Next Article in Journal
Game-Based Flexible Merging Decision Method for Mixed Traffic of Connected Autonomous Vehicles and Manual Driving Vehicles on Urban Freeways
Previous Article in Journal
High-Resolution Identification of Sound Sources Based on Sparse Bayesian Learning with Grid Adaptive Split Refinement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analytical Investigation of the Effects of Additional Load Mass on the Fundamental Frequency of Pedestrian Beam Bridges

by
Marija Spasojević Šurdilović
*,
Andrija Zorić
,
Srđan Živković
and
Dragana Turnić
Faculty of Civil Engineering and Architecture, University of Niš, 18000 Niš, Serbia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(16), 7369; https://doi.org/10.3390/app14167369 (registering DOI)
Submission received: 19 July 2024 / Revised: 17 August 2024 / Accepted: 19 August 2024 / Published: 21 August 2024
(This article belongs to the Section Civil Engineering)

Abstract

:
The aspect of resonant vibrations due to pedestrian movement is of great significance in engineering practice. Therefore, understanding the variations in the natural frequency of bridge structures under varying positions of additional mass is of particular interest. This paper presents a procedure for the straightforward determination of the natural frequencies of a beam pedestrian bridge for various positions of pedestrians or a service vehicle based on derived analytical solutions. The calculation takes into account the inertial effects of the additional load mass, modeled as either uniformly distributed or concentrated. The importance of additional load mass effects on the fundamental frequency of a beam pedestrian bridge and its dynamic response to a moving pedestrian load is demonstrated on a bridge example. The proposed solutions are also applicable to other girder system structures with uniform mass and stiffness along their span.

1. Introduction

Modern pedestrian bridges are evolving to be more innovative in their design, not only serving functional purposes but also acting as prominent urban landmarks. Advances in structural understanding and new materials have enabled longer spans and more slender structures [1,2]. In engineering practice, particularly concerning pedestrian bridges, resonant vibrations due to pedestrian traffic are highly significant. This is particularly understandable because footbridges are increasingly designed with lower frequencies and lighter weights, which can cause their fundamental vibration frequency to align with the walking frequency of pedestrians [3,4]. This alignment often leads to potential issues with vibration comfort, highlighting the importance of understanding how the natural frequencies of bridge structures change with different positions of additional masses.
As a result of specific incidents related to footbridge vibration issues over the past several decades such as the Millennium Bridge in London [5,6,7] and the Solferino Bridge in Paris [8,9], research has focused on enhancing design standards that consider pedestrian loads, traffic density, and comfort requirements [10]. Structures with natural frequencies outside the range of 1–3 Hz, which is typically induced by pedestrian loads [11], are not at risk of resonant vibrations according to many regulations and standards and, therefore, do not require further dynamic analysis.
According to the Eurocode [12], dynamic analysis is necessary for pedestrian bridges if their natural frequencies fall within the critical range of 0.5 to 5.0 Hz. Meanwhile, the AASHTO LRFD Guide Specifications for the Design of Pedestrian Bridges [13] indicate that pedestrian-induced vertical vibrations are negligible if a bridge’s fundamental vertical frequency exceeds 3.0 Hz, and transverse vibrations are negligible if the transverse frequency exceeds 1.3 Hz. If these conditions are not met, AASHTO recommends conducting a dynamic analysis according to European Standards, specifically following SETRA [14] for methods of analysis, loading cases, and criteria for acceleration acceptance. The serviceability assessment is generally based on a comparison between the pedestrian-induced acceleration and a suitably defined limit value [15].
In dynamic analysis, a significant aspect involves monitoring the shift in the fundamental frequency caused by additional masses, such as groups of pedestrians standing in specific sections of a bridge while others are in motion. It is essential to pinpoint the critical mass configuration capable of pushing a system into an undesirable frequency range, potentially inducing resonant vibrations. Such vibrations could compromise pedestrian comfort and raise concerns regarding a bridge’s stability. The first and simplest way to simulate people’s presence on a structure is by modeling human occupants as added masses [16,17], which can effectively explain reductions in natural frequencies. However, to realistically explain a dynamic response, it is preferable to use models that consider human–structure interaction when simulating the presence of people on a footbridge [18,19,20,21,22,23,24,25]. Analytical models for analyzing the vibrations of a beam under the action of a harmonic moving mass [26] and moving mass with constant [27] and variable speed [28], as well as for the evaluation of vertical vibrations of footbridges using response spectrum [29], have been proposed.
Regarding the current codes of practice [14,30], the influence of a static pedestrian mass, corresponding to up to 5% of a bridge mass, can be ignored since it results in a decrease in natural frequency of less than 2.5%. Several studies, including those by Ellis et al. [31] and Maraveas et al. [32], some relying on full-scale measurements, have demonstrated that passive humans (sitting and standing) significantly impact the dynamic properties of a structure. Typical findings indicate a notable reduction in vibration response and slight alterations in a structure’s natural frequency and damping [33,34].
Busca et al. [35] introduced a method to predict changes in a structure’s modal parameters due to passive individuals, requiring accurate knowledge of the empty structure’s modal characteristics, where each passive individual is modeled by their apparent mass and integrated into the structure without restrictions on the degrees of freedom. In their study, Yang et al. [36] investigated frequency variations in a vehicle–bridge interaction system subjected to a moving mass, which can be significant from the aspect of the movement of a service vehicle on a pedestrian bridge. They employed a three-dimensional model that incorporated the effects of inertial and centrifugal forces from the moving mass into the bridge’s motion equations. O’Sullivan et al. [37] suggested that the mass of pedestrians, along with any additional weight they carry in urban environments, could significantly contribute to the vibrations experienced by a bridge.
In addition to the effects of the additional mass on the dynamic characteristics of bridges, the influence of the stiffness of different structural parts on the dynamic properties of a timber bridge was experimentally analyzed [38]. Human–structure interactions were also analyzed in the case of aluminum footbridges [39]. Recently, studies have focused on the vibration control of pedestrian bridges using dynamic vibration absorbers [40,41,42].
This paper introduces a procedure that relies on analytical solutions to straightforwardly determine the natural frequencies of a beam pedestrian bridge with additional mass, which, to the best of the authors’ knowledge, has not been derived before. It applies to various positions of pedestrians or service vehicles on a bridge, where their mass is modeled either as uniformly distributed or concentrated. In one bridge example, the importance of the additional load mass effects on the fundamental frequency of a beam pedestrian bridge and its time–history response due to the dynamic moving pedestrian load is highlighted.

2. Problem Statement

Let us consider the transverse vibrations of a beam with constant bending stiffness EI and mass μ along span L, as shown in Figure 1. It is assumed that the effects of rotary inertia of the cross-section and shear deformation on bending are negligible in the vibration analysis. Furthermore, taking into account the assumption that the energy dissipation linearly depends on the mass and stiffness of the structure, the partial differential equation for the considered problem of the transverse vibrations of the beam can be defined as follows [43,44]:
μ 2 ν x , t t 2 + 2 β μ ν x , t t + E I α t 4 ν x , t x 4 + E I 4 ν x , t x 4 = p x , t .
In Equation (1), v(x,t) represents the deflection at point x and time t. The load on the bridge, whether stationary or moving at a constant speed (c = const.), is described by the force p(x,t), which, without considering its inertial effect, is defined as a deterministic disturbance force. The second and third terms in Equation (1) represent damping forces, where the damping constant β expresses the proportionality of the damping force with respect to the mass μ and constant α expresses that with the stiffness EI of the system.
Using modal analysis and considering the orthogonality of eigenfunctions, Equation (1) can be reduced to a system of independent equations of the following form:
η ¨ r t + 2 ξ r ω r η ˙ r t + ω r 2 η r t = 1 M r F r t ,
where ηr(t) denotes a principal coordinate of the r-th mode, ωr is the natural circular frequency of the r-th mode of the system, and ξr represents the relative damping of the r-th mode.
In Equation (2) Mr stands for the generalized mass of the r-th mode of the system, which is defined as follows:
M r = μ 0 L V r 2 x d x ,
and Fr is the generalized force of the r-th mode:
F r t = 0 L p x , t V r x d x .
Applying the Laplace transform to Equation (2), the most general expression for the dynamic deflection of a beam, as a solution of Equation (1), based on which various specific cases can be analyzed, results in the following [43]:
v x , t = r = 1 V r x M r 0 t 0 L p u , τ V r u d u sin ω d r t τ ω d r e ξ r ω r t τ d τ   + r = 1 V r x 0 L V r 2 x d x [ cos ω d r t 0 L v 0 r u V r u d u   + sin ω d r t 0 L ξ r ω r v 0 r u + v ˙ 0 r u ω d r V r u d u ] e ξ r ω r t ,
where Vr(x) denotes the eigenfunction of the r-th mode, v 0 r x = v r x , 0 a n d v ˙ 0 r x = v ˙ r x , 0 are the initial conditions, τ is the time variable of force p, u is the position variable, and ωdr is the circular frequency of free damped vibrations of the r-th mode of the system, which is defined as follows:
ω d r t = ω r 1 ξ r 2 .
The effects of additional load mass on the free transverse vibrations of an elastic oscillatory system (Figure 2) with constant bending stiffness EI and mass μ along span L are the main focus of this research.
The constant mass μ, per unit length of the beam, is increased by an additional uniformly distributed mass μq, per unit length λ, located at the position defined by the abscissa x0. It is assumed that the contact between the beam and the additional masses is continuously maintained.
Such a configuration of additional masses to a beam structure is most commonly encountered in pedestrian bridges, where pedestrians either stand still or move at a speed represented by c, with x0 = ct. The problems of undamped vibrations, starting from Equation (1), and neglecting the damping forces (second and third terms in Equation (1)), can be described by the following differential equation:
μ 2 ν x , t t 2 + E I 4 ν x , t x 4 = p x , t .
Considering the inertial effect of the additional mass µq as the force p(x,t), whose position is defined by the Heaviside function:
p x , t = μ q 2 ν x , t t 2 H x x 0 + λ 2 H x x 0 λ 2 ,
one can obtain differential equation of undamped free transverse vibrations of a beam with an additional load mass [44]:
μ 2 ν x , t t 2 + E I 4 ν x , t x 4 = μ q 2 ν x , t t 2 H x x 0 + λ 2 H x x 0 λ 2 ,
with boundary conditions dependent on the type of support of the beam system, as well as initial conditions:
v x , 0 = v 0 x ;   v ˙ x , 0 = v ˙ 0 x .
Employing modal analysis and introducing an approximation for the inertial effect of the additional masses as following:
μ q 2 ν x , t t 2 μ q η ¨ r t V r x ,
differential Equation (9) is reduced to a system of r independent equations, which can be symbolically represented in the following form:
μ η r ¨ t V r x + E I η r t d 4 V r x d x 4 + μ q η r ¨ t V r x H x x 0 + λ 2 H x x 0 λ 2 = 0 ,
η r ¨ t μ 0 L V r 2 x + μ q 0 L V r 2 x H x x 0 + λ 2 H x x 0 λ 2 + E I η r t 0 L d 4 V r x d x 4 V r x d x = 0 ,
η ¨ r t + ω ¯ r 2 η r t = 0 ,
where ηr(t) denotes a principal coordinate of the r-th mode and ω ¯ r is the corresponding circular frequency of the r-th mode of the system with the additional mass:
ω ¯ r 2 = E I 0 L d 4 V r x d x 4 V r x d x μ 0 L V r 2 x + μ q 0 L V r 2 x H x x 0 + λ 2 H x x 0 λ 2 d x
ω ¯ r 2 = E I μ 0 L d 4 V r x d x 4 V r x d x 0 L V r 2 x 1 + μ q μ 0 L V r 2 x 0 L V r 2 x H x x 0 + λ 2 H x x 0 λ 2 d x
For ω ¯ r in Equation (16), Relation (17) applies as follows:
ω ¯ r = ω r 1 + G r
where:
G r = 1 M r μ q x 0 λ 2 x 0 + λ 2 V r 2 x d x .
It should be noted that Mr in Relation (18) represents the generalized mass of the r-th mode of the system without the additional mass µq.
Since the contribution of higher modes is typically of lesser significance to the overall vibrations, it is evident that Relation (17) is acceptable for the fundamental mode (r = 1), and for several subsequent modes where deviations become more pronounced, particularly for higher modes when the ratio of additional mass to bridge mass is larger.
In the case of pedestrian bridges, it is practically important to understand the change in the frequency of the fundamental mode due to the pedestrian or service vehicle mass, which Relation (17) satisfies, specifically for r = 1.

3. Analytical Solutions for Beams with One, Two, and Three Equal Spans

The application of the presented theoretical procedure will be demonstrated on beams with span L, specifically for a beam with three equal spans (L = 3l), a beam with two equal spans (L = 2l), and a beam with a single span (L = l).
Analysis was conducted for three characteristic cases of additional mass configurations. The first case involves a uniformly distributed mass μq over length λ, the second case concerns the approach of a uniformly distributed mass μq on the beam, and the third case examines the impact of a concentrated mass m. The dynamic models of those three cases are further presented in Section 3.1, Section 3.2 and Section 3.3.
Sinusoidal vibrations were considered, with the following eigencharacteristics applying to all three beams:
  • The eigenmode shape function for mode r, which satisfies boundary conditions:
V r x = sin r π x l ,
  • Natural frequency of the r-th mode:
ω r = r 2 π 2 l 2 E I μ ,
  • Generalized mass of the r-th mode:
M r = μ 0 L sin 2 r π x l d x = μ L 2 ,                 L = n l ,                 n = 1 , 2 , 3 .
For a beam with a single span, eigencharacteristics (19)–(21) apply to all modes, while for beams with two and three equal spans, they pertain to the fundamental mode and higher modes only of sinusoidal nature. Since emphasis is placed on the fundamental mode, the analysis for higher modes of non-sinusoidal nature is omitted here.

3.1. Influence of Additional Mass over Length λ

The frequency ω ¯ r for this case is determined from Equation (17) in accordance with Relations (19)–(21) as follows:
G r = μ q M r x 0 λ 2 x 0 + λ 2 V r 2 x d x = μ q M r x 0 λ 2 x 0 + λ 2 s i n 2 r π x l d x = μ q M r · l 2 · λ 2 1 r π s i n r π λ l c o s 2 r π x 0 l ,
ω ¯ r = ω r 1 + G r = ω r 1 + μ q M r · l 2 · λ 2 1 r π s i n r π λ l c o s 2 r π x 0 l 1 2 ,
ω ¯ r = ω r 1 + μ q M r l 2 Φ a λ , x 0 1 2 ,
where:
Φ a λ , x 0 = λ l 1 r π sin r π λ l cos 2 r π x 0 l .
Equation (24) is applicable to beams with one, two, or three equal spans. For such beams, the generalized mass Mr should be determined as specified in Equation (21). The specific expressions for the frequencies of systems with these span configurations are provided in Table 1.
Changes in fundamental frequency due to additional mass μq along length λ = 0.2l, obtained using the expressions in Table 1 and depicted through the ratio of circular frequencies, are presented in Figure 3 for bridges with one, two, and three equal spans. One may conclude that the additional mass in the mid-span has a larger effect on the natural frequency. Increasing the ratio of additional mass to bridge self-mass amplifies this effect.

3.2. Influence of the Arrival of a Uniformly Distributed Mass

The frequency ω ¯ r for this additional mass case can be obtained using Equations (24) and (25), where introducing the shift x0 = λ and rearranging with a translation of x0 by λ/2 can be represented as follows:
G r = μ q M r 0 x 0 V r 2 x d x = μ q M r 0 x 0 s i n 2 r π x l d x = μ q M r · l 2 · x 0 l 1 2 r π s i n 2 r π x 0 l ,
ω ¯ r = ω r 1 + G r = ω r 1 + μ q M r l 2 x 0 l 1 2 r π s i n 2 r π x 0 l 1 2 ,
ω ¯ r = ω r 1 + μ q M r l 2 Φ b x 0 1 2 ,
where:
Φ b x 0 = x 0 l 1 2 r π sin 2 r π x 0 l .
Relation (28) is relevant for beams with one, two, or three equal spans. For the analysis of these beams, the generalized mass Mr should be calculated according to Relation (21). Detailed expressions for the frequencies of such systems are listed in Table 2.
Changes in fundamental frequency due to arrival of uniformly distributed mass μq defined based on Table 2 and compared with the circular frequencies are presented in Figure 4 for bridges with one, two, and three equal spans. It is evident that the largest effect on the natural frequency is obtained when additional mass is distributed over the entire span (x0/l = 1 for a bridge with one span, x0/l = 2 for a bridge with two equal spans, and x0/l = 3 for a bridge with three equal spans). Similar to the previous case, increasing the ratio of additional mass to bridge self-mass amplifies this effect.

3.3. Influence of Concentrated Mass

The influence of concentrated mass on a pedestrian bridge structure may be more significant for loading from service or firefighting vehicles, whose movement in some cases may be permitted but practically negligible during pedestrian traffic.
The solution for this case can be obtained from Equations (24) and (25) if one considers λ to be infinitely small and the mass μqλ to have a finite value m. Then, as λ→0, sin r π λ l becomes r π λ l ; thus, the general solution for ω ¯ r in this case results from the following:
G r = μ q M r x 0 λ 2 x 0 + λ 2 V r 2 x d x = μ q M r x 0 λ 2 x 0 + λ 2 s i n 2 r π x l d x = μ q M r · l 2 · λ 2 1 r π · r π λ l c o s 2 r π x 0 l = m M r s i n 2 r π x 0 l
ω ¯ r = ω r 1 + G r = ω r 1 + m M r s i n 2 r π x 0 l 1 2 ,
ω ¯ r = ω r 1 + m M r Φ c x 0 1 2 ,
where:
Φ c x 0 = sin 2 r π x 0 l .
For beams with one, two, or three equal spans, the generalized mass Mr should be computed following Relation (21). The frequency expressions corresponding to these configurations are summarized in Table 3.
For this analyzed case of additional concentrated mass, the changes in the fundamental frequencies of bridges with one, two, and three equal spans, depicted through the ratio of circular frequencies, are presented in Figure 5. Similar conclusions apply to this case as for the scenario with additional mass distributed over length λ.

4. Application Example

The importance of additional load mass effects on the fundamental frequency of a beam pedestrian bridge, as well as its time–history response and acceleration due to dynamic moving pedestrian loads, is illustrated through a specific example. A simply supported timber footbridge with a span length L = 25 m, bending stiffness of EI = 2.016 × 106 kNm2, mass per unit length μ = 400 kg/m, damping ratio β = 0.015, and an additional mass equivalent to that of a light service vehicle (m = 5000 kg) is considered. The numerical demonstration presented here is solely intended as an illustrative tool, utilizing assumed parameters. Importantly, these parameters correspond to a hypothetical pedestrian bridge scenario.
The bridge’s first-order vertical frequency was 5.64 Hz. According to Eurocode [12], it is unnecessary to verify whether the acceleration response under pedestrian loads meets criteria for walking comfort. Considering a pedestrian bridge with an additional mass equivalent to that of a light service vehicle, it is concluded from Figure 5a that placing the additional mass at mid-span would result in the greatest change in the natural frequency. Thus, according to Expression (32) and Table 3, the frequency of the system with the additional mass was determined to be f 0 ¯ = 3.98 Hz. Given that the frequency of the bridge, with a service vehicle present, corresponds to the running frequency, there is a possibility that the bridge could resonate with a pedestrian jogging across it, thereby experiencing excessive vibrations in terms of the serviceability limit state.
The maximum vertical acceleration of the bridge with the additional mass was determined by assuming a resonant pedestrian jogging at a constant speed. This result was compared with the response of the structure without the additional mass under the same pedestrian movement conditions. Both obtained results were evaluated according to the comfort criteria proposed by relevant standards.
The pedestrian’s dynamic load [45,46] was modeled as a pulsating load P(t), moving across the bridge at a constant speed c:
P t = 180   s i n 2 π f 0 ¯ t   N ,
c = 0.9 · f 0 ¯ = 3.59   m s .
The dynamic response was calculated using Equation (5), where the acceleration of the bridge was computed as the second derivative of deflection at x = L/2, and pedestrian load, P(x,t), is defined using the Dirac delta function as follows:
P x , t = P t δ x c t .
Figure 6 shows the time histories of the dynamic response at the mid-span of the pedestrian bridge due to a resonant pedestrian load, considering the bridge with additional mass.
Figure 7 shows the time histories of the dynamic response at the mid-span of the pedestrian bridge due to the same pedestrian load without the additional mass. Therefore, the pedestrian load P(x,t), as given by Equation (36), was not in the resonant domain of the bridge.
Human perception of vibration is complex, with different individuals, and even the same person at different times, experiencing the same vibration source differently [41]. To address pedestrian discomfort from excessive vibrations on footbridges, various countries have developed comfort assessment methods based on acceptable vibration limits for the human body.
SETRA [14] categorizes vertical vibrations by maximum accelerations into three comfort levels: up to 3.535%g (practically imperceptible), between 3.535%g and 7.07%g (merely perceptible), and from 7.07%g to 17.675%g (perceptible but tolerable), where g denotes gravitational constant (approximately 9.81 m/s2). These levels reflect theoretical accelerations of a bridge deck under pedestrian load, with accelerations exceeding 25%g deemed unacceptable.
On the other hand, the European standard [12] defines the maximum acceptable vertical accelerations of a bridge as 0.7 m/s2 to fulfil the serviceability limit state. It should be noted that this clausula is not mandatory, allowing each country to define different criteria in their national annexes. In the case of the national annex of the Republic of Serbia [47], the maximum acceptable vertical acceleration of a footbridge depends on factors such as bridge site usage, route redundancy, bridge structure height, and exposure aspects. For example, if a bridge serves as a primary route in a major urban center and has a height of less than 4 m, according to this national annex, the maximum acceptable vertical acceleration for the footbridge is 1.1 m/s2.
Based on the presented analysis, it can be concluded that the hypothetical bridge structure met the serviceability limit state regarding vibrations when additional load mass was not considered. However, in the presence of additional load mass, this criterion was not fulfilled.

5. Conclusions

The derived analytical solutions for variations in the fundamental frequency of a pedestrian beam bridge due to additional mass, modelled as either uniformly distributed or concentrated mass, can be applied to a specific subset of pedestrian beam bridges where the bending stiffness and mass distribution along the span remain constant. It is acknowledged that the inherent dynamic characteristics of the oscillatory system (bridge without pedestrians) are known. It is essential to emphasize that additional masses always lead to a reduction in the fundamental frequency that depends on their size and position on the bridge.
The presented analytical solutions are highly suitable for both qualitative and quantitative analyses of the problem of fundamental frequency variation, which is always of particular interest in engineering practice. Since the fundamental frequency is a relevant parameter for the vibration analysis of pedestrian bridges, based on the derived analytical solutions or the presented diagrams, it is easy to determine when neglecting the inertial effects of added masses is acceptable.
The application example of a hypothetical pedestrian bridge scenario, with additional mass from a light service vehicle that is 50% of the bridge’s self-weight, showed that the vertical frequency of the bridge decreased by approximately 30% compared with the same bridge without the additional mass. Furthermore, for the same pedestrian excitation, the maximum vertical acceleration of the bridge increased by approximately 50 times due to the additional mass of a light service vehicle at the bridge’s mid-span. This example demonstrates the importance of understanding changes in bridge frequency due to additional mass, illustrating how pedestrian comfort when crossing the same footbridge without and with additional mass can vary from maximally acceptable to totally unacceptable levels. This conclusion points out that presence of an additional mass on the footbridge should not be neglected in the serviceability limit state regarding vibrations, emphasizing the necessity to develop existing standards in this area.
To fully validate the proposed analytical solution for determining the natural frequencies of a pedestrian bridge with additional mass, experimental testing and a comparison of the experimental and analytical results should be conducted. This will be the focus of the authors’ future work. Further research can also be directed toward developing an analytical solution for additional mass effects on footbridges with non-uniform bending stiffness and mass along the span, aspects commonly encountered in real engineering practice but not covered in the analysis presented in this paper.

Author Contributions

Conceptualization, M.S.Š. and A.Z.; methodology, S.Ž. and D.T.; validation, M.S.Š., A.Z., D.T. and S.Ž.; formal analysis, M.S.Š. and A.Z.; investigation, M.S.Š. and A.Z.; resources, S.Ž. and D.T.; data curation, D.T.; writing—original draft preparation, M.S.Š. and A.Z.; writing—review and editing, M.S.Š. and A.Z.; visualization, M.S.Š. and A.Z.; supervision, S.Ž. and D.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author on reasonable request. The data are not publicly available due to privacy.

Acknowledgments

This work was supported by the Ministry of Science, Technological Development and Innovation of the Republic of Serbia through the support of scientific research work of the University of Niš Faculty of Civil Engineering and Architecture (Agreement registration number: 451-03-65/2024-03/200095).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Archibald, P.; Fanning, J.P.; Pavic, A. Interactive horizontal load model for pedestrians crossing footbridges. Bridge Struct.–Assesment Des. Constr. 2005, 1, 169–176. [Google Scholar] [CrossRef]
  2. Shahabpoor, E.; Pavic, A.; Racic, V.; Zivanovic, S. Effect of group walking traffic on dynamic properties of pedestrian structures. J. Sound Vib. 2017, 387, 207–225. [Google Scholar] [CrossRef]
  3. Feng, Y.; Chen, D.; Wang, Z.; Huang, S.; He, Y. Pedestrian-induced footbridge vibration response based on a simple beam analytical method. Adv. Bridge Eng. 2022, 3, 3. [Google Scholar] [CrossRef]
  4. Han, J.; Wang, H.; Liu, Y. Vibration analysis of a floor system based on probabilistic human-induced force model. Eng. Mech. 2014, 31, 81–87. [Google Scholar] [CrossRef]
  5. Dallard, P.; Fitzpatrick, A.J.; Flint, A.; Le Bourva, S.; Low, A.; Ridsdill-Smith, R.M.; Willford, M. The London Millennium Footbridge. Struct. Eng. 2001, 79, 17–33. [Google Scholar]
  6. Strogatz, S.H.; Abrams, D.M.; McRobie, A.; Eckhardt, B.; Ott, E. Crowded synchrony on the Millennium Bridge. Nature 2005, 438, 43–44. [Google Scholar] [CrossRef]
  7. Ingólfsson, E.; Georgakis, C.; Jönsson, J. Pedestrian-induced lateral vibrations of footbridges: A literature review. Eng. Struct. 2012, 45, 21–52. [Google Scholar] [CrossRef]
  8. Blekherman, A.N. Autoparametric Resonance in a Pedestrian Steel Arch Bridge: Solferino Bridge, Paris. J. Bridge Eng. 2007, 12, 669–676. [Google Scholar] [CrossRef]
  9. Ricciardelli, F.; Demartino, C. Design of footbridges against pedestrian-induced vibrations. J. Bridge Eng. 2016, 21, C4015003. [Google Scholar] [CrossRef]
  10. De Luca, A.; Bauer, M.; Pinto, M.; Malsch, E. Vibration analysis of footbridges: An overview of the current practice. In Proceedings of the 14th International Conference on Vibration Engineering and Technology of Machinery VETOMAC XIV, Lisbon, Portugal, 10–13 September 2018. [Google Scholar]
  11. EN 1991-2; Eurocode 1: Actions on Structures–Part 2: Traffic Loads on Bridges. European Committee for Standardization (CEN): Brussels, Belgium, 2003.
  12. EN 1990; Eurocode–Basis of Structural Design. European Committee for Standardization (CEN): Brussels, Belgium, 2005.
  13. AASHTO. LRFD Guide Specifications for the Design of Pedestrian Bridges, 2nd ed.; With 2015 Interim Revisions; American Association of State Highway and Transport Officials: Washington, DC, USA, 2009. [Google Scholar]
  14. Sétra. Footbridges-Assessment of Vibrational Behaviour of Footbridges under Pedestrian Loading. Technical Guide EQ-SETRA–06-ED17–FR+ENG; Service D’études Techniques des Routes et Autoroutes (Sétra): Paris, France, 2006. [Google Scholar]
  15. FIB. Guidelines for the Design of Footbridges. Bulletin 32; Fédération Internationale du Béton: Lausanne, Switzerland, 2005. [Google Scholar]
  16. Fairley, T.E.; Griffin, M.J. The apparent mass of the seated human body: Vertical vibration. J. Biomech. 1989, 22, 81–94. [Google Scholar] [CrossRef]
  17. Matsumoto, Y.; Griffin, M.J. Mathematical models for the apparent masses of standing subjects exposed to vertical whole-body vibration. J. Sound Vib. 2003, 260, 431–451. [Google Scholar] [CrossRef]
  18. Zheng, G.; Qiu, Y.; Griffin, M.J. Vertical and dual-axis vibration of the seated human body: Nonlinearity, cross-axis coupling, and associations between resonances in transmissibility and apparent mass. J. Sound Vib. 2012, 331, 5880–5894. [Google Scholar] [CrossRef]
  19. Sun, C.; Liu, C.; Zheng, X.; Wu, J.; Wang, Z.; Qiu, Y. An analytical model of seated human body exposed to combined fore-aft, lateral, and vertical vibration verified with experimental modal analysis. Mech. Syst. Signal Process. 2023, 200, 110527. [Google Scholar] [CrossRef]
  20. Fouli, M.; Camara, A. Human–structure interaction effects on lightweight footbridges with tuned mass dampers. Structures 2024, 62, 106263. [Google Scholar] [CrossRef]
  21. Colmenares, D.; Costa, G.; Civera, M.; Surace, C.; Karoumi, R. Quantification of the human–structure interaction effect through full-scale dynamic testing: The Folke Bernadotte Bridge. Structures 2023, 55, 2249–2265. [Google Scholar] [CrossRef]
  22. Gallegos-Calderón, C.; Naranjo-Pérez, J.; Díaz, I.M.; Goicolea, J.M. Identification of a Human-Structure Interaction Model on an Ultra-Lightweight FRP Footbridge. Appl. Sci. 2021, 11, 6654. [Google Scholar] [CrossRef]
  23. Zivanovic, S.; Pavic, A.; Reynolds, P. Vibration serviceability of footbridge under human-induced excitation: A literature review. J. Sound Vib. 2005, 279, 1–74. [Google Scholar] [CrossRef]
  24. Racic, V.; Pavic, A.; Brownjohn, J.M.W. Experimental identification and analytical modelling of human walking forces: Literature review. J. Sound Vib. 2009, 326, 1–49. [Google Scholar] [CrossRef]
  25. Shahabpoor, E.; Pavic, A.; Racic, V. Interaction between Walking Humans and Structures in Vertical Direction: A Literature Review. Shock Vib. 2016, 2016, 3430285. [Google Scholar] [CrossRef]
  26. Garinei, A. Vibrations of simple beam-like modelled bridge under harmonic moving loads. Int. J. Eng. Sci. 2006, 44, 778–787. [Google Scholar] [CrossRef]
  27. Stăncioiu, D.; James, S.; Ouyang, H.; Mottershead, J.E. Experimental model of a mass travelling on a continuous beam. J. Phys. Conf. Ser. 2009, 181, 012084. [Google Scholar] [CrossRef]
  28. Karimi, A.H.; Ziaei-Rad, S. Vibration analysis of a beam with moving support subjected to a moving mass travelling with constant and variable speed. Commun. Nonlinear Sci. Numer. Simul. 2015, 29, 372–390. [Google Scholar] [CrossRef]
  29. Mashaly, E.-S.; Ebrahim, T.M.; Abou-Elfath, H.; Ebrahim, O.A. Evaluating the vertical vibration response of footbridges using a response spectrum approach. Alex. Eng. J. 2013, 52, 419–424. [Google Scholar] [CrossRef]
  30. Heinemeyer, C.; Butz, C.; Keil, A.; Schlaich, M.; Goldbeck, A.; Trometor, S.; Lukic, M.; Chabrolin, B.; Lemaire, A.; Martin, P.-O.; et al. Design of Lightweight Footbridges for Human Induced Vibrations; Sedlacek, G., Heinemeyer, C., Butz, C., Geradin, M., Eds.; JRC Scientific and Technical Reports; European Commission: Luxembourg, 2009. [Google Scholar]
  31. Ellis, B.R.; Ji, T. Human-structure interaction in vertical vibration. Proc. Inst. Civ. Eng.-Struct. Build. 1997, 122, 1–9. [Google Scholar] [CrossRef]
  32. Maraveas, C.; Fasoulakis, Z.C.; Tsavdaridis, K.D. A Review of Human Induced Vibrations on Footbridges. Am. J. Eng. Appl. Sci. 2015, 8, 422–433. [Google Scholar] [CrossRef]
  33. Sachse, R. The Influence of Human Occupants on the Dynamic Properties of Slender Structures. Ph.D. Thesis, University of Sheffield, Sheffield, UK, 2002. [Google Scholar]
  34. Gao, Y.A.; Yang, Q.S. A theoretical treatment of crowd-structure interaction. Int. J. Struct. Stab. Dyn. 2018, 18, 1871001. [Google Scholar] [CrossRef]
  35. Busca, G.; Cappellini, A.; Manzoni, S.; Tarabini, M.; Vanali, M. Quantification of changes in modal parameters due presence of passive people on a slender structure. J. Sound Vib. 2014, 333, 5641–5652. [Google Scholar] [CrossRef]
  36. Yang, J.P.; Su, Z.; Yau, J.D.; Yang, D.S. Investigation of Spatial-Varying Frequencies Concerning Effects of Moving Mass on a Beam. Int. J. Struct. Stab. Dyn. 2023, 23, 2340002. [Google Scholar] [CrossRef]
  37. O’Sullivan, D.; Caprani, C.; Keogh, J. The Response of a Footbridge to Pedestrians Carrying Additional Mass. In Proceedings of the Bridge and Concrete Research Ireland (BCRI) Conference, Dublin, Ireland, 6–7 September 2012. [Google Scholar]
  38. Bergenudd, J.; Battini, J.-M.; Crocetti, R. Dynamic analysis of pedestrian timber truss bridge at three construction stages. Structures 2024, 59, 1005763. [Google Scholar] [CrossRef]
  39. Dallali, E.G.; Dey, P. Evaluation of the Equivalent Moving Force Model for Lightweight Aluminum Footbridges in Simulating the Bridge Response under a Single-Pedestrian Walking Load. Eng. Proc. 2023, 43, 22. [Google Scholar] [CrossRef]
  40. Tian, R.; Yang, X.; Zhang, Q.; Guo, X. Vibration reduction in beam bridge under moving loads using nonlinear smooth and discontinuous oscillator. Adv. Mech. Eng. 2016, 8, 1–12. [Google Scholar] [CrossRef]
  41. Li, J.; Liu, X. Human-Induced Vibration Analysis and Reduction Design for Super Long Span Pedestrian Arch Bridges with Tuned Mass Dampers. Appl. Sci. 2023, 13, 8263. [Google Scholar] [CrossRef]
  42. Manolis, G.D.; Dadoulis, G.I. Passive Control in a Continuous Beam under a Traveling Heavy Mass: Dynamic Response and Experimental Verification. Sensors 2024, 24, 573. [Google Scholar] [CrossRef] [PubMed]
  43. Spasojević-Šurdilović, M. Analiza graničnog stanja upotrebljivosti pešačkih mostova u pogledu vibracija indukovanih pešacima. Ph.D. Thesis, University of Niš, Niš, Serbia, 2013. [Google Scholar]
  44. Brčić, V. Dinamika Konstrukcija, 2nd ed.; Građevinska knjiga: Belgrade, Serbia, 1981. [Google Scholar]
  45. Blanchard, J.; Davies, B.L.; Smith, J.W. Design criteria and analysis for dynamic loading of footbridges. In Proceedings of the Symopsium on Dynamic Behaviour of Bridges at the Transport and Road Research Laboratory, Crowthorne, Berkshire, England, 19 May 1977. [Google Scholar]
  46. BS 5400-2; Steel, Concrete and Composite Bridges–Part 2: Specification for Loads. British Standards Institution (BSI): London, UK, 1978.
  47. SRPS EN 1991-2/NA:2019; Eurocode 1–Actions on Structures–Part 2: Traffic Loads on Bridges–National Annex. Institute for Standardization of Serbia (ISS): Belgrade, Serbia, 2019.
Figure 1. Dynamic model for transverse vibrations of a continuous beam system.
Figure 1. Dynamic model for transverse vibrations of a continuous beam system.
Applsci 14 07369 g001
Figure 2. Dynamic model of a continuous beam system with an additional mass µq.
Figure 2. Dynamic model of a continuous beam system with an additional mass µq.
Applsci 14 07369 g002
Figure 3. Variation of the fundamental frequency of a bridge with additional mass μq along length λ = 0.2l: (a) single-span beam L = l; (b) two-span beam L = 2l; (c) three-span beam L = 3l.
Figure 3. Variation of the fundamental frequency of a bridge with additional mass μq along length λ = 0.2l: (a) single-span beam L = l; (b) two-span beam L = 2l; (c) three-span beam L = 3l.
Applsci 14 07369 g003
Figure 4. Variation of the fundamental frequency of a bridge due to the arrival of a uniformly distributed mass μq: (a) single-span beam L = l; (b) two-span beam L = 2l; (c) three-span beam L = 3l.
Figure 4. Variation of the fundamental frequency of a bridge due to the arrival of a uniformly distributed mass μq: (a) single-span beam L = l; (b) two-span beam L = 2l; (c) three-span beam L = 3l.
Applsci 14 07369 g004
Figure 5. Variation of the fundamental frequency of a bridge with additional concentrated mass m: (a) single-span beam L = l; (b) two-span beam L = 2l; (c) three-span beam L = 3l.
Figure 5. Variation of the fundamental frequency of a bridge with additional concentrated mass m: (a) single-span beam L = l; (b) two-span beam L = 2l; (c) three-span beam L = 3l.
Applsci 14 07369 g005
Figure 6. Dynamic response of a timber footbridge with additional mass to a single resonant pedestrian: (a) deflection at mid-span; (b) acceleration at mid-span.
Figure 6. Dynamic response of a timber footbridge with additional mass to a single resonant pedestrian: (a) deflection at mid-span; (b) acceleration at mid-span.
Applsci 14 07369 g006
Figure 7. Dynamic response of a timber footbridge without additional mass to a single non-resonant pedestrian: (a) deflection at mid-span; (b) acceleration at mid-span.
Figure 7. Dynamic response of a timber footbridge without additional mass to a single non-resonant pedestrian: (a) deflection at mid-span; (b) acceleration at mid-span.
Applsci 14 07369 g007
Table 1. Frequency due to the influence of additional mass μq over length λ.
Table 1. Frequency due to the influence of additional mass μq over length λ.
Dynamic Model of the BridgeCircular Frequency
Applsci 14 07369 i001 ω ¯ r = ω r 1 + μ q μ · Φ a λ , x 0 1 2
Applsci 14 07369 i002 ω ¯ r = ω r 1 + μ q 2 μ · Φ a λ , x 0 1 2
Applsci 14 07369 i003 ω ¯ r = ω r 1 + μ q 3 μ · Φ a λ , x 0 1 2
Table 2. Frequency due to the influence of the arrival of uniformly distributed mass μq.
Table 2. Frequency due to the influence of the arrival of uniformly distributed mass μq.
Dynamic Model of the BridgeCircular Frequency
Applsci 14 07369 i004 ω ¯ r = ω r 1 + μ q μ Φ b x 0 1 2
Applsci 14 07369 i005 ω ¯ r = ω r 1 + μ q 2 μ Φ b x 0 1 2
Applsci 14 07369 i006 ω ¯ r = ω r 1 + μ q 3 μ Φ b x 0 1 2
Table 3. Frequency due to the influence of concentrated mass m.
Table 3. Frequency due to the influence of concentrated mass m.
Dynamic Model of the BridgeCircular Frequency
Applsci 14 07369 i007 ω ¯ r = ω r 1 + m 0.5 μ l Φ c x 0 1 2
Applsci 14 07369 i008 ω ¯ r = ω r 1 + m μ l Φ c x 0 1 2
Applsci 14 07369 i009 ω ¯ r = ω r 1 + m 1.5 μ l Φ c x 0 1 2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Spasojević Šurdilović, M.; Zorić, A.; Živković, S.; Turnić, D. Analytical Investigation of the Effects of Additional Load Mass on the Fundamental Frequency of Pedestrian Beam Bridges. Appl. Sci. 2024, 14, 7369. https://doi.org/10.3390/app14167369

AMA Style

Spasojević Šurdilović M, Zorić A, Živković S, Turnić D. Analytical Investigation of the Effects of Additional Load Mass on the Fundamental Frequency of Pedestrian Beam Bridges. Applied Sciences. 2024; 14(16):7369. https://doi.org/10.3390/app14167369

Chicago/Turabian Style

Spasojević Šurdilović, Marija, Andrija Zorić, Srđan Živković, and Dragana Turnić. 2024. "Analytical Investigation of the Effects of Additional Load Mass on the Fundamental Frequency of Pedestrian Beam Bridges" Applied Sciences 14, no. 16: 7369. https://doi.org/10.3390/app14167369

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop