Sei sulla pagina 1di 12

CALIBRATION OF THE UK NATIONAL ANNEX

Chris BARKER David MACKENZIE


Associate Partner
Flint & Neill Partnership Flint & Neill Partnership
London, UK London, UK

Summary
This paper reviews the draft provisions for the UK National Annex for Eurocode EN 1991-2. The Eurocode does not
currently contain provisions for the treatment of pedestrian comfort criteria on footbridges. The proposals given in the
paper describe the background to the procedures and provide guidance in the application of the United Kingdom
National Annex.
The paper lists the draft National Annex provision and provides details of methods that are suitable for the assessment
of the responses for simple structures. These may be used by designers for the design of the majority of footbridges to
ensure that the balance between safety and economy is reached. The paper also recaps the risk based approach
towards pedestrian response, recognising the considerable variation in response between individual pedestrians.
The proposals were developed under a contract with the UK Highways Agency and form the UK’s response to the issue
of pedestrian induced vibrations. Following a peer review process conducted on behalf of the Highways Agency, it was
determined that the codified provisions should provide a framework for designers to work within rather than defining a
single procedure. This enables specialist designers with advanced numerical tools to adapt the rules without penalty.
The contribution of the UK Highways Agency and the British Standards Institution is gratefully acknowledged.
Keywords: footbridge; dynamic; Eurocode; lateral vibration; response; damping; comfort criteria

1. Introduction
There has been considerable effort in recent years to address the shortfall in adequate guidance for dealing with
pedestrian vibration effects. Following on from the well documented episodes of lateral motion exhibited on London’s
Millennium Bridge [1] and Paris’s Solferino Bridge [2] there has been a significant increase in research carried out at a
number of centres culminating in several different approaches to the assessment of pedestrian induced vibration in
footbridges.
This paper sets down the background to the approach adopted in the British National Annex [3] to BS EN 1991-2 [4]. A
key aspect of the National Annex is that it recognises there are a number of valid approaches to the calculation of the
vibration response: the purpose of this paper is to outline a complementary methodology for bridge designers. The
reader is referred to earlier work by Barker [5] and MacKenzie [6] on the development of the theory and practice behind
the standard.

2. Background
The codification of load effects on bridges must attempt to recognise the range of bridge forms and materials likely to be
encountered in design and to ensure that the guidance remains applicable across these forms. In recent years,
footbridges designers have shown a remarkable capacity to adapt and develop conventional bridge forms, producing
some novel and highly complex bridge forms that challenge conventional codes of practice. Allied to this is the
additional complexity of dynamic analysis and, to an extent, the relatively lack of familiarity of many designers with
advanced dynamic concepts.
Therefore in the development of the UK design rules, it was determined that the guidance should be sufficient that it
could take into account the effects of vibration of complicated structures and those in sensitive locations without
imposing undue conservatism that might constrain the designer in achieving and economic solution. It was also

1
recognised that the sensitivity of pedestrians to bridge vibration varied hugely and that a risk based approach was
required where the designer could establish the likely dynamic loads and the sensitivity of the pedestrians to any
potential vibration.
Early forms of the guidance proposed a single probabilistic methodology which has been previously reported [5], but it
was subsequently suggested by the UK Highways Agency that the National Annex should permit the use of acceptable
alternative methodologies. Thus, the National Annex has been modified so that other design approaches may also be
used.

3. UK National Annex to BS EN 1991-2


3.1 Scope
BS EN 1991-2:2003; Eurocode 1: Actions on structures - Part 2: Traffic loads on bridges is the governing standard for
the derivation of bridge loading. The provision for pedestrian sensitivity to bridge vibration has not been included in the
body of the code and has been left to national standards bodies to promulgate national requirements in the National
Annex (NA). These provisions are included in section NA2.44 of the UK National Annex; the titles of the following
sections coincide with those of the NA for ease of reference.
In general, it is assumed that a linear elastic model is appropriate for the calculation of response. Nevertheless,
attention is drawn to the fact that the footbridge as constructed may have natural frequencies which differ from those
calculated (for example due to interaction between structural and non structural parts) and hence respond differently
from predictions. The designer is advised to explore the sensitivity of the contribution of non-structural elements to
investigate potential variation in structural response. As with all structural analysis software, the designer must be
familiar with the modelling methods employed by the software and be aware of the significant variation that can occur in
dynamic response from varying the mass distribution and stiffness of the deck. More complex issues arise when
considering the effects of the cables on deck response.
Typical values for the structural damping values to be used in design are provided below. For cable supported bridges it
will usually be necessary to reduce these values by a further factor of 0.75. Note that structural damping can be difficult
to predict and bridge fittings such as parapets may significantly alter these values. Therefore, it is recommended, before
providing additional damping to a constructed bridge, that its intrinsic damping is measured. Designers should also note
that the value of structural damping may also depend on the mode and load case under consideration.

Table 1 – Recommended damping ratios


Material of construction Damping ratio
Steel 0.03
Steel and concrete composite 0.04
Concrete 0.05
Timber 0.06 to 0.12
Aluminium alloy 0.02
Glass or Fibre Reinforced Plastic 0.04 to 0.08

3.2 General Provisions [NA2.44.1]


The NA assumes an upper limit where the vibration serviceability requirements are deemed to be satisfied: these may be
taken as being when the fundamental natural frequencies of vibration exceed 8Hz for the unloaded bridge in the vertical
direction and 1.5 Hz for the loaded bridge in the horizontal direction. Below these levels, the designer is required to
assess the likely dynamic response of the structure. The approach adopted examines the effects of normal operating
conditions only. It does not consider such effects as mass gathering (for example marathons, demonstrations),
deliberate pedestrian synchronization or vandal loading.
Two distinct analyses are required;
• the determination of the maximum vertical deck acceleration and its comparison with comfort criteria (as
described in NA 2.44.3), and
• an analysis to determine the likelihood of large synchronised lateral responses (as described in NA 2.44.4).

2
3.3 Dynamic Actions to be considered [NA2.44.2]
All bridges are categorised into bridge classes by their usage to determine the actions to be applied. The NA considers
the effects of groups both walking and jogging, as well as specific crowd densities.

Table 2 – Bridge Classification (Table NA.7)

Group Group Crowd density


Bridge Bridge
size size ρ (persons/m2)
Class Usage (walking) (jogging) (walking)
Rural locations seldom used and in sparsely populated
A N=2 N=0 0
areas.
Sub-urban location likely to experience slight variations in
B N=4 N=1 0.4
pedestrian loading intensity on an occasional basis.
Urban routes subject to significant variation in daily usage
C N=8 N=2 0.8
(e.g. structures serving access to offices or schools).
Primary access to major public assembly facilities such as
D N = 16 N=4 1.5
sports stadia or major public transportation facilities.

3.4 Dynamic actions representing single pedestrians and pedestrian groups [NA 2.44.3(a)]
The dynamic actions contained in clause NA 2.44.3(a) represent a group of N pedestrians where;
• The pedestrians in the group make a single crossing of the bridge together.
• One pedestrian in the group is assumed to walk with a pace frequency that is exactly matched to the frequency
of the mode being investigated (the reduced likelihood of this occurring in practice being dealt with
subsequently by the factor k(fv) obtained from figure 2).
• All other pedestrians in the group (N-1) are assumed to walk with phase and pace rates that are randomly
chosen from the pedestrian population model.
• In order to allow for the increased likelihood that pedestrians in small groups may walk with frequencies that are
similar to each other, the standard deviation of the distribution of frequencies used for the group pedestrian
model was been chosen to be 0.1 (as oppose to the value of 0.3 that was used in the development of the crowd
model and the factor k(fv))
In this way the group model aims to provide an assessment of the effect of groups of pedestrians who are walking more
purposefully than those represented by the crowd model.
The method employed in this model attempts to provide a realistic estimate of bridge response by first calculating a
reference response due to a hypothetical standard pedestrian crossing the bridge pacing in time with the natural
frequency of each contributing mode, and then applying a series of modifying factors that deal with population models,
damping, span, group size, etc.
The design maximum vertical accelerations that result from single pedestrians or pedestrian groups should be calculated
by assuming that these are represented by the application of a vertical pulsating force F (N), moving across the span of
the bridge at a constant speed vt, as follows:

F = F0 .k ( f v ). 1 + γ .( N − 1) . sin( 2π . f v .t ) (1)

Where;
N is the number of pedestrians in the group
F0 is the reference amplitude of the applied fluctuating force (N) given in Table 3 (and represents the
maximum amplitude of the applied pedestrian force at the most likely pace frequency)
fv is the natural frequency (Hz) of the vertical mode under consideration

3
k(fv) given in Figure 2, is a combined factor to deal with (a) the effects of a more realistic pedestrian
population, (b) harmonic responses and (c) relative weighting of pedestrian sensitivity to response,
γ is a factor to allow for the unsynchronised combination of actions in a pedestrian group, is a function of
damping and effective span, and is obtained from Figure 3.
Seff is an effective span length (m) equal to the area enclosed by the vertical component of the mode shape
of interest divided by 0.634 times the maximum of the vertical component of the same mode shape.
(In all cases it is conservative to use Seff = S ) – Figure 1

⎛ Area A + Area B ⎞
S eff = ⎜⎜ ⎟⎟
⎝ 0 . 634 .γ max ⎠
γ max
Area A S eff
λ = 0 . 634
S

Area B

Figure 1 Effective span calculation (Figure NA.7)

It may be argued that a pedestrian that is exactly in phase with the mode frequency in region A will become out of phase
with the resulting motion upon entering region B, however in practice (a) for such cases larger responses will occur when
the pedestrian frequency is not exactly matched to that of the mode, and (b) real pedestrian pacing always contains
some natural variation of phase and frequency over time. Thus even though the above allowance for the mode shape
may not model the theoretical pedestrian case perfectly, it is judged to produce responses that are proportional to those
that are likely to occur in practice.

Table 3 – Parameters to be used in the calculation of pedestrian response (Table NA.8)


Load Parameter Walking Jogging
Reference load, Fo (N) 280 910
Pedestrian crossing speed, vt (m/sec) 1.7 3

It should be noted that the pedestrian actions described in the NA correspond to an average pedestrian weight of 700 N.
In general the dynamic action imposed by pedestrian excitation is proportional to the average weight. For bridges that
are to be designed for locations where the national average weight is noticeably different to that of the UK consideration
may be given to adjusting the applied actions accordingly.
Combined population and harmonic

1.4

1.2 Walking
Jogging/running
1
factor, k(fv)

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5 6 7 8
Mode frequency (Hz)

Figure 2 - Relationships between kv(fv) and frequencies fv (Figure NA.8)

4
The above population factor has been obtained by performing a convolution integral of pedestrian crossings that take
into account realistic population parameters and thus allow for the relative likelihood of responses occurring [5].

0.8

0.7
Reduction faction on effective number of pedestrians,γ .

12

0.6
15

0.5
20

0.4
30

0.3

0.2

0.1

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
Structural damping - logarithmic decrement, δ

Figure 3 – Reduction factor, γ, to allow for the unsynchronised combination of pedestrian actions within groups and crowds
(Figure NA.9)
The derivation of the above reduction factor, γ, involves the calculation of convolution integrals of all possible group
responses using pedestrians taken from the assumed population and comparing these with the response that would
occur from N pedestrians all crossing with the same amplitude and frequency, but with random phase. (Because
damping causes the responses of pedestrians whose pace frequency does not coincide well with the resonant frequency
of the mode in question to be less than those that do, the effective number of pedestrians in a group is less than the
number in the group, Neff = 1+γ.(N-1). Increased damping acts to increase group responses because resonant
responses less attenuated and hence the effective number in the group is closer to the actual group size N.)
3.4.1 Simplified method for deriving the maximum vertical acceleration
The code deliberately does not describe methods that engineers may employ to calculate responses from the defined
actions. However, for relatively simple bridges where the response is likely to be dominated by a single mode, simplified
methods may be used. This method described here is very similar to the simplified method of BD37/01 [7] and is valid
only for single span, or two-or-three-span continuous, symmetric superstructures, of constant cross section and
supported on bearings that may be idealised as simple supports.
The maximum vertical acceleration a (m/s2) may be taken as
⎛ F′ 2 ⎞
a = ⎜⎜ .γ i max ⎟⎟.K .Ψ (2)
⎝ Mi ⎠
Where,
F’ is the amplitude of the moving dynamic load (N) given in NA 2.44.3(a)(1).
That is: F ′ = F0 .k ( f v ). 1 + γ .( N − 1)
Mi is the generalised mass of the mode of interest i (kg)
γimax is maximum vertical component of mode shape i.
K is the configuration factor (see Figure 4)
ψ is the dynamic response factor (see Figure 5)

5
Figure 4 – Span configuration factors

60
δ = 0.02
55

50 δ = 0.03
δ = 0.02

45 δ = 0.04
δ = 0.03
Dynamic response factor Ψ

40 δ = 0.05
δ = 0.04

35 δ = 0.05

30

25

20

15

10

0
0 5 10 15 20 25 30 35 40 45 50
Main span length, l (m)

Figure 5. Dynamic response factor ψ

In all other situations the maximum vertical acceleration for single pedestrians and pedestrian groups should be
determined from a time history calculation of the bridge response that explicitly models fluctuating loads passing across
the span using the parameters described in the NA.
3.5 Dynamic load modelling of pedestrians in crowded conditions [NA 2.44.3(b)]
In order to describe the range of possible bridge responses that might be caused by crowd loading it is normal to
describe the responses in a probabilistic manner. A useful measure of the amplitude of the response is the RMS of the
population of possible responses but as this amplitude is exceeded very frequently by large amount the RMS is not
suitable for comparison with comfort criteria. (Note that in the extreme there is a real, but vanishingly small risk that
everyone within a crowd might be fully synchronised for the duration of the loading.)
Mathematical techniques are readily available to obtain design maximum values for given return periods. For a
response frequency around 2Hz it can be shown that in order to provide a design maximum that has an average return
period of 15 minutes requires that 4.0 standard deviations be used. At face value this appears to be a suitable design
maximum however if one examines real response signals it is evident that the use of 4 standard deviations would be

6
unnecessarily penal.
Figure 6 below illustrates a synthetic time history produced from a random response population model centered on 2Hz.
Horizontal lines mark the boundaries of ±4.0, ±2.5 and ±1.0 standard deviations on this typical 7½ minute time history.

a
100 150 200 250 300 350 400 450 500
Time (seconds)

Figure 6 – Synthetic Time History


It has been decided that 2.5 standard deviations is an appropriate design maximum value to be used for crowd
responses, and the loading provided by NA 2.44.3(b) incorporates this assumption. This permits short durations of
response that are above the target comfort criteria but the average level of comfort is significantly better.
Other design guides, e.g. Sétra [2], use design maximum accelerations that are more closely equivalent to the 4
standard deviations illustrated above. In certain circumstances, with the agreement of the relevant authority, the
designer may choose to apply a more stringent limit to the maximum responses by increasing the design maximum
loading from 2.5 to 4 standard deviations.
In very crowded conditions the mean velocity of pedestrians tends to reduce as the density increases and the population
statistics change. In the calculation of vertical responses to NA 2.44.3(b) the maximum crowd density of 1 persons/m2 is
provided to reflect this effect (figure 7).

2.5

2
Density (persons/m2)
Speed

1.5

0.5

0 0.5 1 1.5 2 2.5 0


Density (persons/m2) Flow rate

Figure 7- reduction of pedestrian flow rates with density


At the time of writing some current research suggests that in crowded conditions the pedestrians mass may also act to
add vertical damping to the bridge and so further reduce the responses. Should this effect be proven to be a reliable
source of energy dissipation the extra damping provided may be added to estimate of structural damping and in all other
respects the response calculation will be unaltered. However, under no circumstances should such damping be utilised
in the avoidance of unstable lateral responses.

7
In addition there is rather more evidence to suggest that simple probabilistic models of crowd behaviour such as that
used as the basis for NA 2.44.3(b) tend to produce calculated responses that are somewhat larger than field
measurements of comparable cases; however at the time of writing these cannot be relied on for design and further
research is required in this area.
The design maximum vertical accelerations that result from pedestrians in crowded conditions should be calculated by
assuming that these are represented by a vertical pulsating distributed load w (N/m2), applied to the deck for a sufficient
time so that steady state conditions are achieved as follows:
⎛F ⎞
w = 1.8⎜ 0 ⎟.k ( f v ). γ .N / λ . sin(2π . f v .t ) (3)
⎝ A⎠
Where,
N is the total number of pedestrians distributed over the span S.
N = ρ.A = ρ.S.b
ρ is the required crowd density reference taken from Table 1 but with a maximum value of 1.0 persons/m2.
(This is because crowd densities greater than this value produce less vertical response as the forward
motion slows.)
S is the span of the bridge (m)
A is the area of the walking surface of the bridge (m2)
b is the width of the walking surface of the bridge (m)
γ is a factor to allow for the unsynchronised combination of actions in a crowd and is obtained from Figure
3.
λ is a factor that reduces the effective number of pedestrians in proportion to the enclosed area of the
mode of interest. λ = 0.634(Seff / S).

In order to obtain the most unfavourable effect this loading should be applied over all relevant areas of the footbridge
deck, the sign of the force should be varied to match the sign of the vertical displacements of the mode for which
responses are being calculated.
3.5.1 Method for deriving the maximum vertical acceleration in crowded conditions
Again it should be noted that the code deliberately does not describe methods that engineers may employ to calculate
responses from the actions defined above. This method is generally applicable for spans where the response is
dominated by a single mode.
The maximum vertical acceleration a (m/s2) may be taken as
⎛ S w′.b. γ ( x) .dx ⎞
a=⎜
⎜ 0 ∫ i ⎟ 1
.γ i max ⎟. (4)
⎜ Mi ⎟ 2ξ
⎝ ⎠
Where,
w’ is the amplitude of the vertical dynamic load (N/m2) given in NA 2.44.3(b)(1).
⎛F ⎞
That is: w′ = 1.8⎜ 0 ⎟.k ( f v ). γ .N / λ
⎝ A⎠
b is the width of the walking surface of the bridge (m)
Mi is the generalised mass of the mode of interest i (kg)
γi(x) is the vertical component of mode shape i at position x (m) along the span S (m).
Note that in the above equation the absolute, positive value of the mode amplitude is used in the integration.
γimax is maximum vertical component of mode shape i.
ξ is the structural damping when expressed as a damping ratio, ξ = δ / 2π .

For uniform width spans where the mode shape is approximately sinusoidal the above equation may be simplified further
to,

8
⎛ 0.634.w′.b.S 2 ⎞ 1
a = ⎜⎜ .γ i max ⎟⎟. (5)
⎝ Mi ⎠ 2ξ

3.6 Recommended serviceability limits for use in design [NA 2.44.3(c)]


The responses obtained using the above methods, for both single pedestrians and crowded conditions, represent a
nominal load condition which needs to be modified to take account of the relative sensitivity of the design to differing site
conditions. These remain unchanged from those previously reported by MacKenzie [6].
The maximum vertical acceleration calculated from the above actions should be less than the design acceleration limit
given by:
a limit = 1.0 k1 k 2 k 3 k 4 m/s2 (6)
and 0.5 m/s2 ≤ a lim it ≤2.0 m/s2
Where the response modifiers k1 , k2 and k3 are taken from Tables 4 to 6 in which:
k1 changes the pedestrian sensitivity according to the site usage
k2 is a factor to allow for route redundancy
k3 changes the pedestrian sensitivity according to the height of structure
k4 is an exposure factor is to be taken as 1.0 unless otherwise agreed between the designer and the
relevant authority.

Table 4 – Recommended values for the site usage factor k1 (Table NA.9)
Bridge Function k1
Primary route for hospitals or other high
0.6
sensitivity routes
Primary route for school.
Primary routes for sports stadia or other 0.8
high usage routes
Major urban centres 1.0
Suburban crossings 1.3
Rural environments 1.6

Table 5 – Recommended values for the route redundancy factor k2 (Table NA.10)
Route Redundancy k2
Sole means of access 0.7
Primary route 1.0
Alternative routes readily available 1.3

Table 6 – Recommended values for the structure height factor k3 (Table NA.11)
Bridge Height k3
Greater than 8 m 0.7
4 to 8 m 1.0
Less than 4 m 1.1

Values of k1, k2 and k3 other than those given in Tables 4 to 6 may be agreed between the designer and the relevant

9
authority. These values may be derived using Figure 8 as a guide. Intermediate values may be used and agreed
between the designer and the relevant authority.

2
Response modifier, ki

1.5

0.5

Primary route for Primary route for


Site usage factor, k1
Function Major Urban Centres Suburban Crossings Rural Environments
hospital school
Route redundancy Alternative routes readily available without
Redundancy
factor, k2
Solemeans
Sole meansofofaccess
access Primary Route
additional hazard to user
Height of structure
Structure
factor, k 3
Height Abovethan
Greater 8 m8 m 4-8 m 4 - 8 m Below 4m
Less than 4 m

Figure 8 – Classification of pedestrian response (Figure NA.10)


With the agreement of the relevant authority k4 may be assigned a value of between 0.8 and 1.2 to reflect other
conditions that may affect the users’ perception towards vibration. These may include consideration of parapet design
(such as height, solidity or opacity), quality of the walking surface (such as solidity or opacity) and provision of other
comfort-enhancing features.
For specialist bridges or bridges in remote locations, less onerous design limits may be applied with the agreement of
the relevant authority, subject to the condition that a suitable risk assessment is undertaken.
3.6.1 Damage from forced vibration
Consideration should be given to the possibility of permanent damage to a superstructure by a group of pedestrians
deliberately causing resonant oscillations of the superstructure. As a general precaution, therefore, the bearings should
be of robust construction with adequate provision to resist upward or lateral movement.
For prestressed concrete construction, resonant oscillation may result in a reversal of up to 10% of the static live load
bending moment. Providing that sufficient unstressed reinforcement is available to prevent gross cracking, no further
consideration need be given to this effect.

3.7 The avoidance of unstable lateral responses due to crowd loading [NA 2.44.4]
The stability condition described in NA 2.44.4 is significantly different from the comfort criteria described in respect of
vertical vibrations. While the level of comfort experienced by bridge users is very subjective and varies from person to
person and site to site, the lateral stability condition has a measurable and clearly defined limit that should not be
exceeded if large uncontrolled lateral motions are to be avoided.
Calculate the pedestrian excitation mass damping parameter D:
mbridge .ξ
D= (7)
m pedestrian
Where:
mbridge is the mass per unit length of the bridge
m pedestrian is the mass per unit length of pedestrians obtained from Table 2
ξ is the structural damping when expressed as a damping ratio, ξ = δ / 2π .

For all deck modes of vibration having a significant lateral horizontal component and a frequency below 1.5 Hz compare
the pedestrian mass damping parameter and the mode frequency with the stability boundary defined in Figure 9 below.
If the pedestrian mass damping parameter for the relevant crowd density obtained from Table 2 falls below the indicated
boundary divergent lateral responses may be expected. Values above the line should be stable.

10
At the time of writing reliable test measurements are only available for footbridge lateral frequencies in the range of 0.5
to 1.1Hz. The extensions to the stability curve beyond this region are based upon a theoretical model of response only
and should be used with caution.

1.8
Pedestrian mass damping parameter, D

1.6

1.4

1.2
Stable
1

0.8

0.6
Unstable
0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Frequency of lateral mode (Hz)

Figure 9 – Lateral lock-in stability boundary (Figure NA.11)


The process provided in NA 2.44.4 is based on a uniform deck mass. Lower critical densities may arise where the deck
mass per length is irregular; a conservative approach would be to use the lowest mass per unit length. The method may
also be conservative when other parts of the structure, such as arch ribs or suspension cables add significantly to the
modal mass. In these circumstances specialist advice is sought.

4. Conclusions
The process outlined above provides a framework for the assessment of the response of footbridges to pedestrian
induced vibrations. The method strikes a balance between an over-simplistic empirical basis that would penalise lighter,
more aesthetic solutions and a complex ‘first principal’ approach that would require undue computational effort and
technical understanding for the vast majority of structures.
However, in undertaking the development of the design rules for this project, it has become clear that many of the
innovative structural forms that are emerging in footbridge design regularly push the limits of codified methods.
Therefore the processes given within the NA have been developed to provide designers with the tools to reduce the risk
of unacceptable vibrations. It should be emphasised that the risks are reduced not removed and that designers have a
duty of care to investigate novel and innovative structures sufficiently to confirm that there are no unexpected dynamic
effects. Primary areas of concerns are the use of stay cables, large plan curvature of decks, separation of deck
elevations and unusual articulation requirements.
The work has indicated that there is a need for more research into the effects of crowd loading on footbridges. At
present, most codified methods are an extrapolation of single pedestrian effects with due account made for the
randomness of the input. There is a risk that codified methods can be conservative for large unsynchronised crowds and
it would be of value to current code provision for better data on the positive (and negative) effects of crowd induced
damping.

5. References
[1] DALLARD P.,FITZPATRICK T.,FLINT A., LOW A., RIDSDILL SMITH R., WILLFORD M., ROCHE M., “London
Millennium Bridge: Pedestrian-Induced Lateral Vibration”, J. of Bridge Engineering, Trans. ASCE, 6, 412-417
(2001)..

11
[2] SETRA, “Assessment of vibrational behaviour of footbridges under pedestrian loading” , Paris, October 2006.
www.setra.equipement.gouv.fr/IMG/pdf/0644A_Footbridges.pdf
[3] BRITISH STANDARDS INSTITUTE NA to BS EN 1991-2: UK National Annex to Eurocode 1: Actions on
structures - Part 2: Traffic loads on bridges (to be published)
[4] BRITISH STANDARDS INSTITUTE BS EN 1991-2: 2003: UK National Annex to Eurocode 1: Actions on
structures - Part 2: Traffic loads on bridges
[5] BARKER D., DeNEUMANN S., MacKENZIE D., KO R., “Footbridge Pedestrian Vibration Limits - Part 1:
Pedestrian Input”, Footbridge 2005 International Conference
[6] MacKENZIE D., BARKER C., McFADYEN N., ALLISON B., “Footbridge Pedestrian Vibration Limits - Part 2:
Human Sensitivity”, Footbridge 2005 International Conference
[7] DESIGN MANUAL FOR ROADS AND BRIDGES: Volume 1 Highway Structures: Approval Procedures and
General Design: Section 3 General Design: Part 14: BD 37/01 – Loads for Highway Bridges

12

Potrebbero piacerti anche