Vous êtes sur la page 1sur 11

Biomechanically Inspired Modeling of Pedestrian-Induced

Vertical Self-Excited Forces


Mateusz Bocian1; John H. G. Macdonald2; and Jeremy F. Burn3

Abstract: Although many models of pedestrian dynamic loading have been proposed, possible bidirectional interactions between the
walker and the excited structure are generally ignored, particularly for vertical vibrations. This shortcoming has arisen from scarcity of data
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

on gait-adaptation strategies used in the presence of structural motion and, as a consequence, the absence of a credible fundamental pedestrian
model capable of capturing the underlying relations between the two dynamic systems. To address this inadequacy of current approaches,
a biomechanically inspired inverted-pendulum pedestrian model has been applied to the human-structure interaction problem. The behavior
of the model is studied when subjected to vertical motion of the supporting structure, in particular, in relation to potential self-excited forces
that can be generated. A mechanism has been identied by which the timing of pedestrian footsteps can be altered subtly, giving a net damp-
ing effect on the structure, without necessarily involving full synchronization. It has been found that depending on the ratio between the
bridge vibration frequency and pedestrian pacing frequency, walkers can effectively act as positive or negative dampers to the structural
motion, but it is expected that for a group of pedestrians with distributed parameters, their action is, on average, to add damping and mass.
DOI: 10.1061/(ASCE)BE.1943-5592.0000490. 2013 American Society of Civil Engineers.
CE Database subject headings: Bridges; Human factors; Pedestrians.
Author keywords: Bridges; Human-structure interaction; Biomechanics; Gait; Inverted-pendulum model; Vertical pedestrian forces.

Introduction pedestrian weight and the constant-amplitude rst harmonic only


[International Federation for Structural Concrete (b) 2005] and
Pedestrian excitation of footbridges and other lively structures others acknowledging the signicance of higher harmonics, with
remains of concern. Despite advances in the modeling of human- the amplitudes of the harmonics growing with walking frequency
induced forces in recent years, there are still signicant uncertain- (Inglfsson et al. 2008). A consequence of movement of a pedestrian
ties. In particular, pedestrian loading models generally consider the along the bridge in the context of modal response is usually accounted
dynamic loading to be an external force to the structure (Racic et al. for by application of a moving force to the structure (at the changing
2009) with the vast majority of existing modeling approaches based pedestrian location). To simulate an action of a crowd of walkers,
on an assumption of periodicity of gait. This simplication allows multiplication factors are sometimes applied to the loading function
denition of the time-varying vertical force as a sum of harmonic derived for a single pedestrian [International Organization for Stan-
components in the general form of (Bachmann and Ammann 1987) dardization (ISO) 2007], with other factors adopted to account for
possible frequency and phase synchronization [British Standards In-
P
j   stitution (BSI) 2003; b 2005]. In other cases, extensive simulations of
F G DGi sin 2ip fp t 2 wi (1) pedestrians modeled as inhomogeneous discrete particles crossing the
i1
bridge are conducted to estimate response of a structure to vertical
crowd loading [Technical Department for Transport, Roads and
where G 5 pedestrian weight; i 5 ith harmonic; DGi 5 load Bridges Engineering and Road Safety/French Association of Civil
amplitude of the ith harmonic; fp 5 pedestrian walking frequency; Engineering (SETRA/AFGC) 2006; Butz et al. 2008; Inglfsson et al.
t 5 time; wi 5 phase angle of the ith harmonic relative to the rst 2008]. Although most models treat the pedestrian loading as an external
harmonic (with w1 obviously equal to 0); and j 5 number of force (Dallard et al. 2001; Racic et al. 2009), the motion of the structure
harmonics considered. Different variations of this mathemat- can in fact affect the human behavior, thus modifying the dynamic
ical formulation have been suggested, some comprised of the interaction forces. The most public demonstration of this was the large-
amplitude lateral motion of the London Millennium Footbridge (LMF)
1 on its opening day in 2000. Similar behavior has since been identied
Ph.D. Student, Depts. of Civil Engineering and Mechanical Engi-
neering, Univ. of Bristol, Bristol BS8 1TR, U.K. (corresponding author). on many other bridges. Tests on the LMF itself led to the nding that
E-mail: Mateusz.Bocian@bristol.ac.uk pedestrians acted as negative dampers to the bridge motion (Dallard
2
Reader, Dept. of Civil Engineering, Univ. of Bristol, Bristol BS8 et al. 2001) because of some change in their gait. That is to say, the
1TR, U.K. motion of the bridge gave rise to a component of the pedestrian force in
3
Senior Lecturer, Depts. of Mechanical Engineering and Anatomy, phase with and proportional to the bridge velocity. Such self-excited
Univ. of Bristol, Bristol BS8 1TR, U.K.
forces cannot be modeled as predened external forces but require
Note. This manuscript was submitted on July 26, 2012; approved on
March 20, 2013; published online on March 22, 2013. Discussion period consideration of the bidirectional human-structure interaction.
open until May 1, 2014; separate discussions must be submitted for in- Vertical vibrations of footbridges and other such structures are
dividual papers. This paper is part of the Journal of Bridge Engineering, more commonly of concern because when the negative damping ef-
Vol. 18, No. 12, December 1, 2013. ASCE, ISSN 1084-0702/2013/12- fect just described is not critical, the vertical dynamic forces from
13361346/$25.00. pedestrians are much larger than the lateral forces. However, existing

1336 / JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013

J. Bridge Eng. 2013.18:1336-1346.


methods of predicting vertical vibration responses of bridges from walkers can alter the damping characteristics of bridges (Willford
pedestrian loading are still often inaccurate. This has been highlighted 
2002; Brownjohn et al. 2004; Zivanovi c et al. 2010). Similar ndings
recently in a review of several methods in comparison with full-scale have been made in measurements of the effects of standing and sitting

measurements by Zivanovi c et al. (2010). It was suggested that apart crowds, e.g., on grandstands, where it has been suggested that the
from other shortcomings of some of the methods, the effect of human- mechanical impedance of the human body is a potential source of this
structure interaction was effectively to increase the damping of the behavior [Institution of Structural Engineers (ISE) 2008]. That is to
structure. It was estimated that the effective damping of one bridge, say, the human body does not behave as a rigid mass, but its dynamic
with a natural frequency of 2.04 Hz, was increased by a factor of response, when subjected to vibrations, is more complex, with the
approximately 2.5 as a result of the presence of walking pedestrians. applied force having components other than just the one in phase with
Similar suggestions had been made previously by other authors based acceleration (Grifn 1990). Modeling of the mechanical impedance
on full-scale measurements (Willford 2002; Brownjohn et al. 2004). has been conducted most often with appropriately arranged and
However, such information remains sparse and uncertain, so there is calibrated vertically coupled spring-mass-damper (SMD) systems,
a need to investigate this potential effect further. To address this issue, and this approach is gaining popularity in modeling the effects of
a biomechanical biped pedestrian model is applied to a vertically pedestrian occupancy of structures. Kim et al. (2008) adopted the
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

oscillating structure to investigate the effect of vertical bridge motion two-degree-of-freedom SMD body model in ISO 5982:1981 (ISO
on human gait and hence on the interaction forces. A simple foot- 1981) for analysis of pedestrian-induced vertical bridge vibra-
placement control law satisfying kinematic constraints is adopted to tions. A single-degree-of-freedom SMD body model was used by
ensure forward progression. The results from the model in terms of the Brownjohn et al. (2004), Tavares da Silva and Pimentel (2011), and
equivalent added damping and mass to the bridge are compared with the Caprani et al. (2011). Alexander (2006) added an actuator to the two-
available data from full-scale bridges and experimental investigations. degree-of-freedom SMD crowd-structure system model, although
pedestrians were not treated there as discrete entities. However, there
is a clear shortcoming of these models for walking pedestriansthey
Modeling of Human-Structure Interaction
are not able to fully capture the inuence of the feedback from the
Biomechanically inspired modeling of pedestrian action in the context vertical motion of the walking surface on the kinematics and kinetics of
of engineering structures is a relatively novel direction of research. Its human gait. Specically, the changes in gait parameters resulting from
emergence is the result of serviceability problems experienced by two disturbance to the trajectory of the center of mass (COM) are neglected,
high-prole bridges: the Solferino Footbridge in Paris, opened in 1999 which can potentially modify the pedestrian loading. This originates in
(Danbon and Grillaud 2005), and the previously mentioned LMF the fact that these models ignore the bipedal nature of human locomotion.
(Dallard et al. 2001). When subjected to crowd loading, unforeseen Humans use two lower limbs for normal (i.e., nonpathologic and
large divergent amplitude lateral vibrations developed, causing the unconstrained) walkinga form of locomotion termed bipedalism.
authorities to make decisions of close and later retrot of these bridges The implications of this feature on human gait are profound, inuencing
with damping devices. Since then, many other bridges susceptible to qualities such as speed, acceleration, and maneuverability, endurance,
pedestrian-induced vibrations have been identied, and much effort economy, energy, and at last but not least, stability (Alexander 2003).
has been spent on studying the cause of this phenomenon. Most of the The last of these qualities is possibly the most essential to a functional
proposed explanations focus on synchronization of the pedestrians gait because while different legged animals have gaits that differ in
frequencies with the structural vibration frequency and adjustments of patterns of movement (e.g., walking, running, hopping, trotting, and
the pedestrians phases so as to increase bridge motion. This, perhaps galloping), their gaits share a common characteristicthey are rst and
the most intuitive, solution could not, however, explain instability of foremost stable. The reason for this is simplethe energy losses as-
two other bridges: the Changi Mezzanine Bridge at Singapore Airport sociated with lack of stability are too costly (e.g., recovery of potential
(Brownjohn et al. 2004) and the Clifton Suspension Bridge in energy and buildup of kinetic energy level after a fall). This also reveals
Bristol, United Kingdom (Macdonald 2008), where no evidence of why no credible model of human walking exists in the eld of bio-
synchronization was observed during periods of instability. The mechanics that ignores bipedalism. Therefore, in contrast with other
measurements from these bridges indicated that some other models of vertical human-structure interaction for walking pedestrians,
mechanism is required, conceivably similar to Barkers (2002) a bipedal model is considered here in which the bridge motion perturbs
model, in which the lateral force was obtained from the action of the gait and challenges its stability. In line with the statement that
a mass moving along the bridge in a straight line, supported by an simplicity promotes understanding validated in this context by McGeer
inclined leg alternately switching from right to left. Although this (1990) and Alexander (1992), the focus of this paper is identication of
simple model was not scientically justied, an important discovery the fundamental relations between a pedestrian modeled as a minimal
was madepedestrians can effectively input energy into a vibrating biped and a vertically oscillating bridge. The motion of the bridge is
bridge even if their walking frequencies are far from the frequency shown to modify the passive motion of the pedestrians COM and thus
of the vibrating structure. Direct empirical evidence of this was cause the emergence of self-excited forces on the structure.
obtained from tests during which pedestrians walked on a laterally
oscillating instrumented treadmill by Pizzimenti and Ricciardelli
Minimal Biped Model
(2005) and later by Inglfsson et al. (2011). A rigorous explanation
was provided by Macdonald (2009), who built a biomechanically The simplest possible stepping model of human gait is a rigid inverted
inspired biped inverted-pendulum model in which a stepping pendulum with instantaneous transfer from one leg to the other. This
strategy was adopted based on the requirement of maintaining model is especially valid for the single support phase of the gait,
balance (Hof 2008). Unlike in Barkers model, motion of the mass which, depending on step rate, comprises around 7095% of the
of pedestrians was included, therefore allowing for full pedestrian- whole gait cycle (Inman et al. 1981). In this period, the CoM moves
bridge dynamic interaction. Bocian et al. (2012) found that the along the arc determined by the length of the stance limb, here the
output of this model is in quite good agreement with the experi- pendulum leg. An advantage of this process is that as for a normal
mental investigations and measurements on full-scale bridges. pendulum, energy is conserved. There is transfer between gravita-
A new trend in modeling vertical pedestrian loading has origi- tional potential energy and kinetic energy, but once initiated, no
nated only recently, prompted by observations that the presence of mechanical work is required to sustain motion. Similar efciencies can

JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013 / 1337

J. Bridge Eng. 2013.18:1336-1346.


be made by exploiting passive dynamic characteristics of the swing a vertical impulse is applied at the moment of transition that imitates
leg moving to the new position at the beginning of the next step the heel strike of the leading leg and push-off of the trailing leg. For
under the action of gravity alone (Mochon and Mcmahon 1980) and steady-state walking on stationary ground, the magnitude of this
by treating the entire single-support phase as motion of a coupled impulse is calculated based on an assumption that the forward
two-pendulum system (Kuo 2007). component of velocity at the transition is conserved, whereas the
In accordance with the preceding statements, the mathematical vertical velocity component changes sense.
model adopted in this study is conceptually similar to the minimal The output of the inverted-pendulum model in terms of the
biped built by Alexander (1995). The mass of the pedestrian is con- vertical and longitudinal ground reaction forces (GRFs) for un-
centrated at a point on top of a massless rigid leg and is pivoted at the perturbed walking (z 5 0) is shown in Fig. 2 in comparison with
point where the leg meets the ground (Fig. 1). The equation of motion measured data averaged over 10 subjects. As can be seen, the
for the inverted-pendulum model depicted in Fig. 1 during the single- longitudinal GRF is captured very well by the model. For the
support phase is easily derived by applying dAlemberts principle and vertical GRF, the characteristic peaks in the measured forces near
taking moments about the support point or using the Lagrangian the beginning and end of each step are concentrated in the sim-
formalism plied model in impulses at the transition instants.
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.


u 21 g zcos u (2)
l Self-Excited Forces

where u 5 support-leg inclination angle; l 5 equivalent inverted- To consider the inuence of vertical bridge motion on the interaction
pendulum length; g 5 gravitational acceleration; z 5 vertical dis- forces, sinusoidal motion of the supporting surface was assumed for
placement of the bridge; and dots over the symbols represent all the following analysis, given by:
derivatives with respect to time. Because Eq. (2) is nonlinear, the
zt A sinvb t (4)
MathWorks MATLAB ode45 numerical ordinary differential equa-
tion solver, incorporating the Runge-Kutta algorithm, was used to
where A 5 amplitude of the bridge motion; and vb 5 angular
obtain time-history solutions. Variable time steps were allowed for the
solver for computation efciency, with the upper bound set to 0.01 s. frequency.
The effect of the bridge motion on the pedestrian modies the
The results were then resampled into time series with regular 0.001 s
sampling intervals. The vertical force on the bridge F is given by interaction force [see Eqs. (2) and (3)], which can potentially in-
clude components at the bridge frequency and which from the point
  of view of the bridge equation of motion (not addressed here per se)
F mp g z l
u cos u 2 lu_ sin u
2
(3)
can be considered as equivalent to added mass and added damping
of the structure. Although there is no fundamental difculty in
where mp 5 pedestrian mass. modeling the structure as well, this would involve additional
Clearly, the motion of the bridge z perturbs the motion of the parameters and would detract from the main goal with no real
pedestrian COM u, as given by Eq. (2), which thus modies the benet. From the point of view of the individual studied here, the
interaction force, in accordance with Eq. (3). This is a direct con- bridge is forced by the crowd or other external loading, and the
sequence of the bipedalism and has not been captured by any individuals inuence on bridge motion is minor. The pedestrian
previous model of vertical human-structure interaction.
effectively experiences imposed motion of the walking surface,
For simplicity, the transfer from one leg to the other is taken to be
which would be dominated by one or more natural frequencies of
instantaneous, and the dynamics of the swing leg are omitted. The
latter simplication is made based on the observation by Pandy
(2003) that the vertical ground reaction force during the single-
support phase is mainly a result of the stance-leg dynamics. To
account for the change in vertical velocity necessary at the transfer,

Fig. 2. Performance of the inverted-pendulum pedestrian model walking


at 5 km=h with l 5 1:045 m and fp 5 2 Hz in terms of vertical force
(continuous curve with discontinuous arrowed lines at the transition points
between steps indicating instantaneous impulses) and longitudinal force
Fig. 1. Inverted-pendulum pedestrian model subjected to vertical (dashed curve) in comparison with averaged data measured from subjects
ground motion (gray continuous lines show the conguration of legs at walking at the same speed on a mechanical treadmill (data from Masani
transfer to next step) et al. 2002)

1338 / JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013

J. Bridge Eng. 2013.18:1336-1346.


the structure. For simplicity, here the bridge motion is taken to be at characteristic values for males and females. Subsequently, the equiv-
one frequency [Eq. (4)]. The effect of the interaction is quantied in alent inverted-pendulum length in the sagittal plane (vertical plane con-
terms of the self-excited forces generated by the bridge motions taining the direction of progression) was found from a formula adopted
effect on the pedestrian. These forces could be applied to the from Hof et al. (2005): l 5 1:2 3 leg length, giving l 5 1:045 m. The
equation of motion of the bridge (with any parameter values), if established values of mp and l will be hereafter referred to as repre-
desired. The net effect of a large number of pedestrians applied in sentative pedestrian parameters.
this way can become signicant. This approach was used by The average frequency and velocity of people walking freely on
Bocian et al. (2012) to study human-structure interaction for lateral footbridges and shopping oors have been measured as 1:8 Hz and
vibrations, yielding good agreement with the results of simulations of 1:3 m=s and 2 Hz and 1:4 m=s, respectively (Pachi and Ji 2005).
a large number of individually modeled pedestrians fully interacting Therefore, in this paper, a pedestrian walking frequency of 1.86 Hz
with the equations of motion of two bridge modes. and mean walking velocity of 1:33 m=s were taken as typical for
The average components of the self-excited force at the bridge unrestricted walking obtained from the averages of the preceding
vibration frequency in phase with the bridge velocity Fvvseb and in values weighted toward results measured on footbridges.
phase with the bridge acceleration Favseb are given by (Macdonald
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

2009; Bocian et al. 2012)


Step-Transition Control Law
T0T
2 Apart from bipedalism, another important feature of human walking is
F vvseb F cosvb t dt (5) direction-dependent control of balance. This quality implies variation
T
T0 between strategies adopted for correcting postural instability in the
sagittal and frontal (i.e., vertical plane perpendicular to the sagittal
T0T plane) reference planes. Specically, whereas the dynamic stability in
F vaseb 22 F sinvb t dt (6) the frontal plane is achieved with active control executed by the
T central nervous system (CNS), little, if any, control is required in the
T0
sagittal plane (OConnor and Kuo 2009). This obviously applies to
normal walking conditions, i.e., unperturbed gait. However, it seems
where overbars represent averages; T0 5 arbitrary reference time reasonable for the same control mechanism to be valid in the sagittal
(after any initial transient behavior); and T 5 complete integration plane for walking on a vertically oscillating surface, at least in the
interval. To correctly capture the long-term average, T should cover presence of small vibrations. The reason for this is twofold. First, if the
integer numbers of bridge and pedestrian cycles with a representa- resulting gait were stable, savings of lower demand of processing
tive distribution of relative phases between them. The equivalent power from the CNS would correspond to a lower energetic cost of
added damping DC and added mass DM per pedestrian are then walking (energy optimization) (Alexander 1989). Second, the freed
found from CNS resources could be allocated to other tasks. Therefore, a step-
transition control law applicable to unperturbed walking is considered.
F vvseb
DC (7)
Avb
Constant-Touchdown-Angle Law
F vaseb Because the pendulum legs are rigid and of equal length, the legs
DM (8) must be arranged symmetrically at the transfer between one leg and
Av2b
the other (Alexander 1995). With regard to Fig. 1, this occurs when
Positive equivalent added damping is benecial because it
increases the amount of global damping. Conversely, negative equiv- 180 2 u uTD (9)
alent added damping is adverse. Furthermore, for free (rather than
imposed) vibrations of the structure, positive equivalent added mass can where uTD 5 touchdown angle, i.e., the initial value of u on the next
potentially decrease the vibration frequency, whereas negative equiv- step, taken to be constant for all steps. Note the 180 angle corresponds
alent added mass can cause the vibration frequency to increase. Con- to the constraint of the model because the pendulum is not allowed to
ventionally, the equivalent added mass may be expected to equal the fall below the level of the walking surface, which is considered to be
actual pedestrian mass mp [from the z term in Eq. (2)], but the full at. In the simulations, the touchdown angles uTD 5 70 and
interaction is more complex, giving rise to variable DC or DM uTD 5 72:6 were chosen such that (1) they fall within the range of
depending on the system parameters and the control law at the values measured for human walking (Lee and Farley 1998), and (2)
transition from one foot to the other. the results of simulations of the inverted-pendulum model on sta-
tionary ground yield reasonable gait parameters (e.g., walking speed
and walking frequency). However, to obtain a credible gait, the in-
Parameters
uence of the initial conditions also had to be examined. Therefore, the
To properly assess the dynamics of the model, a number of parameters performance of the model was rst mapped for a range of physio-
have to be dened that are physiologically plausible and at the same logically plausible equivalent pendulum lengths and different initial
time fulll the condition of dynamic similarity of the solution with forward velocities. For each combination of initial conditions, the
actual human gait. The average values of pedestrian mass and height simulations were run for 400 steps, conservatively discarding the rst
were established based on data published by the U.K. National Health 100 steps to avoid transients. The results for uTD 5 70 are presented
Service (NHS) in December 2009 from a representative sample of the in Fig. 3.
English population aged 16 and over (NHS 2009). These were, re- For a representative pedestrian and uTD 5 70, normal walking is
spectively, 76.9 kg and 1.683 m. The corresponding leg length of achieved when the model is given an initial velocity of approximately
0.871 m was obtained using the approximately linear relationship to 1:55 m=s. With this velocity, the model quickly converges on the
pedestrian height as stated by Pheasant (1982), averaging between steady-state motion, and all normal walking parameters are preserved.

JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013 / 1339

J. Bridge Eng. 2013.18:1336-1346.


The impulse contributes to the total self-excited forces in Eqs. (5)
and (6), which become summation terms at the relevant time instants.
Note that the magnitude of the impulse is independent of the adopted
integration interval. In practical implementation of a numerical al-
gorithm, one of the checks of its accuracy can be performed by nding
the average vertical force, including impulses, over the time of the
simulation. Such an average force so obtained should be equal to the
pedestrian weight.

Simulation Period
To determine the magnitude of the long-term average equivalent added
damping and mass, the simulations were run for at least 600 pedestrian
steps. Provision was made for any initial transients to decay; hence the
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

rst 300 pedestrian steps were conservatively discarded.


The gait derived from the model based on the constant-touchdown
angle was found, in most cases, to be nearly (but not normally exactly)
Fig. 3. Map of performance of the model in case of walking on stationary periodic. This means that small differences in single-step durations
ground in terms of average walking frequency (black curves, hertz) were present as a result of the imposed vertical bridge vibrations.
and average walking velocity (gray curves with boxed annotations, Therefore, to ensure that in the whole analysis period all possible
kilometers per hour) for a control law based on constant touchdown phase angles between the bridge and pedestrian were appropriately
angle with uTD 5 70 and for representative pedestrian parameters accounted for (i.e., in total, there being near-integer numbers of bridge
and pedestrian cycles), the phase angle between the bridge and the
pedestrian, identied from the timing of the instantaneous transfer
Note that the constant touchdown angle foot placement law does from one leg to the other with respect to the beginning of the nearest
not in fact allow for feedback control. Regardless of the amplitude preceding bridge vibration cycle [occurring when z 5 0 and z_ is
and frequency of vibration of the bridge, the foot is always placed at positive; see Eq. (4) and Fig. 1] in relation to the bridge vibration
a constant horizontal distance from the COM (and from the preceding period, and the angular velocity of the pedestrian were recorded at
foot, i.e., constant step length). Nonetheless, this foot-placement law the beginning of the 301st step. The end of the simulation time was
seems to be simple enough to be consistent with the notion of gait identied when the corresponding values at least 300 steps later were
being passively stable in the sagittal plane (Collins et al. 2005) without within 2% of these values.
any loss of generality in the case of small bridge vibrations. Mean-
while, the timing of each step is passively modied by the bridge
motion, the time of the foot transition being given when the solution of Results
Eq. (2) satises Eq. (9).
Truncated time history of the total vertical force (continuous curve
Initial Conditions on Each Step and Impulses with discontinuous arrowed lines indicating impulses applied at the
transition points between steps) in the presence of sinusoidal bridge
The velocity of the COM at the end of the step is also modied by the motion (dashed curve; A 5 3 mm, and the bridge vibration fre-
bridge motion in accordance with Eq. (2). It seems reasonable to quency fb 5 1:8 Hz) for a model based on representative pedestrian
assume that there is some correction of this velocity at the foot parameters and uTD 5 70 is presented in Fig. 4. The inuence of the
transition. In the absence of any further data on the actual transition bridge motion on the vertical force is manifested in misshaped
in the presence of perturbations, it is assumed that the initial velocity curvature of the force prole between the instantaneous impulses
on each step is simply reset to the value for walking on stationary applied at the transition points between steps (see Fig. 2), varying
ground. In practice, this task could be performed during the double- amplitude of these impulses, and altered timing of each consecutive
support phase of gait. Hence the initial conditions of u and u_ are the foot placement in relation to the bridge motion. Note that slow
same on each step. evolution of the force prole is caused by the close vicinity of the
To make a provision for the force necessary to redirect the COM bridge vibration frequency (1.8 Hz) and the pedestrian mean walking
to the next pendulum arc, an impulse is applied at the moment of frequency (1.86 Hz).
transition from one supporting leg to the other. The magnitude of The equivalent added damping and mass were found to depend
the vertical component of the impulse is given by on frequency and amplitude. Figs. 5(a and b) include results for
A 5 3 mm and A 5 6 mm and the adopted representative pedestrian
In 2mp y_ nf mp x_ 0n1 cot uTD (10) parameters, with touchdown angle uTD 5 70 and initial velocity of
1:55 m=s, for which the mean forward velocity on stationary ground
where In 5 vertical impulse at the end of the nth step; and y and is 1:33 m=s and the pedestrian frequency is 1.86 Hz. Also presented
x 5 vertical and horizontal displacements of the COM, respectively, in Figs. 5(a and b) are results for A 5 3 mm and the adopted
such that representative pedestrian parameters, but with touchdown angle
uTD 5 72:6 and initial velocity of 1:25 m=s, for which the mean
y_ nf lu_ n cos180 2 uTD
f
(11) forward velocity on stationary ground is approximately 1 m=s and
the pedestrian frequency is approximately 1.6 Hz, which is consis-
x_ 0n [ x_ 01 (12) tent with empirically established frequency-speed relationship for
walking obtained by Bertram and Ruina (2001). The higher forward
where the subscripts represent the step number; and the superscripts velocity is typical for unrestricted walking, whereas the lower ve-
0 and f represent the initial and nal values on that step, respectively. locity is relevant to a moderately dense crowd because the mean

1340 / JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013

J. Bridge Eng. 2013.18:1336-1346.


Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Truncated time history of total vertical force (continuous curve


with discontinuous arrowed lines indicating instantaneous impulses
applied at the transition points between steps) in presence of sinu-
soidal bridge motion (dashed curve; A 5 3 mm and fb 5 1:8 Hz) for
a model based on representative pedestrian parameters uTD 5 70 and
x_ 0n 5 1:55 m=s

pedestrian velocity decreases as crowd density increases [see Fig. 3 in


Bocian et al. (2012)].
According to the current model, the magnitude of the equivalent
added damping and mass is generally larger for lower vibration
amplitudes. It can be seen in Fig. 5 that the pedestrian negative
damping coefcient is negligible for fb  fp . If fb is close to but
lower than jfp , where j is a positive integer, DC is negative,
i.e., detrimental, but it becomes positive, i.e., benecial, when fb
is close to but higher than jfp . The equivalent added mass is close to
the actual pedestrian mass for all frequencies except close to jfp . This
relationship is sustained for the results from all considered values
of initial forward velocity and touchdown angle. Fig. 5. (a) Equivalent added damping; (b) equivalent added mass for
representative pedestrian, uTD 5 70; x_ 0n 5 1:55 m=s (giving mean
walking velocity of 1:33 m=s and fp 5 1:86 Hz for stationary ground);
Discussion A 5 3 mm (black curve and circles); and A 5 6 mm (gray curve and
circles); also presented are results for a representative pedestrian,
The model output has been compared with experimental data avail- uTD 5 72:6; x_ 0n 5 1:25 m=s (giving mean walking velocity of 1 m=s
able in the literature. It was necessary to estimate appropriate values of and fp 5 1:6 Hz for stationary ground); and A 5 3 mm (black curve and
the model parameters for these comparisons. The relationship be- triangles) [dashed horizontal line in (b) denotes pedestrian mass mp ]
tween the walking speed v, walking frequency fp , and step length
d was used: v 5 fp d. In the case of known fp (where reported in the
literature) but unknown v and d, the walking velocity was determined
from the average value identied from 12 subjects by Bertram and tests when the PB vibrated with a displacement amplitude of 2.8 mm
Ruina (2001) for a given fp . Step length was then calculated from the (determined based on peak acceleration of 0:44 m=s2 as averaged
preceding relationship. The equivalent inverted-pendulum length over values of peak acceleration per 150 s obtained from three tests).
l and pedestrian mass mp were taken as equal to the representative The average pedestrian parameters were established as 1.87 Hz for
pedestrian parameters adopted in this study. Given the symmetry at the the average walking frequency and 0.74 m for the average step
transfer from one leg to the other (Fig. 1), the touchdown angle was length. It was identied that the back-calculated damping ratio of
found from the simple trigonometric relation cos uTD 5 d=2l. This the relevant bridge mode increased from 0.26 to 0.67% during
left only one parameter to be dened, i.e., initial velocity x_ 0n , which was tests, suggesting additional benecial damping from the presence
chosen such that fp from the output of the simulations coincided with of walkers. The estimated average additional damping coefcient
the value observed on the real structure. In the case of known fp and d, per pedestrian in this case was between 1,125 and 1,395 Ns=m. The
the remaining problem was optimization of the initial velocity x_ 0n in simulations were run for the same vibratory conditions as measured
compliance with the known v. during the tests (A 5 2:8 mm and fb 5 2 Hz) and for corresponding
To date, the most detailed data on the dynamic behavior of average pedestrian parameters (uTD 5 69:3 and x_ 0n 5 1:6 m=s). The
a bridge in the vertical direction under the action of pedestrians come estimated average equivalent added damping per pedestrian was

from the Podgorica Bridge (PB) (Zivanovi c 2012). The vibration 504 Ns=m, which is approximately 2.232.77 times smaller than the
frequency of the fundamental vertical mode of the bridge was back-calculated values obtained from measurements on the PB. The
measured at 2.04 Hz and decreased to 2 Hz during the crowd loading average equivalent added mass per pedestrian was 33 kg, which is

JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013 / 1341

J. Bridge Eng. 2013.18:1336-1346.


consistent with the measured decrease in the vibration frequency of means that for a crowd of walkers with distributed parameters (e.g.,
the considered mode. walking frequency, mass, and leg length), the net effect on the
No evidence of benecial pedestrian damping was found by structure is likely to add damping and increase mass.

Zivanovi c et al. (2010) during tests on the Reykjavik City Footbridge
(RCF) for the considered second vertical mode (2.33 Hz). The
Phase Relationships
displacement amplitude in this case was approximately 0.33 mm
(assuming RMS acceleration of 0:05 m=s2 in the considered mode). Another interesting way to look at the results is to analyze phase
The average pedestrian parameters were established as 1.95 Hz for angles between the timing of the instantaneous transfer from one leg
average walking frequency and 1:4 m=s for average walking speed. to the other and the beginning of the nearest preceding bridge vi-
The simulations were run for the same vibratory conditions as bration cycle [occurring when z 5 0 and z_ is positive; see Eq. (4) and
measured during the tests (A 5 0:33 mm and fb 5 2:33 Hz) and for Fig. 1] dened here in relation to the bridge vibration period. At this
corresponding average pedestrian parameters (uTD 5 69:8 and moment, energy is supplied to or extracted from the structure
x_ 0n 5 1:61 m=s). The estimated average equivalent added damping through the impulse, ensuring that the motion of the CoM is
per pedestrian was 145 Ns=m, and the average equivalent added redirected to the new pendulum trajectory. The impulses give the
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

mass per pedestrian was 56 kg. Lack of evidence of increased dominant component of the average self-excited forces in the
damping on the RCF was explained by Zivanovi  c et al. (2010) to be integrals in Eqs. (5) and (6). An example of the case where fb is
caused by two factors: insufcient acceleration levels and time of slightly lower than fp is shown in Fig. 6(a), whereas an example of
exposure as a result of the short lengths of the spans on which the case where fb is slightly higher than fp is shown in Fig. 6(b). In
measurements were taken. Another possible explanation for the these plots, densication of the phase angles occurs in the vicinity of
discrepancy between the measured and expected added damping p [Fig. 6(a)] and in the vicinity of 0 or 2p [Fig. 6(b)]. For com-
might be its dependence on the walking frequency that is captured by parison, an example where fb is not close to fp or its positive-integer
the current model (Fig. 5). However, this issue requires further multiples is shown in Fig. 6(c). The phase angles in this case are
investigation based on empirical data. evenly distributed throughout the whole range.
The presented model is also reasonably consistent with the The plots presented in Fig. 6 reveal the reason for the different
results of laboratory investigations on the behavior of a slab sub- effective added damping from the pedestrians in different bridge
jected to the action of 2, 4, 6, and 10 walking people reported by frequency ranges. For bridge frequencies far from integer multiples

Zivanovi c et al. (2009). During testing, the slab with a mass of of the (unperturbed) pedestrian pacing frequency [e.g., Fig. 6(c)], the
approximately 15,000 kg vibrated with a frequency of approxi- step periods are affected little by the bridge motion, and there is
mately 4.45 Hz and a displacement amplitude of approximately a steadily progressing phase difference between the bridge and
2 mm (based on acceleration amplitudes equal to peak acceleration pedestrian, giving an even distribution over all phase angles. Hence,
levels of 1:85, 1:6, 1:46, and 1:19 m=s2 achieved during testing). It on some steps, the work done by the pedestrian on the bridge will be
was identied for all tests that pedestrians acted as a source of positive, but equally on others steps, it will be negative, so on
additional benecial damping to the structure, increasing its damping average overall there is little work done on the bridge, and the ef-
ratio approximately proportionally to the number of walkers from fective added damping coefcient is close to zero [Fig. 5(a);
0.72% in the case of an empty slab to 2.86% in the case of 10 fb 5 0:5 Hz).
pedestrians present on the slab. The simulations were run for the For bridge frequencies slightly above the (unperturbed) pedes-
average vibratory conditions measured during testing (A 5 2 mm and trian pacing frequency [e.g., Fig. 6(b)], for phase angles close to zero
fb 5 4:45 Hz). Given the lack of details regarding the behavior of the (i.e., bridge velocity upward at the beginning of the pedestrian step),
pedestrians and their parameters, model parameters had to be chosen. the bridge is generally accelerating downward for the rst half of the
C. T. Georgakis (personal communication, 2012) found that during step (u , 90; see Fig. 1) and upward for the second half of the step
similar tests on a footbridge set in a laboratory environment, the (u . 90). Both these bridge accelerations tend to accelerate the
preferred pacing rate of pedestrians decreased to 1.38 Hz, and he pedestrian motion [see Eq. (2)]. Hence the pedestrian arrives at
attributed this behavior to physical and psychological constraints the end of the step earlier than for unperturbed walking. Therefore,
associated with a narrow deck and a lack of handrails. It was assumed the relative phase angle between the bridge and pedestrian pro-
herein that pedestrians acted in the same manner during tests con- gresses more slowly than otherwise expected. Conversely, for phase

ducted by Zivanovi c et al. (2009) and that the average pedestrian angles close to p at the beginning of the step, the pedestrian motion is
walking frequency matched that measured by Georgakis. Based on decelerated by the bridge acceleration, and the relative phase angle
these assumptions and using data from Bertram and Ruina (2001), the progresses more rapidly. Hence there is a concentration of phase
corresponding average pedestrian parameters were established angles close to zero. The principal dynamic load from the pedestrian
as uTD 5 74 and x_ 0n 5 1:065 m=s. The estimated average equivalent is the impulse at the foot transition. For the favored phase re-
added damping per pedestrian was 283 Ns=m, and the average lationship, i.e., close to zero, the impulse (downward on the bridge)
equivalent added mass per pedestrian was 61 kg. More detailed occurs when the bridge velocity is upward; hence it has a damping
comparisons would require further speculations as to the behavior effect on the bridge motion, so the net DC is positive [Fig. 5(a);
of pedestrians during testing and therefore are omitted. fb 5 1:9 Hz; for x_ 0n 5 1:55 m=s, fp  1:86 Hz]. Similarly, for bridge
Unfortunately, the authors are not aware of any experimental data frequencies slightly below the (unperturbed) pedestrian pacing
in the frequency ranges where the model predicts equivalent nega- frequency [e.g., Fig. 6(a)], the bridge acceleration modies the
tive damping for further validation of the model. pedestrian step periods so as to concentrate the relative phases near
Note that the peak damping coefcients (Fig. 5) are much larger p, giving net negative effective damping to the bridge [Fig. 5(a);
in magnitude than the value of 2300 Ns=m proposed by Arup for fb 5 1:8 Hz; for x_ 0n 5 1:55 m=s, fp  1:86 Hz].
lateral bridge vibrations from the behavior of the LMF (Dallard et al. The effect of the phase angle between the beginning of step and
2001). Hence the estimated effective damping for vertical vibrations the beginning of bridge vibration cycle [occurring when z 5 0 and z_
could have a major inuence on bridge behavior. Also note that, on is positive; see Eq. (4) and Fig. 1], in relation to the bridge vibration
average, the pedestrians potential to add damping to the bridge is period, on the pedestrian motion is presented in more detail in Fig. 7.
stronger than the potential to effectively subtract damping. This All results were obtained for a model with representative pedestrian

1342 / JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013

J. Bridge Eng. 2013.18:1336-1346.


Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Effect of vertical bridge vibration on pedestrian motion


for a model with representative pedestrian parameters uTD 5 70;
x_ 0n 5 1:55 m=s (giving mean walking velocity of 1:33 m=s and
fp 5 1:86 Hz for stationary ground); A 5 3 mm; and fb 5 1:8 Hz in case
of the initial phase angle between beginning of step and beginning of
bridge vibration cycle of (a) p; (b) 0 [dashed lines denote transition
instants between steps, and uz presented in bottom subplots is the
component of the angular acceleration that depends on bridge accel-
eration (presented in the middle subplots) resolved by a cosine of the leg
inclination angle u (presented in the top subplots) in the equation of
motion uz 5 21=lz cos u; see Eq. (2)]

parameters uTD 5 70(u 2 huTD ; 180+ 2 uTD i), x_ 0n 5 1:55 m=s (giv-
ing a mean velocity of 1:33 m=s and fp 5 1:86 Hz for stationary
ground), A 5 3 mm, and fb 5 1:8 Hz, corresponding to Fig. 6(a). uz
Fig. 6. Phase angle between timing of instantaneous transfer from one presented in the bottom subplots is the component of the angular
leg to the other and beginning of the nearest preceding bridge vibration acceleration of the COM that depends on the bridge acceleration
cycle [occurring when z 5 0 and z_ is positive; see Eq. (4) and Fig. 1] in (presented in the middle subplots), resolved by the cosine of the leg
relation to bridge vibration period (after conservatively discarding rst inclination angle u (presented in the top subplots) in the equation of
300 steps, allowing the transients to decay) for a model with repre- motion [uz 5 21=lz cos u; see Eq. (2)]. Therefore,
uz modies the
sentative pedestrian parameters uTD 5 70; x_ 0n 5 1:55 m=s (giving mean dominant component of the angular acceleration, which depends on
walking velocity of 1:33 m=s and fp 5 1:86 Hz for stationary ground); gravity [
ug 5 21=lg cos u; see Eq. (2)], which denes the behavior
A 5 3 mm and (a) fb 5 1:8 Hz; (b) fb 5 1:9 Hz; (c) fb 5 0:5 Hz of the model in the case of stationary ground. Fig. 7(a) reveals that for
a step starting at a bridge phase angle of p [hence acceleration
z 5 2A sinvb ts 1 p, where ts is the time from the beginning of the
step, as in the middle subplot of Fig. 7(a); see Eq. (4)], the vertical

JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013 / 1343

J. Bridge Eng. 2013.18:1336-1346.


vibration of the bridge inuences movement of the pedestrian by
slowing down the progression of the COM because uz is in this case
generally negative. Conversely, Fig. 7(b) reveals that for a step starting
simultaneously with the bridge vibration cycle [phase angle 5 0,
hence acceleration z 5 2A sinvb ts , as in the middle subplot in
Fig. 7(b); see Eq. (4)], the vertical vibration of the bridge inuences
movement of the pedestrian by speeding up progression of the COM
because uz is in this case typically positive. Note that the step duration
expressed as a proportion of the bridge cycle duration, marked in
Fig. 7 by dashed lines that also denote the instant of the impulse
applied at the step transitions, is longer for a phase angle of p [Fig. 7
(a)] than for a phase angle of 0 [Fig. 7(b)]. Hence the phase evolves
more slowly when close to p than when close to 0, giving the bias
toward the phases close to p in Fig. 6(a). Also note that the impulse
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

occurring at the step-transition instants is applied when the bridge is


traveling downward [having negative velocity; see Eq. (4)] for the
case when phase angle is p, and hence the pedestrian does positive
work on the bridge and amplies its motion [Fig. 7(a)]. For a phase
angle of 0, the impulse is applied when the bridge is traveling upward
[having positive velocity; see Eq. (4)], and hence the pedestrian does
negative work on the bridge and damps its motion [Fig. 7(b)].
These results and their explanation demonstrate that the effective
added damping from pedestrians can be positive or negative
depending on the relationship between the bridge and pedestrian
frequencies. This occurs as a result of the bridge motion subtly al-
tering the timing of the footfall impulses to bias the net effect.
For some combinations of vibration and walking frequencies,
when the walking frequency is very close to integer multiples of the
vibration frequency, the model can yield pedestrian motion syn-
chronized with the bridge, although this is not a necessary condition
for the pedestrian to have a (positive or negative) damping effect on
the bridge. The cases where synchronization occurs for fp 5 1:86 Hz
on stationary ground and fb is close to this value are shown in Fig. 8,
which is a zoomed view of part of Fig. 5. The tendency to synchronize
is greater for larger amplitudes of vibration. For example, in Fig. 8, for
a vibration amplitude of 3 mm, synchronization occurs within the Fig. 8. Zoom of Fig. 5 for bridge frequencies close to pedestrian
bridge vibration frequency range of 1.851.88 Hz, whereas doubling frequency showing cases when synchronization occurs (open circles
the vibration amplitude broadens this range to 1.831.9 Hz. A record mark cases where no synchronization occurs, whereas lled circles mark
of occurrences of synchronization of pedestrians to the vertical ground cases where synchronization is evident): (a) equivalent added damping;
motion reported in the past has been compiled by Willford (2002) and (b) equivalent added mass based on simulations for representative
includes cases described in Bachmann and Ammann (1987) and pedestrian; uTD 5 70, x_ 0n 5 1:55 m=s (giving mean walking velocity of
Grundmann et al. (1993). The current model is able to capture this 1:33 m=s and fp 5 1:86 Hz for stationary ground); A 5 3 mm (black
behavior and predicts an increasing tendency toward synchronization curve); and A 5 6 mm (gray curve) [dashed horizontal line in (b) denotes
with larger amplitudes of vibration. pedestrian mass mp ]

Model Limitations and Other Effects


restoring the velocity at the moment of transition from one to the next
It must be kept in mind that the results presented herein are the step to the characteristic value for unperturbed walking. However, in
average values expected over a long time. For short periods of ex- real gait, this task, performed mostly during the double-support phase,
posure (e.g., for individual pedestrians passing over short bridges), is unlikely to be executed perfectly. Therefore, although this model
the bridge response might vary considerably from one pedestrian indicates that the pedestrian-equivalent added damping and mass
crossing to another. This is not to say that the underlying effects of potentially can be either positive or negative in different conditions,
the pedestrian are different but that force from the pedestrian can the quantitative results should be treated with care. The results are in
effectively excite or dampen the bridge over short periods. However, fact quite sensitive to the control law used for the transition from one
it is expected that for a group or crowd of walkers with the same foot to the other, so this aspect of human behavior needs further
parameters, the results would converge to the average values pre- consideration.
sented herein. Note that a similar line of argument was adopted by The results are sensitive to the walking frequency, which is de-
Dallard et al. (2001) in the derivation of the stability criterion based termined by the initial velocity, touchdown angle, and equivalent
on the average negative damping coefcient per pedestrian from pendulum length. For the current model, different walking fre-
results of tests on the LMF laterally excited by a crowd of walkers. quencies produce similar patterns of the equivalent added damping
The rigid inverted-pendulum pedestrian model subjected to ver- and mass to those presented in Fig. 5 but shifted to the corresponding
tical excitation of the supporting surface has been considered with walking frequency and its harmonics.
a simple control law based on an assumption that the pedestrian will The results presented herein are typical for normal, i.e., un-
try to compensate for the COM trajectory disturbance by fully constrained, walking and walking in moderately dense crowds. Lower

1344 / JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013

J. Bridge Eng. 2013.18:1336-1346.


average velocities and walking frequencies would be more relevant to References
loading from a dense crowd because the mean pedestrian velocity
(hence walking frequency) decreases considerably as crowd density Alexander, N. A. (2006). Theoretical treatment of crowd-structure in-
increases. To calibrate the model for such conditions, the initial ve- teraction dynamics. Proc., ICE Struct. Build., 159(6), 329338.
locity x_ 0n and the touchdown angle uTD would need to be changed Alexander, R. M. (1989). Optimization and gaits in the locomotion of
because they are known to depend on walking velocity. Furthermore, vertebrates. Physiol. Rev., 69(4), 11991227.
for low speeds, the mechanics of walking change in that the duration of Alexander, R. M. (1992). A model of bipedal locomotion on compliant
the double-support phase of gait increases, and hence inclusion of this legs. Phil. Trans. R. Soc. Lond. B, 338(1284), 189198.
feature might be necessary. Alexander, R. M. (1995). Simple models of human movement. Appl.
Other signicant factors that could improve the model are the Mech. Rev., 48(8), 461469.
Alexander, R. M. (2003). Principles of animal locomotion, Princeton
inherent body damping, leg compliance, and motion of the swing
University Press, Woodstock, U.K.
leg. It is well known that the human body exhibits certain resonant Bachmann, H., and Ammann, W. (1987). Vibrations in structures induced
characteristics and internal damping that depend on posture. For by man and machines, 3rd Ed., International Association for Bridge and
example, during whole-body vertical vibrations of standing sub- Structural Engineering (IABSE), Zurich, Switzerland.
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

jects, support conditions, such as knee postures and whether one or Barker, C. (2002). Some observations on the nature of the mechanism that
two feet are in contact with the ground, have been found to inuence drives the self-excited lateral response of footbridges. Proc., Foot-
the resonant frequency of the apparent mass. The main resonance of bridge 2002: 1st Int. Conf., Paris.
a standing human exposed to 1 m=s2 (RMS) vibrations occurs at Bertram, J. A. E., and Ruina, A. (2001). Multiple walking speed-frequency
5.5 Hz and decreases to 3.75 Hz in the case of standing on one leg and relations are predicted by constraint optimization. J. Theor. Biol.,
to 2.75 Hz for a posture in which both legs are bent and in contact 209(4), 445453.
with the ground (Matsumoto and Grifn 1998). Such effects could Bocian, M., Macdonald, J. H. G., and Burn, J. F. (2012). Biomechanically
be represented by changes in the stiffness and damping of the SMD inspired modelling of pedestrian-induced forces on laterally oscillating
human models (Matsumoto and Grifn 2003). Hence, to fully structures. J. Sound Vibrat., 331(16), 39143929.
capture the damping effect of walking pedestrians on the supporting British Standards Institution (BSI). (2003). UK national annex to Eurocode
structure, it may be necessary to include such features along with the 1: Actions on structures. 2: Trafc loads on bridges, London.
Brownjohn, J. M. W., Fok, P., Roche, M., and Omenzetter, P. (2004). Long
bipedal behavior considered here. This could be achieved, for ex-
span steel pedestrian bridge at Singapore Changi Airport. 2: Crowd loading
ample, by adding a spring and a dashpot damper to the current model.
tests and vibration mitigation measures. Struct. Eng., 82(16), 2834.
However, importantly, the current model demonstrates that sig- Butz, C., et al. (2008). Advanced load models for synchronous pedestrian
nicant effective damping of the structure can arise as a result of the excitation and optimised design guidelines for steel footbridges. RFSR-
simple bipedal mechanics of walking, without considering the internal CT-2003-00019, European Commission, Brussels, Belgium.
dynamic characteristics of the human body. Therefore, any model of Caprani, C. C., Keogh, J., Archbold, P., and Fanning, P. (2011). Char-
the human-structure interaction from walking pedestrians should take acteristic vertical response of a footbridge due to crowd loading. Proc.,
into account the stepping behavior and its modication by bridge Eurodyn 2011: 8th Int. Conf. on Structural Dynamics, 978985.
motion. Furthermore, currently used models of human dynamic loading Collins, S., Ruina, A., Tedrake, R., and Wisse, M. (2005). Efcient bipedal
neglect this effect, which could signicantly affect the resulting struc- robots based on passive-dynamic walkers. Science, 307(5712), 1082
tural response amplitudes. 1085.
Dallard, P., et al. (2001). The London Millennium Footbridge. Struct.
Conclusions Eng., 79(22), 1733.
Danbon, F., and Grillaud, G. (2005). Dynamic behaviour of a steel foot-
bridge: Characterisation and modelling of the dynamic loading induced
This paper presents results of a study on interactions between a walking
by a moving crowd on the Solferino Footbridge in Paris. Proc.,
pedestrian and a vertically oscillating structure. It has been shown that
Footbridge 2008: 2nd Int. Conf., Porto, Portugal.
the pedestrian has a potential to act as a source of either positive or Grifn, M. J. (1990). Handbook of human vibration, Academic Press,
negative equivalent added damping and mass to the structure depending London.
on the system parameters. However, the net expected effect from a group Grundmann, H., Kreuzinger, H., and Schneider, M. (1993). Schwin-
of pedestrians with distributed parameters is likely to be an increase in gungsuntersuchungen fr Fugngerbrcken [Vibration tests of pe-
effective damping and mass. A possible subtle entrainment mechanism destrian bridges]. Der Bauingenieur, 68, 215225 (in German).
has been identied that, depending on the pedestrian-to-bridge fre- Hof, A. L. (2008). The extrapolated center of mass concept suggests
quency ratio, could be responsible for providing negative or positive a simple control of balance in walking. Hum. Mov. Sci., 27(1), 112125.
damping without necessarily involving full synchronization of the pe- Hof, A. L., Gazendam, M. G. J., and Sinke, W. E. (2005). The condition
destrian to the bridge frequency. The key nding is that signicant ef- for dynamic stability. J. Biomech., 38(1), 18.
fective damping and variations in effective mass of the structure resulting Inglfsson, E. T., Georgakis, C. T., Ricciardelli, F., and Jnsson, J. (2011).
from pedestrians can arise from their stepping behavior, which has not Experimental identication of pedestrian-induced lateral forces on
been considered previously. This is apart from any effects of resonances footbridges. J. Sound Vibrat. 330(6), 12651284.
Inglfsson, E. T., Georgakis, C. T., and Svendsen, M. N. (2008). Vertical
or internal damping of the human body. However, improvements to the
footbridge vibrations: Details regarding and experimental validation of the
model, based on experimental observations of human gait, are needed to
response spectrum methodology. Proc., Footbridge 2008: 3rd Int. Conf.,
build a more reliable fundamental pedestrian model.
Porto, Portugal.
Inman, V. T., Ralston, H. J., and Todd, F. (1981). Human walking, Williams
Acknowledgments & Wilkins, Baltimore.
Institution of Structural Engineers (ISE). (2008). Dynamic performance
This work was supported by an Engineering and Physical Sciences requirements for permanent grandstands subject to crowd action:
Research Council (EPSRC) Doctoral Training Account studentship Recommendations for management, design and assessment, London.
for Mateusz Bocian and an EPSRC Advanced Research Fellowship International Federation for Structural Concrete (b). (2005). Guidelines
for John H. G. Macdonald. for the design of footbridges. b Bull. 32, Lausanne, Switzerland.

JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013 / 1345

J. Bridge Eng. 2013.18:1336-1346.


International Organization for Standardization (ISO). (1981). Vibration OConnor, S. M., and Kuo, A. D. (2009). Direction-dependent control of
and shock: Mechanical driving point impedance of the human body. balance during walking and standing. J. Neurophysiol., 102(3), 14111419.
ISO 5982-1981, Geneva. Pachi, A., and Ji, T. (2005). Frequency and velocity of people walking.
International Organization for Standardization (ISO). (2007). Bases for Struct. Eng., 83(3), 3640.
design of structures: Serviceability of buildings and walkways against Pandy, M. G. (2003). Simple and complex models for studying muscle
vibrations. ISO 10137:2007, Geneva. function in walking. Phil. Trans. R. Soc. Lond. B, 358(1437), 1501
Kim, S. H., Cho, K. I., Choi, M. S., and Lim, J. Y. (2008). Development of 1509.
human body model for the dynamic analysis of footbridges under pe- Pheasant, S. T. (1982). Anthropometric estimates for British civilian
destrian induced excitation. Int. J. Steel Struct., 8(4), 333345. adults. Ergonomics, 25(11), 9931001.
Kuo, A. D. (2007). The six determinants of gait and the inverted pendulum Pizzimenti, A. D., and Ricciardelli, F. (2005). Experimental evaluation of
analogy: A dynamic walking perspective. Hum. Mov. Sci., 26(4), 617656. the dynamic lateral loading of footbridges by walking pedestrians.
Lee, C. R., and Farley, C. T. (1998). Determinants of the center of mass Proc., Footbridge 2005: 2nd Int. Conf., Venice, Italy.
trajectory in human walking and running. J. Exp. Biol., 201(21), 2935 Racic, V., Pavic, A., and Brownjohn, J. M. W. (2009). Experimental
identication and analytical modelling of human walking forces: Lit-
2944.
erature review. J. Sound Vibrat., 326(12), 149.
Macdonald, J. H. G. (2008). Pedestrian-induced vibrations of the Clifton
Tavares da Silva, F., and Pimentel, R. L. (2011). Biodynamic walking
Downloaded from ascelibrary.org by University of Bristol on 11/27/13. Copyright ASCE. For personal use only; all rights reserved.

Suspension Bridge, UK. Proc. ICE Bridge Eng., 161(2), 6977.


model for vibration serviceability of footbridges in vertical direction.
Macdonald, J. H. G. (2009). Lateral excitation of bridges by balancing
Proc., Eurodyn 2011: 8th Int. Conf. on Structural Dynamics, 10901096.
pedestrians. Proc. R. Soc. A, 465(2104), 10551073. Technical Department for Transport, Roads and Bridges Engineering and
Masani, K., Kouzaki, M., and Fukunaga, T. (2002). Variability of ground re- Road Safety/French Association of Civil Engineering (SETRA/AFGC).
action forces during treadmill walking. J. Appl. Physiol., 92(5), 18851890. (2006) Footbridges: Assessment of vibrational behaviour of footbridges
MATLAB 7.8.0.347 [Computer software]. Natick, MA, MathWorks. under pedestrian loading. Technical Guide 0611, Paris.
Matsumoto, Y., and Grifn, M. J. (1998). Dynamic response of the standing Willford, M. (2002). Dynamic actions and reactions of pedestrians. Proc.,
human body exposed to vertical vibration: Inuence of posture and Footbridge 2002: 1st Int. Conf., Paris.
vibration magnitude. J. Sound Vibrat., 212(1), 85107. 
Zivanovic, S. (2012). Benchmark footbridge for vibration serviceability
Matsumoto, Y., and Grifn, M. J. (2003). Mathematical models for the assessment under the vertical component of pedestrian load. J. Struct.
apparent masses of standing subjects exposed to vertical whole-body Eng., 138(10), 11931202.
vibration. J. Sound Vibrat., 260(3), 431451. 
Zivanovic, S., Daz, I. M., and Pavic, A. (2009). Inuence of walking and
McGeer, T. (1990). Passive dynamic walking. Int. J. Robot. Res., 9(2), standing crowds on structural dynamic properties. Proc., IMAC-XXVII:
6282. Conf. and Exposition on Structural Dynamics: Model Verication and
Mochon, S., and McMahon, T. A. (1980). Ballistic walking. J. Biomech., Validation, Society for Experimental Mechanics, Bethel, CT.
13(1), 4957. 
Zivanovic, S., Pavic, A., and Inglfsson, E. T. (2010). Modeling spatially
National Health Service (NHS). (2009). Health survey for England: 2008 unrestricted pedestrian trafc on footbridges. J. Struct. Eng., 136(10),
trend tables, London. 12961308.

1346 / JOURNAL OF BRIDGE ENGINEERING ASCE / DECEMBER 2013

J. Bridge Eng. 2013.18:1336-1346.

Vous aimerez peut-être aussi