Sie sind auf Seite 1von 13

Design of Footbridges against Pedestrian-Induced Vibrations

F. Ricciardelli, Ph.D., P.Eng., M.ASCE1; and C. Demartino, Ph.D., P.Eng.2

Abstract: In 1999 and 2000, the two vibration incidents at the Paris Passerelle Solferino Bridge and the London Millennium Bridge triggered
a major revision of existing knowledge concerning footbridge response to pedestrian-induced actions. In the last 15 years, an incredibly large
amount of research has emerged on the topic. Although researchers have provided many valuable scientific contributions regarding the under-
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

standing and modeling of pedestrian-induced vibrations of footbridges, there is still a need to determine what real improvements have been
achieved in design procedures. This article provides a critical overview of the methodologies proposed over the last four decades, as well as
their code implementations, and summarizes the actual advances available that consultants can use to their advantage in bridge design.
DOI: 10.1061/(ASCE)BE.1943-5592.0000825. © 2015 American Society of Civil Engineers.
Author keywords: Footbridges; Walking-induced vibrations; Crowd–structure interaction; Walking loading models; Response models.

Introduction compared (Ingólfsson et al. 2012; Venuti and Bruno 2009;


Živanovic et al. 2005; Racic et al. 2009). Many of these models
Researchers have known of the problem of the evaluation of the have also become part of design procedures for footbridges.
dynamic response of footbridges to pedestrian loading at least Design of footbridges against pedestrian-induced vibrations
since the work of Tilden (1913), more than a century ago. requires (a) knowledge of the characteristics of the pedestrian action
Nevertheless, for decades, most of the design procedures and code (b) a method of analysis for pedestrian-induced vibration, and (c) a
specifications have been based on the simplifying assumption that comfort criterion (Wheeler 1982).
walkers apply only a static vertical load to the footbridge. In fact, Fig. 1 shows the classification of the methods of analysis for pe-
pedestrian movement produces a complex system of forces, acting destrian-induced vibrations adopted in this article. A two-way con-
in three directions and varying in time and space; these depend on nection exists between pedestrian action and pedestrian-induced
the characteristics of the pedestrian, but are also modified by the vibration, due to the feedback of the footbridge vibration resulting
interaction with other pedestrians and by the perception of the from the pedestrian behavior. The characteristics of the pedestrian
structural motion. force can be studied using experimental data or analytical models
The approach to the analysis of footbridge response to pedes- and can be described in terms of deterministic or stochastic models.
trian loading dramatically changed at the turn of the millennium. Furthermore, the forcing function can be expressed in the time or fre-
The Passerelle Solferino Bridge (now Passerelle Leopold-Sedar- quency domain. The methods of analysis for pedestrian-induced
Senghor Bridge) in Paris on December 15, 1999, and the vibrations have been classified into linear models, further subdivided
Millennium Bridge in London on July 10, 2000, both on their into stability criteria and response evaluation methods, and nonlinear
opening days, experienced large lateral vibrations, the causes of models, further subdivided into parametric excitation models,
which are clearly found in the interaction between the dynamics of inverted pendulum models, nonlinear damping models, and response
the structures and that of the walkers. Researchers recognized that feedback models (Ingólfsson et al. 2012). Moreover, models can
the action exerted by a walker on a footbridge is modified by the consider or neglect interpedestrian and pedestrian–structure inter-
footbridge vibration. The main effect of pedestrian-induced vibra- action. The output of the models is either the magnitude of the struc-
tions is to cause discomfort, thus impairing serviceability. In recent tural response (usually in terms of maximum acceleration) or the
years, dozens of papers have appeared that deal with the phenom- critical number of pedestrians causing diverging vibration ampli-
enon of crowd–structure interaction, with application to foot- tudes (Ingólfsson et al. 2012).
bridges; these also contain a number of loading and response eval- Finally, tools are needed to establish human tolerance to vibra-
uation models able to account for such interaction. Also, review tion. Tredgold (1825) first gave a commonsense tolerance criterion
papers have appeared, in which these models are summarized and for vibrations, suggesting that “girders over long spans should be
made deep to avoid the inconvenience of not being able to move on
the floor without shaking everything in the room.” Vibration toler-
1
Dept. of Civil, Construction and Environmental Engineering, Second ance depends on the activity carried out during vibration and on the
Univ. of Naples, Via Roma 29, 81031 Aversa (CE), Italy; formerly, Dept. vibration frequency. Usually, reference is made to the acceleration
of Informatics, Infrastructures and Sustainable Energy, Univ. of Reggio limits contained in ISO 10137 (ISO 2007), adjusted to the intended
Calabria, Via Graziella, Feo di Vito, 89122 Reggio Calabria, Italy. E-mail: occupancy. This article does not discuss vibration tolerance.
friccia@unina2.it What instead seems to be missing is a comparison among the
2
Dept. of Structures for Engineering and Architecture, Univ. of Naples recent proposals and a discussion of the differences they bring with
Federico II, Via Claudio 21, 80125 Naples, Italy (corresponding author).
respect to traditional approaches, not accounting for interaction.
E-mail: cristoforo.demartino@unina.it
Note. This manuscript was submitted on December 29, 2014; approved Because a full comparison is not achievable, as the models differ in
on June 24, 2015; published online on January 20, 2016. Discussion period terms of background hypotheses, fields of applicability, and data
open until June 20, 2016; separate discussions must be submitted for indi- used for calibration, this article’s view is that it would be of interest
vidual papers. This paper is part of the Journal of Bridge Engineering, © for the consultant to have some guidance on the choice of the most
ASCE, ISSN 1084-0702. appropriate approach to use in different situations.

© ASCE C4015003-1 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Classification of the methods of analysis for pedestrian-induced vibrations

This article attempts to compare background hypotheses, field of


applicability, and results obtained through a number of different load-
ing and response evaluation models, with the analysis limited to walk-
ing action. Only models having immediate implications for design
practice are considered. More advanced approaches that don’t seem to
already have practical applicability are not presented or discussed. The
different approaches are compared without any application to a specific
case study, with the aim of obtaining general conclusions, not related
to a specific situation. In fact, other articles that compare the results of
the application of design procedures already exist, yet these do not dis-
cuss or compare the different approaches adopted. Thus, this article
provides a synthesis of the results available and a description of how
these can be used in the design of footbridges.

Walking Force Modeling

Ground reaction forces (GRFs) are the forces exerted by a walker Fig. 2. GRFs exerted by a walker to a footbridge and discrete Fourier
on the walking surface. For example, the vertical GRF exerted by a series representation of the vertical component
walker on a footbridge is shown in Fig. 2. GRFs are characterized
by a rather large level of randomness, which is the result of intersub-
ject variability (i.e., the differences existing between different
walkers) and intrasubject variability (i.e., the nonperiodic behavior
of each walker) (Eriksson 1994; Ohlsson 1982). In addition, GRFs mean of 2 Hz and a standard deviation of approximately 0:20 Hz
are modified by the interaction among walkers and of the walkers (Matsumoto et al. 1972; Ricciardelli and Pansera 2010). One possi-
with the structure (as discussed later in the article). ble approach to the modeling of GRFs is to neglect intrasubject vari-
Although it is not fully periodic, walking is a rather repetitive ac- ability, and assume that the force is periodic in time and moving in
tivity, and so are the GRFs. Natural (or undisturbed) walking fre- space at a constant speed; in so doing, the vertical and lateral com-
quency is reasonably described by a Gaussian distribution with a ponents can be written as

" #   
X1
L
FV ðx; tÞ ¼ W 1 þ DLFV; j ðsin ð j  2  p  fw  t  t Þ  f V; j Þ d ðx  vðt  t ÞÞ Hðt  t Þ  H t  t 
j¼1
v
" #   
X1
L
FL ðx; tÞ ¼ W DLFL;j ðsin ðj  p  fw  t  t Þ  f L; j Þ d ðx  vðt  t ÞÞ Hðt  t Þ  H t  t 
j¼1
v (1)

where W is the weight of the walker; DLFV;j and DLFL;j are the jth x is the position of the pedestrian on the bridge and v ¼ fw l is the pe-
dynamic load factors (DLFs); fw is the walking frequency; f j is the destrian speed, with l being the stride length; HðÞ is the Heaviside
phase lag of the jth harmonic; d ðÞ indicates the Dirac delta function; function; t is arrival time; and L is the span length. The equations in

© ASCE C4015003-2 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


Eq. (1) lend themselves to the evaluation of the transient response due shape. For a simple supported beam, the maximum acceleration in
to single or multiple walkers. In the case of multiple walkers, intersub- the first mode was given as
ject variability can be accounted for during superposition. It is worth
nothing that the data concerning these parameters mainly come from €y max ¼ ð2p f1 Þ2  ys  w ð j 1 ; LÞ (5)
biomechanics and only recently from structural engineering (Butz et
al. 2007; Venuti and Bruno 2009; Racic et al. 2009); independently where ys ¼ WL3 =48EI is the static deflection due to the weight of
from the source, these data show large scatter in the results due to the the walker located at midspan, j 1 is the damping ratio in the first
complex and variable (inter- and intrasubject) walking biomechanics. mode, and w ð j 1 ; LÞ [the symbol c was originally used by
On the other hand, the GRFs can be described in the frequency Blanchard et al. (1977), which has been modified here to make it
domain through their power spectral density functions (PSDFs), consistent with the remaining equations] is a factor accounting for
which are suitable for the evaluation of the stationary response to a the transient dynamic behavior; this was evaluated for a simply sup-
stream of walkers, including intra- and intersubject variability. ported beam through numerical integration of the equation of
Such an approach was introduced by Brownjohn et al. (2004b) for motion, and is given in the form of a graph for four different values
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

the vertical GRF and by Pizzimenti and Ricciardelli (2005) and of j 1 , and for L in the range of 10–50 m. The model of Blanchard et
Ricciardelli and Pizzimenti (2007) for the lateral GRF. al. (1977) is a simple correction to a static load, and was easily
Brownjohn et al. (2004b) obtained the PSDFs of the first six ver- incorporated into codes of practice (discussed later in this article).
tical DLFs from the Fourier spectra of measured forces; converted From a mathematical point of view, the same approach can be
to GRF, this becomes applied, and indeed was applied, also to the lateral response; in that
h  i2 case, however, the idea of the correction of a static deflection loses
j f =ð jfw Þ1jCV; j
AV; j þ BV; j exp DV; j its physical meaning. What the authors failed to recognize is that
SFV; j ð f Þ ¼ 32  W 2 (2) the maximum transient response does not depend on L and j 1 sepa-
fw
rately, but rather on their product (Ricciardelli and Briatico 2011).
Nevertheless, the graphs they produced are quite accurate and con-
where AV;j , BV;j , CV;j , and DV;j are curve-fitting parameters. For
each harmonic, f is in the range of ðj  0:25Þ  fw –ðj þ 0:25Þ  fw . sistent with the fact that different pairs of L and j j having the same
Živanovic et al. (2007) extended the results of Brownjohn et al. product must produce the same value of w .
(2004b) by considering also the DLFs of the first five subharmonics Generally speaking, the maximum acceleration in the ith mode
of the walking force. in resonant conditions (i.e. jfw ¼ fi ) can be written as
Pizzimenti and Ricciardelli (2005) and Ricciardelli and Pizzimenti DLFj  W
(2007) obtained the PSDFs of the lateral GRFs around the first five har- €y max;i;j ¼  w ðɛi Þ ¼ €y stat;i;j  w ðɛi Þ (6)
2 j i mi
monics of the walking frequency as
2 !2 3 where ystat;i;j ¼ ðDLFj  WÞ=ð2 j i mi Þ is the stationary acceleration
F~ L; j 2AL;j
2
f =ð j  fw Þ  1 5
SFL; j ð f Þ ¼ pffiffiffiffiffiffiffi exp 42 (3) amplitude in the ith mode due to the jth resonant harmonic, mi is the
f 2p  BL;j BL; j ith modal mass, and w ðɛi Þ is a transient resonant response
Coefficient (TRRC) that accounts for the load motion. The TRRC is
where F~ L;j is the characteristic value of the area of the PSDF around
2 a stationarity index, and has been found to depend on the parameter
the jth harmonic, AL; j is a parameter used to normalize the PSDF, (Ricciardelli and Briatico 2011), as follows:
and BL;j is a bandwidth parameter. For each harmonic, f is in the 1 2L 1
range of 0:9  j  fw –1:1  j  fw . ɛi ¼   ji ¼  n  ji (7)
i l i

Transient Response to a Single Crossing where n is the number of effective load cycles. In particular, large
values of ɛi are associated with long-span footbridges, a short stride
The simplest approach to the analysis of walking-induced foot- length, a low vibration mode, and high damping, in which case the
bridge vibrations is that of evaluating the transient response to the TRRC approaches 1. On the other hand, small values of ɛi are asso-
crossing of a walker in resonance with one of the bridge vibration ciated with short footbridges, a long stride length, a high vibration
frequencies. This condition is likely to occur for footbridges hav- mode, and low damping, in which case the TRRC is close to 0 and
ing vertical vibration frequencies lower than approximately 5 Hz the response highly nonstationary. Nonetheless, few researchers
or horizontal vibration frequencies lower than approximately recognized ɛi as the governing parameter for the TRRC; instead,
2:5 Hz, and may prove to be the worst load scenario. No feedback many used the variables n and j i separately when studying the
of the footbridge vibration to the walking parameters is consid- transient response. In the following discussion, the maximum
ered, as discussed later in the article. acceleration derived by different authors will be written in the
The first study that recognized the crossing of a resonant walker form of Eq. (6).
as a critical condition is that of Blanchard et al. (1977). The action Blanchard et al. (1977) did not mention, and therefore seem not
of the walker was modeled as a vertical sinusoidal force acting at to have been aware of, the work of Fryba (1973) published four
the footbridge fundamental frequency, moving at a speed corre- years earlier, in which a closed-form solution of w was given as a
sponding to a stride length l ¼ 1:8 m: function of j i and v; this solution proves rather inaccurate. In partic-
  ular, it is based on the wrong assumption that the maximum acceler-
l ation occurs when the walker crosses midspan, but this is not the
FV ðtÞ ¼ DLFV;1  W  sinð2p f1 tÞ  f f1 t (4)
2 only source of inaccuracy. The solution of Fryba (1973) does not
acknowledge the fact that the TRRC depends on ɛi , as it expresses it
where W ¼ 700 N, f1 is the footbridge fundamental frequency, as function of L and j separately; in fact, it can be manipulated to
DLFV;1 ¼ 0:257, and f ðxÞ is the normalized fundamental mode take the form

© ASCE C4015003-3 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


   
ɛi p light-gray area shows the typical range of ɛ1 for common foot-
w ðɛi Þ ¼ exp  ɛi þ ɛi (8)
1 þ ɛ2i 2 bridges. The numerical solution of Blanchard et al. (1977) is in quite
good agreement with the closed-form solution of Ricciardelli and
Eleven years later, still not realizing the dependency of w on ɛi , Briatico (2011). The solution of Fryba (1973) is less accurate and
Rainer et al. (1988) gave a graph containing values of w as a function on the unsafe side for all values of ɛ1 . The solutions of Rainer et al.
of n for different values of j 1 ; they did not explain how the values (1988) and of Young (2001) are also less accurate, but on the safe
were derived, but a later article of Pimentel and Fernandes (2002) side for all values of ɛ1 . The solutions of Allen and Murray (1993)
giving the same results reveals that they must have used the approxi- and of Grundmann et al. (1993) are on the safe side for small values
mated equation of the displacement of a simple supported beam of ɛ1 and on the unsafe side for large values of ɛ1 , and have an
crossed by a moving sinusoidal load given by Fryba (1973). The val- unjustified limitation for large values of ɛ1 . Finally, the response of
ues found differ from those of Blanchard et al. (1977), but differ also a single degree of freedom (SDOF) system loaded by a transient
from Eq. (8), as the assumption that the maximum acceleration harmonic load stationary at midspan and acting for a duration equal
to the crossing time T ¼ L=v gives a solution largely on the safe
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

occurs when the walker is at midspan was removed, and the exact
time when this occurs was evaluated numerically. Nevertheless, side for all values of ɛ1 . This model was adopted by some codes of
the solution of Rainer et al. (1988) also proves inaccurate, prob- practice (discussed later in this article).
For footbridges having more than one span, numerically
ably due to the inaccuracy of the response equation it is based on.
Based on the measurements carried out on two footbridges, obtained correction factors exist. In particular, Blanchard et al.
(1977) proposed a correction coefficient for up to three spans;
Allen and Murray (1993) suggested that the constant value w ¼ 0:7
Pimentel and Fernandes (2002) proposed a correction coefficient
can be used, regardless of the bridge characteristics.
for the case of two spans, to be used in the model of Rainer et al.
Grundmann et al. (1993) suggested the approximated value
(1988). The coefficients proposed are in agreement with each other
w ðn; j 1 Þ ¼ 0:6½1  exp ð2p n j 1 Þ (9) and are for reducing factors: For the case of two spans, they are in
the range of 0:7–0:96 for ratio between the lengths of the spans in
derived from interpolation of the solution of the modal equation of the range of 1–0:2.
motion (in square brackets), which was then reduced to 60%. This
model was the first to acknowledge the dependency of w on Response to Multiple Walkers
ɛ1 ¼ n j 1 , but the reason for the reduction was not explained.
For floors, Young (2001) proposed an equation for the TRRC An alternative situation is that of a footbridge crossed by more than
similar to that of Grundmann et al. (1993), as follows: one walker. In that case, the characteristics of the walkers have to
be described on a probabilistic basis, and the response can be eval-
w ðn; j 1 Þ ¼ 1  exp ð2p  0:275n j 1 Þ (10)
uated either in the frequency domain through random vibration
theory, or in the time domain through Monte Carlo simulations.
Finally, Ricciardelli and Briatico (2011) found an approximated The first attempt to evaluate the load exerted by a number N of
closed-form for the TRRC, as follows: walkers is that of Matsumoto et al. (1978), who noted that when the
qffiffiffiffiffiffiffiffiffiffiffiffiffi    
footbridge is crossed by a stream of walkers having equal frequency
ɛi 2 þ exp ɛ  p þ arctan 1
w ðɛi Þ ¼ 1 þ ɛ i and phases uniformly distributed between 0 and 2p (uncorrelated
1 þ ɛ2i i
2 ɛi walkers), the RMS response varies with the square root of N (or with
(11) the square root of the walker density d ¼ N=BL, B being the breadth
of the deck). An opposite situation is that of fully correlated walkers
which proves quite accurate, with an error with respect to the nu- (i.e., walkers stepping in phase), for which the RMS response varies
merical solution never exceeding 6  103 for ɛi in the broad range with N.
of 104 –10. From experiments on a 10-m footbridge, Eriksson and Ohlsson
In Fig. 3 the values of w ðɛ1 Þ proposed by different authors are (1988) found that when the walkers step in phase the total vertical
compared, indicating a clear disagreement among one another. The response is obtained multiplying the RMS response to a single

Fig. 3. Comparison of the TRRC from different authors; the light-gray area shows the typical range of ɛ1 for common footbridges

© ASCE C4015003-4 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


pffiffiffiffi
walker by N 0:9 for the first harmonic, and by N for higher harmon- ratio; TR is the return period; m1 is the actual first modal mass of the
ics; when the walkers are free to choose their walking rate, they tend bridge; λ is the actual pedestrian flow rate; and b j , b L , and b U are
to be fully correlated for N < 5 and tend to lose correlation for a corrective parameters that depend on the actual damping, bridge
larger number; it is suggested that for N > 5 the effective number of length, and mode shape, respectively.
pedestrians is 4:5 þ ðN=10Þ. This result is rather different from that
of Matsumoto et al. (1978), and is possibly the effect of the very short
span of the footbridge and of the experimental approximations. Interaction
Combining Eq. (2) and a Gaussian probability-density function
(PDF) of the walking frequency, Brownjohn et al. (2004b) derived Two types of interaction are of interest when dealing with the load-
an approximated equation for the PSDFs of the first six harmonics ing of structures by humans; the first is that taking place among the
of the vertical force exerted by a group of N uncorrelated walkers, different individuals and the second is that taking place between an
which can be put in the form individual and the structure. Often, these interactions can be
"  2 # neglected, but it is becoming more and more common to include
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

2  f them. It is now widely accepted that some kind of inter-walker and


f  SFVej ðf Þ ¼ N  W  DLFVj  f  exp 12:5  2
2
(12)
j walker–structure interaction almost inevitably occurs.

where f ¼ f =1 Hz. Inter-Walker Interaction


Similarly, through convolution of Eq. (3) and a Gaussian PDF of
the walking frequency, Ricciardelli (2005) derived an approximated Inter-walker interaction can be defined as the violation of the
equation for the PSDFs of the first five harmonics of the horizontal superposition principle: Two or more walkers who interact with
force exerted by a group of N uncorrelated walkers, bringing each other apply a load that is not the sum of those they would
" apply if they were alone. The occurrence of such interaction is
 2 #
f related to the perception that each walker has of the surrounding
f  SFLej ðf Þ ¼ N  W  kj  exp 62  1
2
(13) walkers, and to the influence that this has on the walker’s
j
behavior.
It is a quite complicated topic, to be investigated with the help
where k1 ¼ 6:6  103 , k2 ¼ 3:2  104 , k3 ¼ 2:2  103 , k4 ¼ 2:3 of knowledge and tools from social and medical sciences. The
104 , and k5 ¼ 4:78  104 . main parameter governing such interaction is the walker density d,
Eqs. (12) and (13) can be used to calculate the peak acceleration
which characterizes a uniform crowd. Indeed, that of a uniform
in the ith mode. In the worst-case scenario, in which the ith vibra-
crowd is just one case, and other cases of interest are that of single
tion frequency coincides with the jth load harmonic, and under the
individuals interacting when their distance reduces below a thresh-
assumption that the mode shape is approximated as a sinusoid, then
old value, and that of pairs and groups of individuals walking to-
!0:5 gether. Although very interesting research has been and is pres-
p fi  SFej ðfi Þ ently being carried out, mainly aiming to model the different types
^a i ¼ ga  (14)
8 j i m2i of interactions, it does not seem that any significant and definitive
result has been reached that can be used to advantage in engineer-
where ^a i is the peak acceleration in the ith mode (either vertical or ing practice.
horizontal), fi  SFej ðfi Þ is the spectrum of the jth harmonic For a stationary and homogeneous walker, the flow equation
pffiffiffi of the ith
modal force (either vertical or horizontal), and ga ¼ 2 is the peak holds (Venuti and Bruno 2009)
factor.
A different approach is that of Butz (2006), who carried out q¼d v (16)
Monte Carlo simulations for different values of the walker density
and for different bridge geometries, to derive the 95th-percentile In this context, a critical density d c , a capacity density d ca , and a
peak modal acceleration. Step frequency, pedestrian mass, force jam density d m can be defined. For d < d c , walking is undisturbed;
magnitude, and pedestrian arrival time were selected on a probabil- for d c < d < d ca , the walking speed decreases with increasing
istic basis. This model was incorporated in human-induced vibra- density; for d ca < d < d m , congestion occurs; and for d ¼ d m , the
tion of steel structures (HIVOSS) guidelines (Heinemeyer et al. maximum admissible density is reached and the walkers stop. For
2009), and is described later in the article. d c , values are available in the literature—for example, d c ¼
Ingólfsson et al. (2008) proposed a response spectrum method 0:2 walkers=m2 (Nakamura 2004), d c ¼ 0:2–0:5 walkers=m2
inspired by earthquake engineering. Through Monte Carlo simula- (Butz et al. 2007; Oeding 1963), d c ¼ 0:3–0:6 walkers=m2
tions, they evaluated a reference vertical acceleration and then cor- (Piccardo and Tubino 2012).
rected it through empirical factors. The reference case is a bridge A more sophisticated approach is that of Butz et al. (2007), who
with modal mass m1;o ¼ 100 ; 000 kg, j 1;o ¼ 0:005, Lo ¼ 50 m, carried out tests on a platform 30 m long and found a relationship
loaded by two reference populations with step frequencies equal to between the mean and standard deviation of the walking frequency
either 1:8 or 2:0 Hz, with a standard deviation of 0:1 Hz and and the walker density (Fig. 4), as follows:
Poisson-distributed arrivals with a mean arrival rate of 1 walker=s.
The peak modal acceleration is m fw ¼ 0:79  ½0:13  d 2  0:82  d þ 1:80 þ 0:79
m1;o pffiffiffi s fw ¼ 0:08  ½0:13  d 2  0:82  d þ 1:80  0:03 (17)
^aVi ðTR ; rÞ ¼ ^
aRi ðTR ; rÞ λb j b L b U (15)
m1
These results were derived using the relationship between the
where ^aRi ðTR ; rÞ is the reference acceleration found using extreme average walking velocity and density given by Oeding (1963), and
value analysis of the simulated response; r ¼ f1 =fw is the frequency apply for 0:2  d  1:5.

© ASCE C4015003-5 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Variation of the mean and standard deviation of the walking frequency with the walker density, d (data from Butz et al. 2007)

Walker–Structure Interaction synchronize to the lateral vibration if the walking frequency is


no more than 0:1 Hz away from the vibration frequency, and
Walker–structure interaction is the feedback of the structural vibra-
confirmed that the DLFL;1 increases with increasing amplitude
tion to the walker dynamics. The straightforward way to model it
of vibration. They did not find any stable synchronization with
would be that of describing the walking action as a function of the
vertical vibrations, even at large amplitudes of approximately
footbridge displacement and velocity; nevertheless, following a
10 mm; limited synchronization was found at larger ampli-
common approach in aeroelasticity, it is more convenient to model
tudes, although these are outside of the serviceability limits of
it as a modification of the structural properties of inertia and
footbridges.
damping.
Similar results were found by Nakamura et al. (2008), who
Walker–structure interaction manifests itself in a totally differ-
observed an increase of the lateral GRF with increasing amplitude
ent way depending on whether vibration occurs in the vertical or lat-
of vibration: At 1:0 Hz, the lateral force can be 10% of the weight
eral direction. For lateral vibrations, some researchers claim that
at an amplitude of 10 mm, and 16% at an amplitude of 70 mm.
synchronization is the cause of instability (e.g., Okamoto et al.
They also found a probability of synchronization of 20% at a fre-
1992; Fujino et al. 1993), whereas others claim that it is not neces-
sary (e.g., Macdonald 2008; Brownjohn et al. 2004a). In the quency of 0:87 Hz and of 50% at a frequency of 1 Hz, for the
authors’ opinion, what everybody agrees on and what is relevant to amplitudes in the range of 10–70 mm. Sun and Yuan (2008)
structural design is that the walking forces are modified by the foot- found that DLFV;1 is not affected by the lateral motion in the range
bridge vibration. of 0–30 mm, whereas DLFL;1 depends on the lateral amplitude
Rönnquist (2005) found that the DLFL;1 has two different trends, ys ½m as
at resonance and away from resonance. In the first case, synchroni- DLFL;1 ¼ 1:1811  ys þ 0:0517 ys  0:05 m (20)
zation was observed, and DLFL;1 increases with increasing vibra-
tion amplitude until the maximum level of synchronization is
On the other hand, knowledge of the effects of vibration on the
reached. In the latter case, no synchronization was observed, and
DLFs is not sufficient for calculation of the response, as the phase
DLFL;1 linearly increases with vibration amplitude. The first lateral
of the excitation is also needed. Alternatively, one would need to
DLF was given for the resonant and nonresonant cases as
know the in-phase (inertia) and in-quadrature (damping) compo-
DLFL;1 ðf Þ ¼ 0:145 nents of the GRF. To the authors’ knowledge, no data are available
" " for vertical vibrations.
 2 # ! #
For lateral vibrations, Ingólfsson et al. (2011) and Ricciardelli et
1 f  f1
0:1 exp  0:45 þ 1:5 exp  €y max
1:35
(18) al. (2014) provided some experimental results; the additional damp-
2 0:07
ing and mass coefficients were given as

where €y max is expressed in m=s2 . Comparing full-scale observations D  3  2  3


F fs fs fs
and laboratory results, Rönnquist (2005) evaluated the effective jp ¼  ¼ aj þ b j þ c j þ dj
2ð2p fs Þ2 mo xo fw fw fw
number of walkers, that is, the number of fully synchronized    
walkers giving rise to the measured response, as follows: FI br fs
rp ¼ ¼ a r  sin  exp c r þ dr
"   # ð2p fs Þ2 mo xo fs =fw fw
1:6
N (21)
Ne ðNÞ ¼ 35  34  exp  (19)
60
 D and F
where F  I are the in-quadrature and in-phase (with displace-
Butz et al. (2007) measured GRF forces on a laterally or ment) phase components of the motion-induced forces, mo is the
vertically driven platform. They found that walkers potentially mass of the walker, and fs is the floor oscillation frequency, and in

© ASCE C4015003-6 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


which the coefficients depend on the amplitude of oscillation in 5–30 mm; for an amplitude of 10 mm, it is suggested that a value of
millimeters, as follows: 0:4 be used, which is a compromise between the values of 0:28 and
0:46 measured at 0:96 Hz and 0:75 Hz, respectively. Opposed to
aj ðxs Þ ¼ 1  104  y3s þ 5:3  103  y2s  5:69  102  ys  1:77 this, a value of 0:2 is reported in Fujino et al. (1993).
bj ðxs Þ ¼ 2  104  y3s  4:8  103  y2s  0:12  ys þ 7:78 Strogatz et al. (2005) calculated the critical number of walkers
cj ðxs Þ ¼ 5  105  y3s  3:9  103  y2s þ 0:35  ys  9:56 by considering these as oscillators weakly coupled with the foot-
bridge. Assuming a Gaussian PDF of the walking frequency, they
dj ðxs Þ ¼ 4  106  y3s þ 2:6  103  y2s  0:15  ys þ 3:19
found

a r ðxs Þ ¼ 2:5  102  y2s þ 0:66  ys  9:17 pffiffiffiffiffiffiffi m1 f12 j 1 s fw


Nc ¼ 2p  32p 2 (25)
b r ðxs Þ ¼ 1  104  y2s  5:9  103  ys þ 1:11 DLFL1 mo gc
c r ðxs Þ ¼ 4:9  103  y2s  2:29  102  ys þ 2:89
where g is the acceleration of gravity and c ¼ 16 m1 s1 is a coeffi-
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

d r ðxs Þ ¼ 3  104  y2s  2:28  102  ys þ 0:56 (22) cient calibrated based on Millennium Bridge data. The model is ap-
plicable for vibration frequencies in the range 0:75–1:25 Hz, and it
Eq. (21) can be used to calculate the footbridge response is suggested that a value of DLFL1 ¼ 0:044 be used in combination
accounting for walker–structure interaction. with an average body mass of 70 kg and a standard deviation of the
walking frequency of 0:1 Hz.
Roberts (2005) solved the walker–footbridge coupled equations
Stability of motion, setting the displacement of the walker equal to that of the
footbridge. With uniform distribution of fw =f1 between 0:8 and 1:2,
When the footbridge vibrates laterally, walker–structure interaction the critical number of pedestrians turns out to be
can bring dynamic instability.
The full-scale measurements carried out by Arup in the after- c1 m1 j 1
Nc ¼ 4 (26)
math of the Millennium Bridge failure showed that the amplitude of mo
the in-quadrature component of the lateral force was proportional to
the bridge velocity (Dallard et al. 2001), thus having the characteris- where c1 is a coefficient taking into account the averaging of the fre-
tics of a viscous damping force. This led to the now well-known quency ratio; it depends on the footbridge damping, and for practi-
Arup formula, as follows: cal values of this in the range of 0:01–0:02, it varies between 7:27
and 9:06; other values are given in Roberts (2005) in the form of a
m1 f1 j 1 table.
Nc ¼ 8p (23)
k Ingólfsson et al. (2012), based on the criteria of the Service
d’Études Techniques des Routes et Autoroutes (SETRA 2006),
Eq. (23) allows calculation of the number Nc of pedestrians that, Danbon and Grillaud (2005), and Charles and Bui (2005), consid-
when uniformly distributed on a footbridge having a sinusoidal first ered as critical the condition that the acceleration amplitude reaches
mode shape, produces an in-quadrature component of the lateral a limit value a1 . Assuming that the walkers are uncorrelated, at reso-
force that balances the structural damping force, thus producing nance they obtained
instability. The coefficient k in Eq. (23) was found experimentally
to be equal to 300 Ns=m, in a range of vibration frequencies of p2 m21 j 1 a21
Nc ¼ (27)
0:5–1:0 Hz. 10:82 DLFL1
2 m2 g2
o
In the following discussion, other stability criteria that appeared
after that of Eq. (23) will be discussed, and for purpose of compari- It is suggested that the limit acceleration be set equal to
son, these have all been translated in terms of critical number of 0:10 m=s2 , and that for sparse and moderately dense crowds the
pedestrians and put in a form similar to that of Eq. (23) and with the value DLFL1 ¼ 0:05 be used.
same symbols. They all refer to vibrations occurring in the first Finally, Piccardo and Tubino (2008) investigated parametric res-
mode. onance as the cause of instability, and from solution of the Mathieu
Newland (2003) wrote the modal equilibrium equation of the equation they found
footbridge subjected to the inertia force produced by resonant
walkers, assuming that these are all in quadrature with the bridge m1 f12 j 1
Nc ¼ 32p 2 (28)
motion, and obtained an expression of the critical number of b a1 mo g
walkers, as follows:
in which a1 is the slope of the variation of the DLF with the ampli-
m1 j 1
Nc ¼ 4 (24) tude of the oscillation of the footbridge. Results of experiments pre-
ab mo sented in Dallard et al. (2001) suggest a value of a1 ¼ 2 m1 .
Eqs. (23)–(28) contain some similarities, and some differences,
where mo is the mass of each walker, a is the ratio of the amplitude that have been widely discussed in Ricciardelli (2014).
of oscillation of the walker’s center of mass to the amplitude of os- Eqs. (23)–(28) can all be written in the form
cillation of the footbridge, and b is the probability of a pedestrian
being synchronized with the footbridge motion. In a range of vibra- m1 j 1
Nc ¼ a  (29)
tion frequencies of 0:75–0:95 Hz, a value of a ¼ 2=3 is suggested. mo
As to b , reference is made to the experiments carried out by Arup
(Dallard et al. 2001), which show a strong variability, between 0:28 with a depending on the criterion adopted; these are summarized in
and 0:87, when the amplitude of oscillation is in the range of Table 1, together with the values it takes when the basic parameters

© ASCE C4015003-7 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


Table 1. Comparison of Stability Criteria (Data from Ricciardelli 2014)

Eq. a Suggested by authors Min Max


a,b a,b a
(23) f1 2:9–5:8 2.9 5:8a
8 p mo
k
(24) 4 15 6:9 30c
ab
(25) pffiffiffiffiffiffiffi f1 s fw 6:45–17:9d 3:1 19:5c
2p  31p 2
DLFL1 gc
(26) 4c1 29:1–36:2 29:1 57:2
(27) p2 m1 a21 7:8  105 ms a —e —e
2 DLF2 m g2
10:8 L1 o
(28) 2 36:2–40:2 15:0 80:5
f
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

1
32p 2
b a1 g
a
mo ¼ 70 kg.
b
f1 ¼ 0:5–1 Hz.
c
Phase lag ¼ p =2; a different phase lag increases this value.
d
fs ¼ 0:75–1:25 Hz.
e
This depends on the structure, as an example, for the Millennium Bridge
south span a ¼ 8:5, and for the Millennium Bridge center span a ¼ 11:3.

suggested by the authors are used, and with the minimum and maxi-
mum values coming from the broader combination of the values of
the basic parameters available in the literature.
First, it is observed that there is no consistency in the depend-
ency of a on the oscillation frequency. In Eqs. (24), (26), and (27), a
is independent of f1 ; in Eq. (23), it is proportional to f1 ; and in Eqs. Fig. 5. Variation of the motion-induced added mass, j p , and damping,
(25) and (28) it is proportional to the square of f1 . Then, it is noticed r p : (a) with the range of variation of the phase lag of the motion-
that there is a quite large scatter in the values of a. Some differences induced force; (b) as a function of the ratio of the oscillation frequency
can be explained based on the hypotheses used to derive the stability to the undisturbed walking frequency over a fixed floor (data from
criterion. As an example, the larger values of a in Eq. (26) with Ricciardelli et al. 2014)
respect to those in Eqs. (24) and (25) can be ascribed to the fact that
Eq. (26) was derived under the assumption of uniformly distributed
phases of the walkers, whereas Eqs. (24) and (25) were derived Assuming that unstable vibrations start from rest, the values of
under the assumption that the walkers are all in quadrature with the r p ðfc Þ and j p ðfc Þ appearing in Eqs. (30) and (32) can be obtained
footbridge motion. Some other differences are more difficult to from extrapolation of Eq. (21) to a zero amplitude of vibration, as
explain. follows:
An alternative criterion was presented in Ricciardelli (2014). This  3  2  3
leads to the finding that under the same hypotheses made earlier of si- f f f
j p ðf Þ ¼ 1:77 þ 7:77  9:55 þ 3:19
nusoidal vibration mode and uniformly distributed walkers, the fw fw fw
vibration frequency of the combined walker–footbridge system is    
1:11 f
r p ðf Þ ¼ 9:17  sin  exp 2:89 þ 0:56 (33)
fs f =fw fw
fc ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (30)
1 þ m  r p ðfc Þ
which are plotted in Fig. 5(a).
The results coming from Eq. (32) can then be checked with those
with m ¼ ðd BLmo Þ=ð2m1 Þ being the ratio of the total walker mass of other stability criteria given in Table 1. When the minimum value
to the footbridge mass. Eq. (30) must be solved iteratively, as the of the pedestrian-induced damping j p ¼ 0:4 in Fig. 5(a) is used,
added mass coefficient at the right-hand side of the equation needs then a ¼ 5. This result compares very well with Eq. (23), which
to be evaluated at the vibration frequency. A similar equation had confirms the quite good agreement of the results coming from the
already been used in Bocian et al. (2012). full-scale measurements on the Millennium Bridge (Dallard et al.
Then, the stability condition is 2001) and those coming from laboratory tests using a treadmill
(Ricciardelli et al. 2014), also in consideration of the totally differ-
j s þ m  j p ðfc Þ > 0 (31) ent experimental approach used for the two studies. If the average
mass of 74:4 kg and the average walking frequency of 0:857 Hz of
where j p is the damping coefficient associated with the walker the walkers participating in the tests is used, then the maximum neg-
action, referred to the walker mass (Ricciardelli et al. 2014). The ative j p of 0:40 is translated into cp ¼ 281 Ns=m, whereas the
critical number of walkers can be calculated as value of j p of 0:36 occurring for fs =fw ¼ 1 is translated into
cp ¼ 288 Ns=m. These differ only very little from the value of
m1  j 1 300 Ns=m derived from full-scale tests. Comparison can also be
Nc ¼ 2 (32)
 j p ðfc Þ  mo made with the value cp ¼ 140 Ns=m found numerically in
Macdonald (2008), from a model excluding the possibility of syn-
which can be put in the form of Eq. (29), with a ¼ 2= j p ðfc Þ. chronization. The ratio of the cp value found in Macdonald (2008)

© ASCE C4015003-8 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


and those found in Dallard et al. (2001) and in Ricciardelli et al. Maximum Acceleration
(2014) is in the order of 0.47–0.50, and this can be seen as a measure
The maximum acceleration due to a single, a group, or a stream of
of the percentage of walkers synchronizing with the footbridge and
walkers, either in the vertical or in the lateral direction, is usually
with the treadmill motion during the full-scale and laboratory tests,
given as
respectively.
Fig. 5(b) shows the range of variation, with varying amplitude of €y max ¼ Ne  €y max ð1Þ  Wðf1 Þ (34)
oscillation between 4:5 and 38 mm, of the phase lag of the motion-
induced force (data are derived from Ricciardelli et al. 2014). It is where Ne is the equivalent number of walkers on the bridge and
observed that this is roughly between p =4 and 3=8p for fs =fw ¼ accounts for the level of synchronization, €y max ð1Þ is the maxi-
0:8 and between p =8 and p =4 for fs =fw ¼ 0:8 to 1:1; this is the mum transient acceleration due to one walker resonant with the
range of frequency ratios of interest for synchronization. This result lowest harmonic that can be matched by the walking frequency,
can be used with the stability criteria brought to Eqs. (24) and (25); and Wðf1 Þ is a coefficient that reduces the response when the
when the value f ¼ p =8 is considered, the values of a in the third
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

fundamental frequency of the bridge f1 is away from the walk-


column of Table 1 increase to 39:1 for Eq. (24), and to the range ing frequency. Table 2 contains Ne and €y max ð1Þ, as shown in
16:9–46:8 for Eq. (25); when the value f ¼ p =4 is considered, the Eq. (5), for a number of codes and guidelines; for purpose of
values of a in the third column of Table 1 increase to 21:2 for Eq. comparison, all of the equations are written in the same form as
(24), and to the range 9:1–25:3 for Eq. (25). that of Eq. (34).
BS5400 (BSO 1978) and the Ontario Highway Bridge Design
Code (OHBDC 1983) are based on the model of Blanchard et al.
Design Procedures (1977) and consider only the case of one walker. For the single
walker, the design manual of the Highways Department of the gov-
Guidelines and codes of practice consider three aspects related to ernment of Hong Kong (Highways Dept. 2002) uses the same
the design of footbridges against pedestrian loading: (a) the range of approach as the BS5400 and OHBDC and additionally considers
natural frequencies to be avoided for footbridges; (b) the maximum the case of a stream of walkers; €y max ð1Þ in Highways Dept. (2002)
acceleration response due to the crossing of one, a group, or a is the maximum deflection of a simply supported beam with a uni-
stream of walkers; and (c) the critical number of walkers bringing form load corresponding to the weight of one walker distributed
lateral instability; additional specifications are in some cases pro- over the footbridge span, multiplied by a function of the walker ve-
vided for the loading deriving from other human activities, such has locity and of the footbridge span.
jogging, jumping, or vandalism; these are beyond the scope of this AASHTO (1997, 2009) and AISC (Murray et al. 1997) consider
article and are not discussed. only the case of a single walker and use the response model of Allen
and Murray (1993).
The European prestandard version, designated ENV, of Eurocode
Frequency Limitations 5 (EC5; CEN 1995) considers both the case of a group of three
synchronized walkers and that of a stream of walkers, but the for-
Pioneering standards prescribe that natural frequencies close to the
mer becomes relevant only for very small footbridges, with a deck
first and second walking harmonics, and to the jogging frequency, smaller than 37 m2 . Subsequently, the European Standard version,
should be avoided. BS5400 of the British Standards Institution (BSI designated EN, of EC5 (CEN 2005) modifies the equations, but
1978) prescribes that dynamic analyses should be carried out if the the case of a group of walkers remains dominant for very small
lateral fundamental frequency is less than 1:5 Hz. SIA160 of the footbridges, with a deck area in this case smaller than 22 m2 .
Societe suisse des Ingenieurs et des Architectes (SIA 1989) and For vertical and horizontal vibrations, ENV EC5 (CEN 1995),
the Model Code for Concrete Structures of the Federation EN EC5 (CEN 2005), and FIB (2005) consider the case of a
Internationale du Beton (FIB 2010) require avoiding vertical natural small group of three synchronized walkers and that of a stream of
frequencies in the range of 1:6–2:4 Hz (first walking harmonic), in walkers. Here again the case of a group of walkers becomes rele-
the range of 3:5–4:5 Hz (second walking harmonic), and in the vant for small footbridges, in particular when the number of
range of 2:4–3:5 Hz (running/jogging). If these requirements are walkers is lower than 13; this corresponds to the same condition
not fulfilled, the acceleration response of the structure is to be given in ENV EC5 when a density of 0:36 walkers=m2 is consid-
checked. ered. Moreover, for vertical vibrations, the case of stream of
SETRA (2006) defined levels of risk of vibration in both vertical walkers is considered.
and horizontal directions, depending on frequency. The maximum SETRA (2006) guidelines use a stationary harmonic load
level of vibration risk is associated with vertical vibration frequen- model for the evaluation of the maximum acceleration induced by
cies in the range of 1:7–2:1 Hz and lateral vibration frequencies in one pedestrian; the equivalent number of pedestrians is given as a
the range of 0:5–1:1 Hz; a medium level of vibration risk is associ- function of d .
ated with vertical vibration frequencies in the range of 1:0–1:7 Hz HIVOSS (Heinemeyer et al. 2009) adopts two approaches for
and 2:1–2:6 Hz and lateral vibration frequencies in the range of the evaluation of the maximum acceleration: the stationary har-
0:3–0:5 Hz and 1:1–1:3 Hz; a low level of vibration risk is associ- monic load model and the response spectral model. In the harmonic
ated with vertical vibration frequencies in the range of 2:5–5:0 Hz load model, Ne and €y max ð1Þ coincide with those given by SETRA
and lateral vibration frequencies in the range of 1:3–2:5 Hz. Other (2006), but Wðf1 Þ differs. The response spectral model was derived
vibration frequencies should not be considered a concern. from Monte Carlo simulations based on numerical time-step simu-
HIVOSS (Heinemeyer et al. 2009) defined the critical ranges of lations of various pedestrian streams on various bridges geometries
natural frequencies for footbridges excited by the first harmonic of (Butz 2006). In the equations in Table 2, kp is an empirical peak fac-
the walking load as 1:25–2:3 Hz for the vertical direction and tor that converts the RMS into a 95th-percentile peak acceleration,
0:5–1:2 Hz for the lateral direction, and by the second harmonic as C describes the maximum of the load spectrum, k1 ðfj Þ and k2 ðfj Þ are
2:5–4:6 Hz for the vertical direction. two polynomial functions depending on the frequency of the jth

© ASCE C4015003-9 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


Table 2. Maximum Acceleration from Different Codes

Standard Dir Ne €y max ð1Þ


BS5400 Vert 1 Blanchard et al. (1977)
OHBDC Vert 1 Blanchard et al. (1977)
Highway Dept. Vert 1 Blanchard et al. (1977)
Highway Dept. Vert N 5  q1  L4 p 2 v2 a
384EI 2L2
AASHTO Vert 1 Allen and Murray (1993)
AISC Vert 1 Allen and Murray (1993)
ENV EC5 Vert 3  0:027  L  B 165
½1  exp ð2p n j Þb
Mj
ENV EC5 Lat 3  0:027  L  B 40
½1  exp ð2p n j Þb
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Mj
EN EC5 Vert 0:23  0:6  L  B 200
Mj
EN EC5 Lat 0:18  0:6  L  B 50
Mj
FIB Lat/Vert 3 Rainer et al. (1988) or Grundmann et al. (1993)
FIB Vert 0:225  N Rainer et al. (1988) or Grundmann et al. (1993)
pffiffiffiffiffiffiffi
SETRA and HIVOSS Vert jffiN
10:8pffiffiffi d < 1 w=m2 2 280 1
 
1:85 N d  1 w=m2 p m1 2 j
pffiffiffiffiffiffiffi
SETRA and HIVOSS Lat jffiN
10:8pffiffiffi d < 1 w=m2 2 35 1
 
1:85 N d  1 w=m2 p m1 2 j
pffiffiffiffi c rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
HIVOSS Lat/Vert N k ðf Þ
kp  mC2 k1 ðfj Þ j j 2 j kf
j

Note: Ne ¼ 1 refers to the single walker, Ne ¼ 3 refers to the group of walkers, Ne 6¼ 1; 3 refers to a stream of walkers, Dir = the direction of vibration,
Vert = vertical, Lat = lateral).
a
v ¼ 3 m=s.
b
M is the total mass of the footbridge; n ¼ L=0:9 m.
c
In the HIVOSS model, the parameters contained in €y max ð1Þ are function of d , that is, N.

Fig. 6. Comparison of Wðf1 Þ for the vertical component; the gray and light-gray areas show the typical range of frequency for the first and second har-
monics for a typical walker

mode, and kf represents the variance of the loading of a single pe- In Figs. 6 and 7, the values of Wðf1 Þ for the vertical and lateral
destrian; these values are given in table form for the vertical and lat- components proposed by the different codes and guidelines are
eral components for crowd densities equal to: d < 0:5 walkers=m2 , compared. The gray and light-gray areas show the typical range of
d ¼ 1 walkers=m2 and d ¼ 1:5 walkers=m2 . In this model it is not frequency for the first and second harmonics for a typical walker.
possible to isolate the function Wðf1 Þ, because it is incorporated into The frequencies at which Wðf1 Þ take large values qualify the ranges
k1 ðfj Þ and k2 ðfj Þ. where resonance may occur, and Wðf1 Þ ¼ 1 means perfect resonant

© ASCE C4015003-10 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Comparison of Wðf1 Þ for the lateral component; the gray and light-gray areas show the typical range of frequency for the first and second har-
monics for a typical walker

Fig. 8. Comparison of Ne ðNÞ ( j a ¼ 0:01,


pffiffiffijffi b ¼ 0:02, j c ¼ 0:03); the light-gray area shows the range of Ne between the cases of full correlation
(Ne ¼ N) and of full uncorrelation (Ne ¼ N )

pffiffiffiffiffiffiffiffiffiffiffi
response. ENV EC5 (CEN 1995), EN EC5 (CEN 2005), and FIB Ne ¼ 10:8  j  N ; the second appliesptoffiffiffiffilarge densities and is inde-
(2005) give only one plot for Wðf1 Þ, which applies to the first and pendent of damping: Ne ðNÞ ¼ 1:85  N . The two equations pro-
second harmonics; therefore, the lower values appearing in Figs. 6 vide the same results for j ¼ 0:03. The scatter among the results pro-
and 7 in the second harmonic range are the effect of the ratio vided by the different equations is quite large; for example, for
between the DLFs, that is, DLF2 =DLF1 , being the resonant acceler- N ¼ 20, Ne is in the range of 2–8, and for N ¼ 100, Ne is in the range
ation, and €y max ð1Þ, computed using DLF1 in both cases. As a matter of 10–19.
of fact, the differences between the different curves in Figs. 6 and 7
are minor.
Fig. 8 compares the values of Ne ðNÞ proposed by different guide-
Lateral Stability
lines and codes of practice, indicating a clear disagreement among
the various proposals. The light-gray area shows the range of Ne The problem of lateral stability is also considered by some codes
ffiffiffiffi of full correlation (Ne ¼ N) and of full uncorrela-
between thepcases and guidelines. In particular, FIB (2005), SETRA (2006), HIVOSS
tion (Ne ¼ N ). FIB (2005) suggests to use Ne ¼ 3 for N  10, (Heinemeyer et al. 2009), and AASHTO (1997, 2009) give a value
Ne ¼ 0:225  N for a low-density stream, and Ne ¼ 0:27  0:46  N to for the critical number of walkers as defined in Dallard et al. (2001)
account for synchronization in higher-density streams; in the latter
pffiffiffiffi (described earlier in this article).
two cases the values of Ne lie below the lower limit given by N , up The U.K. National Annex to Eurocode 1 (BSI 2008) pre-
to N ¼ 70 and N ¼ 20, respectively. SETRA (2006) and HIVOSS scribes that all the modes with lateral natural frequencies below
(Heinemeyer et al. 2009) give two equations, depending on d (see 1:5 Hz have to compare their pedestrian mass damping parameter,
Table 2); the first applies to low densities and depends on damping: as follows:

© ASCE C4015003-11 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


j1  M results are found to be in quite good agreement with the Arup
D¼ (35)
N  mo formula.
The design procedures were compared in terms of equivalent
number of walkers on the bridge Ne , maximum transient accelera-
with the stability boundary expressed in terms of D given in the
tion due to one resonant walker €y max ð1Þ, and reduction coefficient
form of graph as a function of the frequency of the lateral mode.
accounting for nonresonant conditions Wðf1 Þ. For Ne and €y max ð1Þ, a
Considering the first sinusoidal mode shape and a linear interpolat-
quite large scatter was found, whereas only minor differences were
ing function of D in the range of 0:65 Hz  f1  1:70 Hz, Eq. (35)
leads to a critical number of walkers of found for Wðf1 Þ from one procedure to another.
In summary, this article has shown the clear need for a critical re-
vision of design procedures that, although inspired by the same
m1 j 1 2 m j
Nc ¼ a  ¼h i 1 1 (36) principles and applying common rules, bring rather different
mo 0:52  1:05 ðf1  0:65Þ
0:52 mo results. Additionally, the need for a validation of these procedures
through a systematic analysis of available full-scale data is evident.
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

Conclusions
References
This article provided a critical review of the available methods for
assessing footbridge serviceability against walking action. In partic- Allen, D. E., and Murray, T. M. (1993). “Design criterion for vibrations due
ular, single-walker models, multiple-walker models, interaction to walking.” AISC Eng. J., 30(4), 117–129.
models (inter-walker and walker-structure), and instability models AASHTO. (1997). Guide specifications for design of pedestrian bridges,
were addressed, together with current design procedures incorpo- Washington, DC.
rated into codes of practice and guidelines. Models originally pre- AASHTO. (2009). Guide specifications for design of pedestrian bridges,
sented in different forms were made homogeneous for purpose of Washington, DC.
comparison. The investigation was limited to the models allowing Blanchard, J., Davies, B. L., and Smith, J. W. (1977). “Design criteria and
practical use, although they are also of great interest, the more analysis for dynamic loading of footbridges.” Proc., Symposium on
Dynamic Behaviour of Bridges, Transport and Road Research
advanced approaches that are not yet ready for implementation into
Laboratory, Berkshire, England, 90–106.
design procedures were not dealt with. Some inconsistencies and
Bocian, M., Macdonald, J. H. G., and Burn, J. F. (2012). “Biomechanically
inaccuracies were pointed out. inspired modelling of pedestrian-induced forces on laterally oscillating
Two general conclusions can be drawn, the first being that in structures.” J. Sound Vib., 331(16), 3914–3929.
spite of the impressive amount of research developed in the last 15 Brownjohn, J. M. W., Fok, P., Roche, M., and Omenzetter, P. (2004a).
years, very few changes have appeared in design procedures. The “Long span steel pedestrian bridge at Singapore Changi Airport—Part
second is that when similar procedures are compared with one 2: Crowd loading tests and vibration mitigation measures.” Struct. Eng.,
another, they often show a quite large and unjustified scatter in the 82(16), 28–34.
results. Brownjohn, J. M. W., Pavic, A., and Omenzetter, P. (2004b). “A spectral
Single-walker models were compared in terms of the TRRC, density approach for modelling continuous vertical forces on pedestrian
and it was shown that the best-fitting numerical solutions is that of structures due to walking.” Can. J. Civ. Eng., 31(1), 65–77.
BSI (British Standards Institution). (1978). “5400—Part 2: Steel, concrete
Ricciardelli and Briatico (2011); this, however, is applicable only to
and composite bridges. Specification for loads.” BS5400, London.
the case of a sinusoidal mode shape. Other approaches seem to be BSI (British Standards Institution). (2008). U.K. national annex to
rather inaccurate. Eurocode 1, London.
For multiple walkers, stochastic frequency domain response Butz, C. (2006). “Beitrag zur Berechnung fußgänger induzierte Brückensch-
models exist for both vertical and lateral vibrations. Although pro- wingungen.” Ph.D. thesis, Shaker Verlag Aachen, Aachen, Germany (in
posed by different authors, these are in formal agreement with one German).
another. Time-domain models also exist, based on the results of nu- Butz, C., et al. (2007). “Advanced load models for synchronous pedestrian
merical simulations. excitation and optimised design guidelines for steel foot bridges
As to the possibility of accounting for interaction, it appears that (SYNPEX).” Project RFS-CR-03019, Final Rep., Office for Official
inter-walker interaction models seem to not yet be ready to describe Publications of the European Communities, Luxembourg.
the variation of GRFs and to model synchronization; the only infor- CEN (European Committee for Standardization). (1995). Eurocode 5,
Brussels, Belgium.
mation available for application seems to be the relationship
CEN (European Committee for Standardization). (2005). Eurocode 5,
between the walker density and the mean and RMS walking fre- Brussels, Belgium.
quency (Fig. 4) proposed by Butz et al. (2007).Walker–structure Charles, P., and Bui, V. (2005). “Transversal dynamic actions of pedes-
interaction is accounted for through the variation of the DLFs and trians. Synchronization.” Proc., Footbridge 2005, 2nd Int. Conf.,
the in-phase and in-quadrature components (with displacement) of Venice, Italy.
the GRFs with the floor acceleration and velocity; however, it Dallard, P., et al. (2001). “The London Millennium Footbridge.” Struct.
appears that only the latter lend themselves to stability analyses. Eng., 79(22), 17–21.
Stability criteria available in the literature were manipulated so Danbon, F., and Grillaud, G. (2005). “Dynamic behaviour of a steel foot-
as to express the critical number of walkers though one common bridge. Characterization and modelling of the dynamic loading induced
equation, and the values of the coefficient a appearing in such equa- by a moving crowd on the Solferino Footbridge in Paris.” Proc.,
Footbridge 2005, 2nd Int. Conf., Venice, Italy.
tions were compared, using the calibration parameters recom-
Eriksson, P. E. (1994). “Vibration of low-frequency floors: Dynamic forces
mended by the authors, as well as a wider range of these; a quite and response prediction.” Ph.D. thesis, Dept. of Structural Engineering,
large scatter of the results was found. Moreover, the data available Chalmers Univ. of Technology, Göteborg, Sweden.
for the amplitude and phase of the motion-induced force caused by Eriksson, P. E., and Ohlsson, S. V. (1988). “Dynamic footfall loading
a walker on a laterally vibrating floor allowed calibration on an orig- from groups of walking people.” Proc., Symposium Workshop on
inal criterion, in which stability is checked by looking at the sign of Serviceability of Buildings, Vol. 1, National Research Council,
the total damping of the combined footbridge–walker system; the Ottawa, Canada, 497.

© ASCE C4015003-12 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003


FIB (Federation Internationale du Beton). (2005). Guidelines for the design Piccardo, G., and Tubino, F. (2012). “Equivalent spectral model and maxi-
of footbridges, Lausanne, Switzerland. mum dynamic response for the serviceability analysis of footbridges.”
FIB (Federation Internationale du Beton). (2010). Model code for concrete Eng. Struct., 40, 445–456.
structures, Lausanne, Switzerland. Pimentel, R., and Fernandes, H. M. B. (2002). “A simplified formulation for
Fryba, L. (1973). Vibration of solids and structures under moving loads, vibration serviceability of footbridges.” Proc., Int. Conf. on the Design
Noordhoff, Groningen, Netherlands. and Dynamic Behaviour of Footbridges, Paris, 20–22.
Fujino, Y., Pacheco, B. M., Nakamura, S., and Warnitchai, P. (1993). Pizzimenti, A. D., and Ricciardelli, F. (2005). “Experimental evaluation of
“Synchronization of human walking observed during lateral vibration of the dynamic lateral loading of footbridges by walking pedestrians.”
a congested pedestrian bridge.” Earthquake Eng. Struct. Dyn., 10.1002 Proc., 6th European Conf. on Structural Dynamics—EURODYN 2005,
/eqe.4290220902, 741–758. Vol. 1, Millpress Science Publishers, Netherlands, 435–440.
Grundmann, H., Kreuzinger, H., and Schneider, M. (1993). “Dynamic cal- Racic, V., Pavic, A., and Brownjohn, J. M. W. (2009). “Experimental iden-
culations of footbridges.” Bauingenieur, 68(5), 215–225. tification and analytical modelling of human walking forces: Literature
Heinemeyer, C., et al. (2009). “Design of lightweight footbridges for human review.” J. Sound Vib., 326(1–2), 1–49.
induced vibrations: Background document in support to the implemen- Rainer, J. H., Pernica, G., and Allen, D. E. (1988). “Dynamic loading and
Downloaded from ascelibrary.org by University of Central Florida on 11/23/20. Copyright ASCE. For personal use only; all rights reserved.

tation, harmonization and further development of the Eurocodes.” JRC- response of footbridges.” Can. J. Civ. Eng., 51(1), 66–71.
ECCS Technical Rep., Office for Official Publications of the European Ricciardelli, F. (2005). “Lateral loading of footbridges by walkers.” Proc.,
Communities, Luxembourg. Footbridge 2005, 2nd Int. Conf., Venice, Italy.
Highways Dept. (2002). Structures design manual for highway and rail- Ricciardelli, F. (2014). “Lateral stability of footbridges subjected to crowd
ways, Highways Dept., Government of the Hong Kong Special loading.” Proc., 9th Int. Conf. on Structural Dynamics—EURODYN
Administrative Region, Kowloon Bay, Hong Kong. 2014, A. Cunha, E. Caetano, P. Ribeiro, G. Müller, eds., Porto, Portugal.
Ingólfsson, E. T., Georgakis, C. T., and Svendsen, M. N. (2008). “Vertical Ricciardelli, F., and Briatico, C. (2011). “Transient response of supported
footbridge vibrations: Details regarding and experimental validation of beams to moving forces with sinusoidal time variation.” J. Eng. Mech.,
the response spectrum methodology.” Proc., Footbridge 2008, 3rd Int. 10.1061/(ASCE)EM.1943-7889.0000241, 422–430.
Conf., Porto, Portugal. Ricciardelli, F., Mafrici, M., and Ingólfsson, E. T. (2014). “Lateral pedes-
Ingólfsson, E. T., Georgakis, C. T., Ricciardelli, F., and Jönsson, J. (2011). trian-induced vibrations of footbridges: Characteristics of walking
“Experimental identification of pedestrian-induced lateral forces on forces.” J. Bridge Eng., 10.1061/(ASCE)BE.1943-5592.0000597,
footbridges.” J. Sound Vib., 330(6), 1265–1284. 04014035.
Ingólfsson, E. T., Georgakis, C. T., and Jönsson, J. (2012). “Pedestrian- Ricciardelli, F., and Pansera, A. (2010). “An experimental investigation into
induced lateral vibrations of footbridges: A literature review.” Eng.
the interaction among walkers in groups and crowd.” Proc., 10th Int.
Struct., 45, 21–52.
Conf., Recent Advances in Structural Dynamics 2010, M. J. Brennan, I.
ISO. (2007). “Bases for design of structures—Serviceability of buildings
Kovacic, V. Lopes, K. Murphy, B. Petersson, S. Rizzi, and T. Yang, eds.,
and walkways against vibration.” ISO 10137, Geneva.
Southampton, U.K.
Macdonald, J. H. G. (2008). “Lateral excitation of bridges by balancing
Ricciardelli, F., and Pizzimenti, A. D. (2007). “Lateral walking-induced
pedestrians.” Proc. R. Soc. A, 465(2104), 1055–1073.
forces on footbridges.” J. Bridge Eng., 10.1061/(ASCE)1084
Matsumoto, Y., Nishioka, T., Shiojiri, H., and Matsuzaki, K. (1978).
-0702(2007)12:6(677), 677–688.
“Dynamic design of footbridges.” IABSE Proc., International
Roberts, T. M. (2005). “Lateral pedestrian excitation of footbridges.” J.
Association for Bridge and Structural Engineering, Zurich, Switzerland,
Bridge Eng., 10.1061/(ASCE)1084-0702(2005)10:1(107), 107–112.
1–15.
Rönnquist, A. (2005). “Pedestrian induced lateral vibrations of slender foot-
Matsumoto, Y., Sato, S., and Shiojiri, H. (1972). “Study on dynamic design
of pedestrian over-bridges in consideration of characteristics of pedes- bridges.” Ph.D. thesis, Norwegian Univ. of Science and Technology,
trians.” Proc. Jpn. Soc. Civ. Eng., 1972(205), 63–70. Trondheim, Norway.
Murray, T. M., Allen, D. E., and Ungar, E. E. (1997). Floor vibrations due SETRA (Service d’Études Techniques des Routes et Autoroutes). (2006).
to human activity: Steel design guide 11, AISC, Chicago. Footbridges: Assessment of vibrational behaviour of footbridges under
Nakamura, S. (2004). “Model for lateral excitation of footbridges by syn- pedestrian loading, Paris.
chronous walking.” J. Struct. Eng., 10.1061/(ASCE)0733 SIA (Societe suisse des Ingenieurs et des Architectes). (1989). SIA160
-9445(2004)130:1(32), 32–37. (Swiss Standard SIA160: Actions on structures), Zurich, Switzerland.
Nakamura, S., Kawasaki, T., Katsuura, H., and Yokoyama, K. (2008). Strogatz, S. H., Abrams, D. M., McRobie, A., Eckhardt, B., and Ott, E.
“Experimental studies on lateral forces induced by pedestrians.” J. (2005). “Theoretical mechanics: Crowd synchrony on the Millennium
Constr. Steel Res., 64(2), 247–252. Bridge.” Nature, 438(7064), 43–44.
Newland, D. E. (2003). “Pedestrian excitation of bridges—Recent Sun, L., and Yuan, X. (2008). “Study on pedestrian-induced vibration of foot-
results.” Proc., 10th Int. Congress on Sound and Vibration, bridge.” Proc., Footbridge 2008, 3rd Int. Conf., Porto, Portugal, 2–4.
International Institute of Acoustics and Vibration, Stockholm, Tilden, C. J. (1913). “Kinetic effects of crowds.” Proc., American Society of
Sweden, 533–547. Civil Engineers, Vol. 76, ASCE, New York, 2107–2126.
Oeding, D. (1963). “Verkehrsbelastung und Dimensionierung von Tredgold, T. (1825). A practical treatise on railroads and carriages,
Gehwegen und anderen Anlagen des Fußgängerverkehrs.” Ph.D. Taylor, London.
thesis, Bundesminister für Verkehr, Abt. Strassenbau, Bonn-Bad Venuti, F., and Bruno, L. (2009). “Crowd-structure interaction in lively
Godesberg, Germany (in German). footbridges under synchronous lateral excitation: A literature review.”
OHBDC (Ontario Highway Bridge Design Code). (1983). Highway Phys. Life Rev., 6(3), 176–206.
Engineering Division, Ministry of Transportation and Communication, Wheeler, J. E. (1982). “Prediction and control of pedestrian-induced vibra-
Ontario, Canada. tion in footbridges.” J. Struct. Div., 108(9), 2045–2065.
Ohlsson, S. V. (1982). Floor vibrations and human discomfort, Chalmers Young, P. (2001). “Improved floor vibration prediction methodologies.”
Univ. Technology, Division of Steel and Timber Structures, Göteborg, Proc., Arup Vibration Seminar, Vol. 4, Institution of Mechanical
Sweden. Engineers, London.
Okamoto, S., Abe, M., Fujino, Y., and Nakano, Y. (1992). “Characteristics Živanovic, S., Pavic, A., and Reynolds, P. (2005). “Vibration serviceability
of human walking on a laterally shaking floor.” Proc., Jpn. Soc. Civ. of footbridges under human-induced excitation: A literature review.” J.
Eng., 441(18), 177–184 (in Japanese). Sound Vib., 279(1–2), 1–74.
Piccardo, G., and Tubino, F. (2008). “Parametric resonance of flexible Živanovic, S., Pavic, A., and Reynolds, P. (2007). “Probability-based pre-
footbridges under crowd-induced lateral excitation.” J. Sound Vib., diction of multi-mode vibration response to walking excitation.” Eng.
311(1–2), 353–371. Struct., 29(6), 942–954.

© ASCE C4015003-13 J. Bridge Eng.

J. Bridge Eng., 2016, 21(8): C4015003

Das könnte Ihnen auch gefallen