Because of the limited
Feltrin 2010) the following: lateral and longitudinal displacements associated with this ver-
0 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi11 tical bending mode, the corresponding accelerations are well
2 below the maximum comfort requirements for horizontal vibra-
j j;eff ¼ j j þ @2 A (9) tions (n0:1 m=s2 ).
m ð1 þ m Þ
The footbridge is located in a city center where urban use and
high pedestrian densities are expected. It was decided, therefore, to
Hence, the mass of the TMD determines what effective damping increase the effective damping ratio of the fundamental mode by
ratio can be achieved (see Fig. 8) and, therefore, the amplitude of adding a number of TMDs to the footbridge. To enable optimal tun-
the maximum steady-state response. ing of the damping devices, the spring and damping characteristics
The required minimum effective damping ratio ð j j;eff Þ is deter- are determined after the footbridge is built (i.e., based on the experi-
mined by the limit value of the acceleration levels € u lim or each of mentally identified modal parameters). The chosen mass ratio of
the pedestrian densities. From Eqs. (1)–(6), it can be shown that the 3.8% is expected to result into an effective damping ratio of 7%
predicted acceleration levels ð€
u j;max Þ depend on the following: under steady-state conditions, assuming the devices to be optimally
tuned (see Fig. 8).
For d < 1 person=m
pffiffiffiffi
2
For d  1 person=m2 Fig. 10(b) compares the calculated frequency response function
(10)
€u j;max ð j j Þ  1 u j;max ð j j Þ  1
€ (FRF) relating the vertical accelerations and an input force at mid-
span with and without the TMD units. The maximum amplification
Because of the assumed resonant conditions, the dynamic ampli- factor is reduced by a factor of 15 by including the TMD units,
fication factor is inversely proportional to the damping ratio. which is perfectly in line with the experimentally identified effec-
However, in the case of low pedestrian densities, the amplitude of tive damping ratio. Because the natural frequency of the damping
the maximum accelerations is inversely proportional to the square devices is slightly lower than optimal, the effective damping ratio is
root of the modal damping ratio [Eq. (10)]. In this case, the damping somewhat lower than the maximum of 7%, which theoretically can
ratio also affects the amplitude of the equivalent load through the be attained with the present mass ratio. Note that in the experimental
equivalent number of pedestrians, Neq [Eq. (1)]. The relation study, the reductions observed in structural response for free walk-
between the structural damping and the equivalent number of ing, synchronized walking, and jumping, factors of 1.5, 5.1, and 3.5
pedestrians as defined by Eqs. (1) and (2), respectively, is shown in respectively, are significantly less.
Fig. 9. For values of the structural damping higher than 2.93%, the The vibration serviceability of the footbridge is reassessed while
equivalent number of pedestrians, Neq , is higher for sparse than for taking into account the effectiveness of the TMDs. The latter is
dense pedestrian densities. The acceleration levels predicted by accounted for by the identified effective damping ratio of 6%. The
using this simplified load model in this case are higher for sparse results of the vibration serviceability assessment are presented in

© ASCE C4016001-6 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


Fig. 10. The vibration serviceability assessment according to (a) the design guides (Association Française de Genie Civil 2006; Heinemeyer et al. 2009)
and (b) the calculated FRF for a vertical input force and output accelerations at midspan before (dark line) and after (light line) installation of the TMDs
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

considered stochastic load model. To verify the validity of this


approach for the highly damped structures obtained by the introduc-
tion of a TMD, the characteristics of the pedestrian flows generated
in this study are set to match the simulation model that is applied by
the Setra and HiVoSS design guides in deriving this simplified load
model. Intra-subject variability and synchronization phenomena are
disregarded. As far as human–structure interaction is concerned,
only the added mass of the considered pedestrian density is
accounted for.
In the following is a discussion of the crowd-induced vibrations
Fig. 11. Time history of the single-step load according to Li et al. of the Charleroi footbridge, which are analyzed on the basis of the
(2010) for three different step frequencies: 1.6 Hz (dashed line), 2.0 Hz Monte Carlo simulation of the pedestrian traffic. First, the simula-
(solid line), and 2.4 Hz (dotted line) tion model is discussed. Second, the structural vibrations induced
by a single pedestrian crossing the footbridge with and without
TMD units are examined. Third, the dynamic performance of the
Fig. 10(a), which illustrates that the installed dampers ensure maxi- Charleroi footbridge before the installation of the TMDs is dis-
mum vibration comfort for all pedestrian densities. cussed. Finally, there is a discussion on the effectiveness of the
TMD units under pedestrian traffic, which is evaluated and com-
Stochastic Load Model for Pedestrian Excitation pared to the in situ observations.

In practice, footbridges are subjected to the simultaneous actions of


groups of pedestrians or crowds (Association Française de Genie
Civil 2006). Crowd-induced loading has to account for inter-subject Simulation Model
and intra-subject variability (Caprani et al. 2012; Van Nimmen et The simulation model consist of three essential elements: (1) the
al. 2014b). In addition, a crowd is in fact a group of mechanical sys- force model describing the walking loads of a single pedestrian; (2)
tems that interact with the supporting structure and thereby modify the crowd model generating the pedestrian traffic; and (3) the struc-
the dynamic behavior of this structure (Brownjohn et al. 2004; ture model describing the dynamic behavior of the footbridge.
Bocian et al. 2013). Because this type of mechanical interaction has These elements are discussed in the following paragraphs.
not been fully characterized (Živanovic et al. 2009; Bocian et al.
2013), numerical models usually account only for the additional
mass on the structure. Furthermore, pedestrians will enter a bridge
Force Model for a Single Pedestrian
at different times, and they may also synchronize their motion with
In this study, only the vertical component of the walking load is
that of other pedestrians or the motion of the bridge itself (Bocian et
considered, because this is the only component relevant for the con-
al. 2012; Carroll et al. 2012). Hence, a statistical analysis is required
sidered footbridge. The generalized single-step load model by Li et
to account for the inherent variability of pedestrian excitation in the
al. (2010) is assumed for the vertical induced forces:
evaluation of the predicted structural response. To this end, recently
developed stochastic load models by Ingólfson and Georgakis   
p1 ðtÞ X 5
pn 0  t  Tc
(2011), Caprani (2014), Piccardo and Tubino (2012), Tubino and ¼ An ðfs Þsin t for (12)
Piccardo (2015), and Živanovic et al. (2010) and available walking G n¼1
T c 1:6 < fs < 2:4 Hz
parameter distributions (Sahnaci and Kasperski 2011; Kasperski
and Czwikla 2012) can be applied. The design procedures of the with p1 ðtÞ (in N) the vertical single-step load as a function of time
Setra and HiVoSS, however, propose a single deterministic analysis t; G ðin NÞ the weight of the pedestrian; n the order number of the
for predicting the pedestrian-induced vibrations based on an equiva- harmonic; fs ðin HzÞ the step frequency; An ðÞ the Fourier coeffi-
lent simplified load model. This load model consists of an equiva- cient of the nth harmonic normalized to the weight of the pedestrian;
lent number of perfectly synchronized pedestrians, Neq , uniformly and Tc ðin sÞ the duration of the contact between the foot and the
distributed over the bridge deck, which is defined such that the ground. The typical time history of the vertical single-step load is
acceleration level obtained is the same as the 95th percentile value shown in Fig. 11 for three different step frequencies. For every indi-
of the peak accelerations due to N pedestrians as predicted by the vidual in the crowd, a weight of 700 N is assumed.

© ASCE C4016001-7 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


Crowd Model Setra (Association Française de Genie Civil 2006) and HiVoSS
To define the pedestrian traffic, assumptions are required with (Heinemeyer et al. 2009) design guides.
respect to the step frequency, the walking trajectory, the walking Sparse crowd conditions are used to simulate four pedestrian
speed, and the arrival time of every pedestrian. The considered densities that vary from 15 persons on the bridge deck up to
walking trajectories consist of straight lines parallel to the longitudi- 0.2, 0.5, and 0.8 person/m2. For the case of dense crowds, den-
nal axis x of the bridge deck. The offsets in the lateral direction are sities of 1.0 and 1.5 persons/m2 are considered. As a result,
chosen randomly along the bridge deck width. All pedestrians are there are a total of six load cases to be considered for each natu-
assumed to move at the same walking speed of 1.5 m/s. For a given ral frequency of the footbridge below 5 Hz. In the case of the
pacing rate, fs , the step length, ls ðin mÞ, follows from Charleroi footbridge, only the fundamental mode has a natural
vs frequency below 5 Hz.
ls ¼ (13)
fs
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

The time taken by a pedestrian to cross the bridge can be


expressed as follows: Structure Model
The uncoupled governing equations of motion for a linear finite ele-
L ment model with proportional damping in modal coordinates read
T¼ (14)
vs
€ _ þ X2 zðtÞ ¼ U> PðtÞ
z ðtÞ þ CzðtÞ (16)
where L ðin mÞ is the length of the bridge. The arrival rate
λ ð in persons=sÞ is computed from the pedestrian density d where zðtÞ 2 RnDOF is the modal coordinate vector; X2 2 RnDOF nDOF
is a diagonal matrix containing the square of the natural frequen-
N Aeff d
λ¼ ¼ (15) cies v j (in rad/s); C 2 RnDOF nDOF is a diagonal matrix containing
T T the terms 2 j j v j with j j ðÞ as the modal damping ratios; and
U 2 RnDOF nDOF is a matrix that has the mass-normalized mode
The arrival times are assumed to follow a Poisson distribution
shapes f j as columns and U> PðtÞ as the modal projection of the
(Živanovic 2012). For the special case in which all pedestrians have
external forces PðtÞ 2 RnDOF . By retaining only the relevant modes in
the same step frequency and walking load, p1 ; these Poisson arrival
a modal decomposition (nm nnDOF ), the model order is drastically
times result in a random phase distribution among the walking loads
reduced by zr ðtÞ 2 Rnm as uðtÞ ¼ Ur zr ðtÞ and Ur 2 RnDOF nm . The
of the individuals in the crowd.
mass of the pedestrians on the footbridge is assumed to be uniformly
For low pedestrian densities ðd < 1 person=m2 Þ, spatially unre-
distributed on the bridge deck and taken into account by lumping the
stricted traffic conditions, and thus free movement of the pedes-
trians, is assumed. In the case of dense crowds ð1  d  additional mass with the corresponding translational DOF.
1:5 persons=m2 Þ, normal walking behavior gets obstructed, which When, in addition, a number of TMDs are coupled to the foot-
causes the forward movement of the stream to slow down and the bridge, the model of the structure is extended accordingly. Each
level of synchronization to increase. Beyond the upper limit value TMD can be represented by a SDOF system characterized by a
of 1.5 persons/m2, walking of pedestrians is considered to be impos- mass (mTMD ), damping (cTMD ), and stiffness (kTMD ). The differen-
sible, which significantly reduces the dynamic effects (Heinemeyer tial equations of motion for the main structure (footbridge) and the
et al. 2009). The sparse crowd conditions are simulated assuming TMDs are solved while ensuring compatibility and equilibrium
a truncated Gaussian distribution of the step frequencies, cen- conditions at the contact points. The equations of motion for the
tered around one of the natural frequencies of the footbridge fj combined system consisting of the footbridge and the TMDs are
with a standard deviation of 0.175 Hz [fs;k N ð m fs;k ; s fs;k Þ ¼ obtained by eliminating the contact forces
N ðfj ; 0:175Þ ðin HzÞ]. For dense crowd conditions, all pedes-  >
trians are given the same step frequency. In the simplified load  z€ ðtÞ þ C
K  2 z ðtÞ ¼ Ur Sp pðtÞ; with
 z_ ðtÞ þ X (17)
0
of the design guides, these assumed distributions of step fre-
quencies result in an equivalent number of pedestrians, which is
larger for dense crowd conditions [see Eqs. (1) and (2)]. The " # " #
zr ðtÞ 2 ¼ X2 U⊤
r STMD kTMD
stated properties of sparse and dense crowd conditions are cho- z ¼ ; X ⊤
(18)
sen in agreement with the stochastic load model applied by the uTMD ðtÞ kTMD STMD Ur kTMD

Fig. 12. Time history of the simulated vertical acceleration levels at midspan for a single pedestrian crossing the bridge at a walking speed of 1.0 m/s
(dark lines) or 1.5 m/s (light lines), leaving the footbridge at the time indicated by the vertical line (dashed lines), and the amplitude of the correspond-
ing steady-state response (horizontal line): (a) fs ¼ f1 ; (b) fs ¼ 1:2f1

© ASCE C4016001-8 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


 
 ¼ I 0  ¼ C U⊤
r STMD cTMD
The dynamic behavior of the Charleroi footbridge is simulated
K ; C
0 mTMD cTMD S⊤
TMD Ur cTMD on the basis of the natural frequencies and mode shapes of the
(19) calibrated FE model and the experimentally identified modal
damping ratios. All modes with a frequency up to 15 Hz are taken
into account. It has to be acknowledged that in many cases,
in which uTMD 2 RnTMD collects the absolute displacements of the human–structure interaction phenomena cause damping of the
nTMD installed TMD units; STMD 2 RnDOF nTMD is a matrix that indi- coupled crowd–structure system to increase in comparison to the
cates the DOF of the footbridge at which the TMD is coupled; and inherent structural damping ratio (Wang et al. 2011; Busca et al.
mTMD 2 RnTMD nTMD ; kTMD 2 RnTMD nTMD , and cTMD 2 RnTMD nTMD 2014; Bocian et al. 2013; Pedersen 2015). It is expected that these
are diagonal matrices containing the mass (mTMD ), damping interaction effects would be nonnegligible for high pedestrian
(cTMD ), and stiffness (kTMD ) parameters of the different TMD units. densities (Jones et al. 2011; Dougil et al. 2006; Van Nimmen et
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Time history of the simulated vertical acceleration levels at midspan for a single pedestrian crossing the bridge with TMDs at a walking
speed of 1.0 m/s (dark lines) or 1.5 m/s (light lines), leaving the footbridge at the time indicated by the vertical line (dashed lines) and the amplitude of
the corresponding steady-state response (horizontal line): (a) fs ¼ f1 ; (b) fs ¼ 1:2f1

Fig. 14. For a pedestrian density of 1 person/m2: (a) the time history of the acceleration levels at midspan; (b–d) the location of the first (circle) and
the following (dots) pedestrians on the bridge deck, crossing the bridge for the first (dark dots) and second (light dots) time at three different points in
time: (b) 0.3T; (c) 0.9T; (d) 1.5T

Fig. 15. The empirical CDF of the maximum acceleration (light lines) and maximum RMS (dark lines) values at midspan of the centerline (circle)
and sidelines (square and ) of 500 random pedestrian streams for footbridge traffic without TMDs: (a) sparse traffic conditions (0.8 person/m2); (b)
dense traffic conditions (1.5 persons/m2)

© ASCE C4016001-9 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


al. 2015); however, these effects are beyond the scope of this in a significant increase in the predicted acceleration levels.
paper. Therefore, the walking speed is set to 1.0 m/s.
Figs. 13(a and b) present a similar comparison for the foot-
Crossing of a Single Pedestrian bridge with TMD units. For resonant conditions, the step fre-
quency is chosen to match the frequency at which the peak value
Figs. 12(a and b) present the acceleration levels at midspan for a sin- of the FRF is found (see Fig. 10). Although the acceleration levels
gle pedestrian crossing the Charleroi footbridge considering reso- are again higher under resonant conditions, the difference in maxi-
nant ðfs ¼ f1 Þ and nonresonant conditions ðfs ¼ 1:2 f1 Þ, respec- mum amplitude in this case is limited to a factor of 2. These fig-
tively. It is clear that the acceleration levels are significantly higher ures also illustrate that under both resonant and nonresonant
(>10 times) when resonant conditions are considered. In addition, loading conditions, acceleration amplitudes that are very close
Figs. 12(a and b) show the steady-state response for a pedestrian to the corresponding steady-state amplitude are attained, which
walking on spot at midspan. The amplitudes of the steady-state is because the number of cycles required to reach the steady-
responses in both cases are higher than the maximum acceleration state is significantly lower.
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

level obtained by a pedestrian crossing the bridge. Under resonant Comparing the results in Figs. 12(a) and 13(a) shows that the
conditions, the steady-state response reaches an acceleration level TMD units significantly reduce the response under resonant loading
that is twice as high. The lower response obtained for the moving conditions. For nonresonant loading conditions [Fig. 12(b) and
pedestrian is a result of the fact that the time until steady-state is 13(b)], the effect of the TMD units on the maximum attained accel-
much larger than the time it takes the pedestrian to cross the span of eration level is insignificant.
the bridge. The number of cycles required to reach a certain percent-
age of the steady-state amplitude for resonant loading is inversely Dynamic Performance of the Footbridge without TMDs
proportional to the damping ratio (Chopra 1995). For the lowly
damped fundamental mode of the Charleroi footbridge, 116 on- Fig. 14(a) shows the time history of the acceleration levels at mid-
spot cycles (and therefore steps) are needed to reach 95% of the span induced by a stream of pedestrians. Figs. 14(b–d) present the
steady-state amplitude under resonant conditions. Although the locations of the pedestrians on the bridge deck at three different
maximum response under nonresonant conditions is less sensitive points in time, corresponding to distances travelled by the first pe-
to the walking speed, similar conclusions can be drawn from Fig. destrian of 30%, 90%, and 150% of the span length, respectively.
12(b). These figures show that all pedestrians are walking in the same
Because the maximum acceleration levels are highly sensitive to direction (i.e., from the left to the right side of the span). Note that a
the number of loading cycles, the walking speed of the pedestrians pedestrian, who left the bridge on the right side, is immediately
is an important parameter. As discussed in the Crowd Model, the reentering the bridge from the left side on his next step. In this way,
design guides consider a fixed walking speed of 1.5 m/s. the continuity of the pedestrian flow and preservation of the pedes-
Experiments have shown, however, that the walking speed is related trian density is guaranteed.
to the step frequency. For the low step frequencies ð1:5 < fs < The simulated response caused by a stream of pedestrians is rele-
1:65 HzÞ considered here, a walking speed of 1.0 m/s is more realis- vant once the desired pedestrian density and steady-state response
tic (Bertram and Ruina 2001; Kuo 2001). Fig. 12(a) shows that are reached. As discussed in Crowd Model, the desired pedestrian
reducing the walking speed to 1.0 m/s in resonant conditions results density ðd ¼ N=Aeff Þ is reached within the time, T, needed for one
pedestrian to cross the bridge. The steady-state response, however,
is reached only at time T þ DT, where DT is dependent on the free
decay response. Analysis showed that for this application, the
steady-state response is reached at t  2T [see Fig. 14(a)]. For ev-
ery simulated stream of pedestrians, the maximum acceleration
level and corresponding maximum RMS value over the relevant
time window ½2T; 3T are retained for each node of the effective
bridge deck area.
Figs. 15(a and b) present the empirical cumulative distribution
Fig. 16. Selected output locations on the bridge deck: the centerline
functions (CDFs) of the maximum acceleration levels for three
(circle) and two sidelines (square and ) with the corresponding output
selected output locations at midspan (as indicated in Fig. 16) for
location at midspan (dark marks)
500 pedestrian streams under sparse and dense traffic conditions,

Fig. 17. Ratio of the 95th percentile of the maximum acceleration at location xð€ u x95 Þ and the overall maximum u€ max95 (light shapes) and maximum
RMS (dark shapes) values at the centerline (circle) and sidelines (square and x) along the bridge deck length (x) for the Monte Carlo–simulated pedes-
trian traffic with: (a) sparse or (b) dense conditions for the footbridge without TMDs

© ASCE C4016001-10 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


respectively. To ensure convergence of the 95th percentile value of frequency of the footbridge. The (near-)resonant loading with the
the peak acceleration levels, a 5% tolerance is set at 95% confi- second (or third) harmonic of the walking load increases the con-
dence. Thus, the number of Monte Carlo samples is increased tribution the higher modes to the overall structural response.
until the probability that the true percentile of the distribution is
within 65% of the corresponding percentile of the simulated data
accumulated thus far is higher than 95% [Statistics Toolbox Comparison with the Response Predicted According to
(MATLAB 2014)]. The converged 95th percentile value of the the Design Guides
maximum acceleration levels and corresponding maximum RMS
value are retained for each load case. Fig. 18 compares the maximum acceleration levels predicted by the
Figs. 17(a and b) present the 95th percentile values of the peak design guides to the corresponding 95th percentile values of the
acceleration levels for the selected output locations on the bridge Monte Carlo–simulated crowd-induced vibrations. These results
deck (see Fig. 16) under sparse and dense traffic conditions, respec- show that the design guides overestimate the response due to the
tively. The results presented in Fig. 17(b), almost perfectly display assumed steady-state response. For dense traffic conditions, the val-
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

a half-cycle sine shape, whereas the ratios of the corresponding ues predicted by the design guides are approximately twice as high
maximum and maximum RMS values of the acceleration levels are as the corresponding 95th percentile values of the Monte Carlo sim-
approximately equal to the square root of 2. These results indicate ulations. A better agreement in the predicted acceleration levels is
that for these dense traffic conditions, in which all pedestrians are obtained for sparse pedestrian densities as in this case the consid-
walking with a step frequency fs ¼ f1 , the structural response is ered loading is mainly nonresonant. The number of cycles required
clearly dominated by the contribution of the fundamental bending to reach the steady-state amplitude is in this case significantly
mode, as expected. smaller and reached within the time needed to cross the bridge.
Fig. 17(a) shows that for sparse pedestrian densities, the maxi-
mum acceleration levels are still attained at midspan. However,
the ratios of the corresponding maximum and maximum RMS val- Dynamic Performance of the Footbridge with TMDs
ues of the acceleration levels, calculated with a time window of
1=~f 1 (in s), are lower than the square root of 2. This observation Figs. 19(a and b) present the 95th percentile values of the maximum
indicates that the relative contribution of the higher modes is acceleration levels and maximum RMS values for the selected out-
increased, which results from the fact that for sparse crowd condi- put locations on the bridge deck (see Fig. 16), for sparse and dense
tions, the step frequencies of the individuals in the crowd follow a traffic conditions respectively. The half-cycle sine shape of the
Gaussian distribution centered around the fundamental natural results presented in Fig. 19(b), suggests that for dense traffic condi-
tions the fundamental bending mode still dominates the structural
response. However, the ratios of the maximum and maximum RMS
value of the acceleration levels are considerably higher than the
square root of 2 while the acceleration levels at the central line are
significantly lower than those at the outer sides. This is due to the
fact that the relative contribution of the torsional mode has
increased as a result of the (reduced) contribution of the fundamen-
tal bending mode.
For sparse traffic conditions, the maximum acceleration levels
and maximum RMS values no longer display the shape of the fun-
damental vertical mode but rather a combination of the first and the
second bending mode, in particular along the centerline of the
bridge [Fig. 19(a)]. In addition, the ratios of the corresponding max-
Fig. 18. Maximum vertical acceleration predicted according to the imum peak and maximum RMS value of the acceleration levels are
design guides (Association Française de Genie Civil 2006; Heinemeyer considerably higher than the square root of 2. These results reveal
et al. 2009) (dark bars), the 95th percentile value of the Monte Carlo– that for sparse pedestrian densities, the second vertical bending
simulated pedestrian traffic (light bars), and the steady-state response mode and the fundamental torsional mode are dominating the struc-
obtained for the pedestrians walking on spot () tural response. This is again due to the reduced contribution of the
fundamental bending mode.

Fig. 19. Ratio of the 95th percentile of the maximum acceleration levels at location xð€ € max95 (light) and maximum
u x95 Þ and the overall maximum u
RMS (dark) values at the centerline (circle) and sidelines (square and ) along the bridge deck length (x) for the Monte Carlo–simulated pedestrian
traffic with (a) sparse or (b) dense conditions for the footbridge including the TMD units

© ASCE C4016001-11 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


experimentally for free and synchronized walking (i.e., 1.5 and 5.1,
respectively). The assessment presented in Fig. 21 shows that by
installing the TMD units, mean and maximum comfort levels are
obtained for low and high pedestrian densities, respectively. It
should be emphasized that the effectiveness of the TMD units
highly depends on the contributions of the remaining (lowly
damped) low-frequency modes of the structure. Consequently,
effectiveness is strongly case specific.

Conclusions
Fig. 20. Maximum vertical acceleration predicted by the design
guides (dark bars) and the 95th percentile of the maximum acceleration The Charleroi footbridge is a small-span slender steel footbridge.
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

levels (light bars) of the Monte Carlo–simulated pedestrian traffic for Because of the low value of the fundamental natural frequency,
the Charleroi footbridge with TMDs resonance at the first harmonic of the walking load may occur.
To ensure the vibration comfort of users, two TMD units are
installed.
In this work, the dynamic performance of the Charleroi foot-
bridge with and without the TMD units is examined in situ on the
basis of an extensive measurement campaign. After installation of
the TMD units, an increase by a factor of 15 is found for the effec-
tive damping ratio of the fundamental mode. Experiments involv-
ing excitation by a limited number of persons show that the TMD
units are effective in reducing the structural response. The
obtained reduction, however, depends highly on the type of human
excitation.
A comprehensive numerical analysis is performed to evaluate
the effectiveness of the damping devices for crowd densities up to
1.5 persons/m2. The results are compared with predictions based on
Fig. 21. The 95th percentile of the maximum acceleration levels of the
the simplified load model proposed in the Setra and HiVoSS design
Monte Carlo–simulated pedestrian traffic considering different pedes-
guides. For the footbridge without TMD units, the structural
trian densities for the Charleroi footbridge without (dark bars) and with
response predicted by the simplified load model is significantly
(light bars) TMD units
higher than that in detailed numerical simulations, because the
steady-state conditions assumed by the design guides are not
Effectiveness of the TMD Units attained by pedestrians crossing short-span and lowly damped
bridges. For the footbridge with TMD units, the structural response
Fig. 20 compares the maximum acceleration levels of the footbridge predicted by the simplified load model is considerably lower
with the two TMD units as predicted by the design guides based on because the fundamental mode was highly damped and no longer
the effective damping ratio of 6%, to the 95th percentile value dominated the structural response. In addition, the effective damp-
obtained from the Monte Carlo–simulated pedestrian traffic. These ing ratio overestimated the effectiveness of the damping devices
results show that a significantly lower structural response is pre- because crowd-induced loading is near-harmonic and transient, not
dicted by applying the simplified load model. The latter predomi- harmonic.
nantly results from the fact that the design guides assume that the The effectiveness of the TMD units predicted by the detailed nu-
resonant mode dominates the structural response. The present anal- merical analysis is in excellent agreement with the in situ observa-
ysis shows, however, that this is no longer the case when the reso- tions for free and synchronized walking.
nant mode is highly damped. In addition, the effectiveness of the The main conclusion of this investigation is that a simplified pro-
TMD units is overestimated by the effective damping ratio, as cedure based solely on the contribution of the resonant mode is no
crowd-induced loading is near-periodic and transient as opposed to longer appropriate for evaluating the dynamic performance of a
harmonic (Weber 2013). footbridge with TMDs. A more detailed analysis that accounts for
The effectiveness of the TMD units, therefore, is evaluated more both the variability of pedestrian excitation and the dynamic behav-
accurately by comparing the 95th percentile value of the maximum ior of the coupled system is required.
acceleration levels of Monte Carlo–simulated pedestrian traffic for
the footbridge with and without TMD units. This comparison is pre-
sented in Fig. 21, which illustrates that the effectiveness of TMD Acknowledgments
units is much larger for dense traffic conditions. Analysis shows
that for sparse and dense pedestrian densities, the structural This research is funded by the Agency for Innovation by Science
responses are reduced by factors of approximately 1.3 and 5.2, and Technology in Flanders (IWT); this financial support is
respectively. This difference in reduction is explained by the fact gratefully acknowledged.
that for sparse pedestrian densities, the relative contribution of the
higher (lowly damped) modes is much higher than for dense traffic References
conditions. The reduction is slightly different for different pedes-
trian densities, because the optimality of the tuning of TMD units Ainsworth, I., Franklin, K., and Burnton, P. (2011). “Kurilpa bridge—A
also depends on the considered pedestrian density. Note that the case study.” Proc., 4th Int. Footbridge Conf., L. Debell and H. Russel,
attained reductions are in excellent agreement with those identified eds., Wroclaw, Poland.

© ASCE C4016001-12 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


Asami, T., Nishihara, O., and Baz, A. (2002). “Analytical solutions to Kuo, A. D. (2001). “A simple model of bipedal walking predicts the preferred
Hinfinity and H2 optimization of dynamic vibration absorbers attached speed-step length relationship.” J. Biomech. Eng., 123(3), 264–269.
to damped linear systems.” J. Vib. Acoust., 124(2), 284–295. Li, Q., Fan, J., Nie, J., Li, Q., and Chen, Y. (2010). “Crowd-induced random
Association Française de Genie Civil. (2006). Setra: Evaluation du com- vibration of footbridge and vibration control using multiple tuned mass
portement vibratoire des passerelles pietonnes sous l’action des pietons, dampers.” J. Sound Vib., 329(19), 4068–4092.
AFGC, Paris (in French). Marano, G., Greco, R., and Sgobba, S. (2010). “A comparison between dif-
Bertram, J. E. A., and Ruina, A. (2001). “Multiple walking speed— ferent robust optimum design approaches: Application to tuned mass
Frequency relations are predicted by constrained optimization.” J. dampers.” Probab. Eng. Mech., 25(1), 108–118.
Theor. Biol., 209(4), 445–453. MATLAB. (2014). Statistics Toolbox (R2014a). MathWorks Inc., Natick,
Bocian, M., Macdonald, J. H. G., and Burn, J. F. (2012). “Biomechanically MA.
inspired modelling of pedestrian-induced forces on laterally oscillating Maurer-Sohne. (2013). Extract of reference list for tuned mass dampers.
structures.” J. Sound Vib., 331(16), 3914–3929. Reference list, Munich, Germany.
Bocian, M., Macdonald, J. H. G, and Burn, J. F. (2013). “Biomechanically- Meinhardt, C. (2012). “Detailed numerical and experimental dynamic anal-
inspired modeling of pedestrian-induced vertical self-excited forces.” J. ysis of long-span footbridges to optimize structural control measures.”
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Bridge Eng., 10.1061/(ASCE)BE.1943-5592.0000490, 1336–1346. Proc., 6th Int. Conf. on Bridge Maintenance, Safety and Management,
Brownjohn, J., Fok, P., Roche, M., and Omenzetter, P. (2004). “Long span Taylor & Francis.
steel pedestrian bridge at Singapore Changi Airport—Part 2: Crowd Occhiuzzi, A., Spizzuoco, M., and Ricciardelli, F. (2008). “Loading models
loading tests and vibration mitigation measures.” Struct. Eng., 82(16), and response control of footbridges excited by running pedestrians.”
28–34. Struct. Control Health Monit., 15(3), 349–368.
Busca, G., Cappellini, A., Manzoni, S., Tarabini, M., and Vanali, M. Papadimitriou, C., Katafygiotis, L. S, and Au, S.-K. (1997). “Effects of
(2014). “Quantification of changes in modal parameters due to the pres- structural uncertainties on TMD design: A reliability-based approach.”
ence of passive people on a slender structure.” J. Sound Vib., 333(21), J. Struct. Control, 4(1), 65–88.
5641–5652. Pedersen, L. (2015). “Implications of interaction between humans and
Butz, C., Feldmann, M., Heinemeyer, C., and Sedlacek, G. (2008). structures.” Proc., IMAC 33, Int. Modal Analysis Conf , Springer.
“SYNPEX: Advanced load models for synchronous pedestrian excita- Peeters, B., and De Roeck, G. (1999). “Reference-based stochastic subspace
tion and optimised design guidelines for steel footbridges.” Technical identification for output-only modal analysis.” Mech. Syst. Sig. Process.,
Rep., Research Fund for Coal and Steel, Brussels, Belgium. 13(6), 855–878.
Butz, C., Schuermann, C., and Benicke, O. (2011). “Tuned mass dampers Piccardo, G., and Tubino, F. (2012). “Equivalent spectral model and maxi-
for the footbridge of VW Autostadt in Wolfsburg, Germany.” Proc., 4th mum dynamic response for the serviceability analysis of footbridges.”
Int. Footbridge Conf., J. Biliszczuk, J. Bien, P. Hawryszków, and T. Eng. Struct., 40, 445–456.
Kaminski, eds., Wroclaw, Poland. Reynders, E., and De Roeck, G. (2008). “Reference-based combined deter-
Caetano, E., Cunha, A., Magalhães, F., and Moutinho, C. (2010). “Studies ministic–stochastic subspace identification for experimental and opera-
for controlling human-induced vibration of the Pedro e Inõs footbridge, tional modal analysis.” Mech. Syst. Sig. Process., 22(3), 617–637.
Portugal. Part 2: Implementation of tuned mass dampers.” Eng. Struct., Reynders, E., Pintelon, R., and De Roeck, G. (2008). “Uncertainty bounds
32(4), 1082–1091. on modal parameters obtained from stochastic subspace identification.”
Caprani, C. C. (2014). “Application of the pseudo-excitation method to Mech. Syst. Signal Process., 22(4), 948–969.
assessment of walking variability on footbridge vibration.” Comput. Sahnaci, C., and Kasperski, M. (2011). “Simulation of random pedestrian
Struct., 132, 43–54. flow.” Proc., 8th Int. Conf. on Structural Dynamics of EURODYN, De
Caprani, C. C., Keogh, J., Archbold, P., and Fanning, P. (2012). Roeck, G. Degrande, G. Lombaert, and G. Müller, eds.,. Leuven, Belgium.
“Enhancement for the vertical response of footbridges subjected to sto- Setareh, M., Ritchey, J. K., Baxter, A. J., and Murray, T. M. (2006).
chastic crowd loading.” Comput. Struct., 102–103, 87–96. “Pendulum tuned mass dampers for floor vibration control.” J. Perform.
Carroll, S. P., Owen, J. S., and Hussein, M. F. M. (2012). “Modelling Constr. Facil., 10.1061/(ASCE)0887-3828(2006)20:1(64), 64–73.
crowd-bridge dynamic interaction with a discrete defined crowd.” J. Tubino, F., and Piccardo, G. (2015). “Tuned mass damper optimization for
Sound Vib., 331(11), 2685–2709. the mitigation of human-induced vibrations of pedestrian bridges.”
Chopra, A. K. (1995). Dynamics of structures, Prentice Hall, Upper Saddle Meccanica, 50(3), 809–824.
River, NJ. Van den Broeck, P., and De Roeck, D. G. (2012). “Tricon: Prediction and
Colombel, M. (2014). “Montage des ADAs sur la passerelle.” Tech. rep., control of human-induced vibrations of civil engineering structures.”
GERB Engineering, Berlin. Van Nimmen, K., Lombaert, G., De Roeck, G., and Van den Broeck, P.
Den Hartog, J. P. (1985). Mechanical vibrations, Dover Publications, (2014a). “Vibration serviceability of footbridges: Evaluation of the cur-
Mineola, NY. rent codes of practice.” Eng. Struct., 59, 448–461.
Dougil, J. W., Wright, J. R., Parkhouse, J. G., and Harrison, R. E. (2006). Van Nimmen, K., Lombaert, G., Jonkers, I., Roeck, D., and Van den
“Human structure interaction during rhythmic bobbing.” Struct. Eng., Broeck, G. P. (2014b). “Characterisation of walking loads by 3D inertial
84(2), 32–39. motion tracking.” J. Sound Vib., 333(20), 5212–5226.
Ellis, B. R., and Ji, T. (2004). “Loads generated by jumping crowds: Van Nimmen, K., Maes, K., Živanocic, S., Lombaert, G., De Roeck, G., and
Numerical modelling.” Struct. Eng., 82(17), 35–40. Van den Broeck, P. (2015). “Identification and modelling of vertical
GERB Engineering. (2014). Tuned mass dampers. Reference list, Berlin. human–structure interaction.” Proc., IMAC 33, Int. Modal Analysis Conf.
Heinemeyer, C., et al. (2009). “Design of lightweight footbridges for Varela, W. D., and Battista, R. C. (2011). “Control of vibrations induced by
human induced vibrations—Background document in support to the people walking on large span composite floor decks.” Eng. Struct.,
implementation, harmonization and further development of the 33(9), 2485–2494.
Eurocodes.” hhttp://www.stb.rwth-aachen.de/projekte/2007/HIVOSS Viguie, R. (2010). “Tuning methodology of nonlinear vibration absorbers
/download.phpi. coupled to nonlinear mechanical systems.” Ph.D. thesis, Univ. of Liège,
Ingólfson, E. T., and Georgakis, C. T. (2011). “A stochastic load model for Liège, Belgium.
pedestrian-induced lateral forces on footbridges.” Eng. Struct., 33(12), Wang, D., Gao, S., Kasperski, M., Liu, H., and Jon, L. (2011). “Simulation
3454–3470. of the dynamic characteristics of the coupled system structure.” Appl.
Jones, C. A., Reynolds, P., and Pavic, A. (2011). “Vibration serviceability Mech. Mater., 71–78, 1507–1510.
of stadia structures subjected to dynamic crowd loads: a literature Weber, B., and Feltrin, G. (2010). “Assessment of long-term behavior of
review.” J. Sound Vib., 330(8), 1531–1566. tuned mass dampers by system identification.” Eng. Struct., 32(11),
Kasperski, M., and Czwikla, B. (2012). “A refined model for human 3670–3682.
induced loads on stairs.” Topics on the dynamics of civil structures. Weber, F. (2013). “Dynamic characteristics of controlled MR-STMDs of
Proc., Society for Experimental Mechanics Series, Springer, 27–39. Wolgograd Bridge.” Smart Mater. Struct., 22(9), 095008.

© ASCE C4016001-13 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001


Weber, F., Feltrin, G., and Olaf, H. (2006). “Samco final report—F05 guide- Živanovic, S., Diaz, I., and Pavic, A. (2009). “Influence of walking and
lines for structural control.” Structural Engineering Research Laboratory, standing crowds on structural dynamic properties.” Proc., IMAC 27, Int.
Swiss Federal Laboratories for Materials Testing and Research. Modal Analysis Conf., Springer.
Živanovic, S. (2012). “Benchmark footbridge for vibration serviceability Živanovic, S., Pavic, A., and Ingólfsson, E. (2010). “Modeling spatially
assessment under the vertical component of pedestrian load.” J. Struct. unrestricted pedestrian traffic on footbridges.” J. Struct. Eng., 10.1061
Eng., 10.1061/(ASCE)ST.1943-541X.0000571, 1193–1202. /(ASCE)ST.1943-541X.0000226, 1296–1308.
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

© ASCE C4016001-14 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4016001

Das könnte Ihnen auch gefallen