Open access
Technical Papers
Nov 9, 2022

Using Full-Scale Observations on Footbridges to Estimate the Parameters Governing Human–Structure Interaction

Publication: Journal of Bridge Engineering
Volume 28, Issue 1

Abstract

The further development and improvement of prediction models for crowd-induced vibrations of footbridges requires detailed information on representative operational loading data. In this paper, an inverse method is used to estimate the parameters that govern human–structure interaction from the resulting structural response. The parameters of interest concern the dynamic characteristics of a mass-spring-damper (MSD) system, applied to describe the mechanical interaction between the pedestrian and the structure. The dynamic characteristics of the MSD interaction model are estimated by minimizing the discrepancy between the observed and the simulated power spectral density of the structural response. The parameter estimation procedure assumes that the dynamic behavior of the empty structure, the average weight, and the distribution of step frequencies in the crowd are known. The proposed approach is verified using numerical simulations and the influence of modeling errors is investigated. The results show that as footbridges and the human body are by nature lightly (2%) and strongly (≈30%) damped, respectively, the structural response is most sensitive to small variations in the natural frequency of the MSD interaction model. The results, furthermore, show that the parameter estimation problem is mostly sensitive to errors related to the mean value of the distribution of step frequencies and the structural modes’ natural frequency and modal mass. The impact of the structural modeling errors is found to decrease as the impact of human–structure interaction increases. Next, the approach is applied to two real footbridges where the walking behavior and the structural response induced by high pedestrian densities are observed. The results show that an estimate of the natural frequency (≈3.0 Hz) and damping ratio (34%) of the MSD interaction model is obtained that is in accordance with recent findings in the literature. These estimates are, however, for the first time ever, it is believed, based on full-scale observations involving high pedestrian densities.

Introduction

Footbridges are often conceived as slender and lightweight structures. In consequence, these structures can be highly sensitive to dynamic pedestrian excitation (Živanović et al. 2005). Where the static traffic load for footbridges is typically 5 kN/m2 (≈6 persons/m2) (NBN EN 1991-2 ANB 2013), a vibration serviceability assessment is generally performed, starting from 15 persons on the bridge deck, up to 1.5 persons/m2 (AFGC 2006; Heinemeyer et al. 2009). Dynamically speaking, pedestrian densities of 0.5 persons/m2, and beyond, can be considered high (AFGC 2006; Heinemeyer et al. 2009). Currently, the response of these structures to human-induced loading is predicted based on equivalent load models, upscaled from single-person force measurements (AFGC 2006; Heinemeyer et al. 2009). Although the dynamic performance under high crowd densities is often imperative for the design, these loading conditions have virtually never been verified (Georgakis and Ingólfsson 2014). Concerns about these load models are further strengthened by the fact that human–structure interaction (HSI) phenomena are not well understood (Bruno and Venuti 2009). The further development and calibration of pedestrian excitation models requires in-field observations, for they are the only way to obtain detailed and accurate information on representative operational loading data (Georgakis and Ingólfsson 2014; Živanović 2012). As direct force measurements are, in this case, practically infeasible, inverse methods, where the input or unknown model parameters are estimated from the resulting vibration response and a dynamic model of the structure, constitute a promising alternative.
HSI may refer to two different phenomena. The first phenomenon, also known as active HSI (Van Nimmen et al. 2021a) or structure-to-human (S2H) interaction (Ahmadi et al. 2018), refers to the case where a pedestrian changes walking behavior as a result of the (un)consciously perceived vibrations of the supporting structure. As, in the vertical direction, these active HSI phenomena arguably only occur for unacceptably high structural vibration levels, whereby pedestrians automatically stop walking (Dang and Živanović 2016), they are not taken into account in this study. The second phenomenon, also known as passive HSI (Van Nimmen et al. 2021a) or human-to-structure (H2S) interaction (Ahmadi et al. 2018), describes the mechanical interaction between the human body and the supporting structure. In the vertical direction, this phenomenon has been shown to have a significant impact on the dynamic behavior of the coupled crowd–structure system (Pimentel and Waldron 1996; Brownjohn and Fu 2005; Živanović et al. 2010; Dong et al. 2011; Salyards and Hua 2015; Shahabpoor et al. 2016a) and, in turn, on the vibration serviceability assessment of footbridges (Shahabpoor et al. 2017; Van Nimmen et al. 2017a, 2021a). This HSI phenomenon is focused on in the following.
In the literature, different approaches have been adopted to model and identify HSI for walking persons. Inspired by developments in biomechanics, multibody (Maca and Valasek 2011) and inverted pendulum (IP) (Bocian et al. 2013; Dang and Živanović 2013; Qin et al. 2013, 2014) models have been introduced to simulate the vertical pedestrian excitation and interaction. Experimental validation of these models is lacking and difficult, owing to the large number of model parameters involved.
Recent developments increasingly make use of a mass-spring-damper (MSD) model to describe vertical HSI (Van Nimmen et al. 2017a; Shahabpoor et al. 2016a; Caprani and Ahmadi 2016; Venuti et al. 2016; Zang et al. 2016). This model owes its popularity to its (1) simplicity and (2) ability to describe the low-frequency (0–10 Hz) dynamic behavior of the human body (ISO 1981; Brownjohn 2001; Matsumoto and Griffin 2003; Agu and Kasperski 2011; Caprani and Ahmadi 2016; Shahabpoor et al. 2016b; Cappellini et al. 2016; Van Nimmen et al. 2017a). This model is composed of a sprung mass mh1, an unsprung mass mh0, a spring kh1, and a damping element ch1. The nominal mass of the person is given by mh = mh1 + mh0 with μh1 = mh1/mh. Silva and Pimentel (2011) used the MSD model to define the relation between the amplitudes of the first three harmonics of the ground reaction forces (GRFs) induced by a pedestrian on a rigid surface and the acceleration levels registered on the pedestrian’s lower back. Acceleration data were collected for the lower backs of 33 persons walking on a rigid surface. The GRFs were simulated using a model given in the literature (Živanović et al. 2005). Considering a body mass of 70 kg and a pacing rate of 1.8 Hz, in that study, a natural frequency of 2.64 Hz and a damping ratio of 55% are reported for the MSD model.
Jimenez-Alonso and Saez (2014), Zang et al. (2016), and Zhang et al. (2015) reported on experiments performed on a 16-m-long double steel U-beam laboratory setup. The laboratory setup was excited by up to 10 pedestrians, together with a known shaker input force. Jimenez-Alonso and Saez (2014) fitted the MSD model parameters to minimize the mean square error between the experimentally identified and numerically predicted fundamental vertical natural frequency of the occupied footbridge. This approach resulted in a natural frequency of 2.75 Hz and a damping ratio of 47% for the MSD model. Zang et al. (2016) and Zhang et al. (2015) conclude that as the shaker frequency was closer to the natural frequency of the laboratory setup, the excitation of the pedestrians can be considered negligible. Using the known shaker input force, the frequency response function (FRF) of the primary system (footbridge) was derived from experiments. The parameters of the MSD model were then found by fitting the predicted FRF to the identified FRF. In this way, a natural frequency of 1.85 Hz and a damping ratio of 30% were found for the MSD model.
Shahabpoor et al. (2016a) used a laboratory setup consisting of an in situ cast post-tensioned concrete slab with a length of 11.2 mm and a width of 2 m. The experiments considered 2–15 pedestrians walking on the slab. The walking force of each person was set to be equal to the forces registered by an instrumented treadmill for the pedestrian walking on a rigid surface at a self-selected walking speed. In addition, the slab was excited by a known shaker input force. To identify the parameters of the MSD model, the error between the predicted and experimentally identified peak FRF magnitude and the modal parameters of the occupied structure (natural frequency, modal mass, and modal damping ratio) was minimized. In doing so, only a single mode of the concrete slab was taken into account. That study reports results obtained for the MSD model for natural frequencies in the range 2.75–3.00 Hz and damping ratios in the range 27.5%–0%.
So far, the calibration of the MSD interaction model involved laboratory-scale setups and a limited number of pedestrians. This work aims to estimate the dynamic characteristics of an MSD interaction model from full-scale observations involving high pedestrian densities. To this end, a parameter estimation procedure is proposed. The procedure is applied to two real footbridges, where the walking behavior and the structural response induced by high pedestrian densities are observed. The novelty and contribution of this work are in the formulation of the parameter estimation procedure specifically for the problem of HSI on footbridges, the analysis of its sensitivity to modeling errors, and the application to full-scale experimental observations.
The outline of this paper is as follows. In the next section, the parameter estimation procedure is presented and the impact of modeling errors is investigated. Then the procedure is applied to two real footbridges. The final section summarizes the conclusions.

Parameter Estimation Procedure

Mathematical Framework

This study uses the crowd–structure model introduced by Van Nimmen et al. (2017a). The dynamic behavior of the empty structure is represented by its modal parameters. Each pedestrian in the crowd is modeled as the superposition of an autonomous force term, corresponding to the forces induced on a rigid surface, and a linear MSD system used to simulate the vertical mechanical interaction with the supporting structure. This framework assumes that there is no active interaction between the pedestrians and the structure. In other words, it is assumed that the walking behavior (and thus, the autonomous force term) is not influenced by the structural vibrations perceived by the pedestrians.
The first-order continuous-time state-space equation of motion of the coupled crowd–structure model is
x˙(t)=Acx(t)+BcSp(t)pf(t)
(1)
withx(t)=[z(t)uh(t)z˙(t)u˙h(t)]
(2)
where x(t)Rens is the state vector; ns = 2 (nm + nh), z(t)Renm, is the modal coordinate vector; nm = number of modes; uhRenh is the vector of displacements of the MSD interaction models; nh = number of pedestrians; pf(t)Renh is the force vector collecting the time history of the autonomous force term pf,k(t) of each pedestrian k of the nh pedestrians as rows; and Sp(t)Rendof×nh is a selection matrix that transfers the forces to the corresponding ndof degrees of freedom (DOFs) of the model of the structure. The autonomous force term pf,k(t) is modeled using the probabilistic single-person force model developed by Živanović et al. (2007) and depends on the mass (mh,k) and the step frequency (fs,k) of the pedestrian. The arrival times of the pedestrians are assumed to follow a Poisson distribution (Helbing and Molnar 1995; Živanović 2012). The system matrices AcRens×ns and BcRens×ndof are defined as
Ac=[0IM¯hs1K¯hsM¯hs1C¯hs]
(3)
Bc=[0Tp]
(4)
where M¯hs, K¯hs, and C¯hs are the generalized mass, stiffness, and damping matrices of the coupled crowd–structure system, and Tp is the generalized input transformation matrix, as defined in Van Nimmen et al. (2017a).
The power spectral density (PSD) of the structural accelerations (Gu¨) can be derived from Monte Carlo simulations, given that the PSD is representative of the considered pedestrian density (see also the next section).
The mathematical framework is visualized in Fig. 1.
Fig. 1. Mathematical framework, parameter estimation problem, and input from experimental observations.

Parameter Estimation Problem

The objective is to identify the parameters that govern the effect of HSI on the resulting structural response: the dynamic characteristics of the linear mechanical systems representing the pedestrians in the crowd. The parameter estimation procedure assumes that the following inputs are known.
1.
The dynamic behavior of the empty structure (the structure without pedestrians). The natural frequencies and mode shapes can be predicted using a finite-element (FE) model and the corresponding modal damping ratios estimated based on the used materials and the construction type (AFGC 2006). A much more accurate representation of the dynamic behavior of the structure can be obtained from experimental vibration data using system identification techniques (Peeters and De Roeck 1999; Reynders and De Roeck 2008).
2.
The structural response induced by the pedestrians. This requires vibration measurements on the structure during operation. The duration of the measurements should be sufficiently long that the collected data can be considered representative of the involved pedestrian density. This condition can be verified by checking that the PSD of the structural accelerations has converged.
3.
The average weight of the pedestrians. This information can be easily collected for each participant using a scale on site.
4.
The distribution of step frequencies in the crowd. This can be derived from empirical relations (e.g., as a function of the pedestrian density Bruno and Venuti 2009) but is preferably based on additional information collected in situ (e.g., the body motion registered for each pedestrian) (Van Nimmen et al. 2021b).
Van Nimmen et al. (2017b) show that when a sufficiently long duration of the structural response induced by a certain pedestrian density is considered, the statistical effect of the remaining unknown stochastic parameters, i.e., the effect of the randomness in the autonomous term, on the PSD of the resulting structural response can be averaged out. The randomness of the autonomous force term results from the arrival times and the intraperson variabilities. As these parameters are much more difficult to identify reliably from experimental observations, it is a great advantage to be able to average out their statistical effect.
Next, an optimization problem is formulated, where the objective function measures the difference between the PSD of the experimentally observed (G~u¨) and numerically simulated (Gu¨) structural responses within a frequency range defined by a lower (ωl [rad/s]) and upper bound (ωu [rad/s]). For now, it is assumed that the structural response is observed in a single output location. The residuals are computed at discrete frequencies ωk [rad/s]: ωk=ωl+kΔωfork=0,,N, where N + 1 = (ωuωl)/Δω + 1 = number of samples; and Δω = selected frequency resolution. The decision variables in this optimization problem are the natural frequency fh1, the modal damping ratio ξh1, and the mass ratio μh1 of the selected linear mechanical system (see the introduction). It is shown by Van Nimmen et al. (2017a) that the impact of the interperson variability of the parameters of the human body models on the resulting structural response is low. Therefore, these parameters are assumed to be identical for the different persons on the structure and are subject to the following physical constraints: fh1 > 0, ξh1 > 0, and 0 < μh1 < 1. The problem, subject to four inequality constraints, is formulated in standard form:
minimizefh1,ξh1,μh1g(fh1,ξh1,μh1)=G~u¨Gu¨(fh1,ξh1,μh1)22G~u¨22subject tofh1>0ξh1>0μh1>0μh1<1
(5)
and can be categorized as a nonlinear programming problem, owing to the nature of the objective function (Nocedal and Wright 2006). The inequality constrained optimization problem can be solved using the lsqnonlin solver (MATLAB 2018) by means of the trust-region-reflective algorithm.
In the case where the structural response is registered at mdof DOFs, the many objectives can be combined into a single-objective scalar function as follows:
minimizefh1,ξh1,μh1l=1mdofγlgl(fh1,ξh1,μh1)withl=1mdofγl=1γl>0l=1,,mdof
(6)
where γl is the weight factor for the structural response registered at DOF l. If desired, the objective function in Eq. (6) can be further extended with a frequency weight factor γω to increase or decrease the contribution of certain frequency ranges.
Given the low variation that is reported in the literature for the mass ratio μh1, it is reasonable to consider this parameter fixed and equal to μh1 = 0.95 (Matsumoto and Griffin 2003; Shahabpoor et al. 2016b). In this way, the numbers of parameters and inequality constraints in the optimization problem in Eq. (5) are reduced to two (fh1, ξh1) and two (fh1 > 0, ξh1 > 0), respectively.
Fig. 1 visualizes the applied mathematical framework, the parameter estimation problem (objective function), and the input from the experimental observations.

Numerical Verification Example

The parameter estimation procedure is illustrated and verified using a numerical example. The structure under consideration is a footbridge with a span L of 50 m and a deck width of 3 m. The structure is modeled as a simply supported Euler–Bernoulli beam for which only the fundamental vertical bending mode is considered, with realistic values of its modal parameters (AFGC 2006; Heinemeyer et al. 2009): a sinusoidal mode shape; a modal mass of 50 × 103 kg; a modal damping ratio of 0.5%; and a natural frequency of 2 Hz. In this work, the modal mass refers to the proportion of the physical mass of the structure that is engaged in that mode (Brownjohn and Pavić 2007). It is noted that this is different from the unit-value modal mass that is conventionally related to mass-normalized mode shapes (Brownjohn and Pavić 2007). The lower the modal mass, the more sensitive that mode is to dynamic excitation.
A pedestrian density of 0.5 pedestrians/m2 is considered, with a corresponding walking speed of 1.30 m/s, a mean step frequency of 1.89 Hz (Bruno and Venuti 2009), and a standard deviation of the step frequencies of 0.175 Hz, as for the case of spatially unrestricted traffic (AFGC 2006) (N(1.89,0.175)). The phase shifts between the different pedestrians are uniformly distributed in [0, 2π] (Van Nimmen et al. 2020) and their arrival times are assumed to follow a Poisson distribution (Helbing and Molnar 1995; Živanović 2012). Fig. 2 presents the PSD of the pedestrian load. The latter is determined based on the procedure detailed in Van Nimmen et al. (2020), using the dynamic load factors for the main and subharmonics of the walking load defined in Živanović et al. (2007). The mechanical properties of the MSD interaction model are set to approximate the dynamic behavior of a person with one or two legs slightly bent [fh1 = 3.0 Hz; ξh1 = 0.30; mh1 = 70 kg; and μh1 = 0.95].
Fig. 2. PSD of pedestrian load (GF) for pedestrian density of 0.50 persons/m2.
Fig. 3 presents the PSD of the structural response at midspan, predicted without and with HSI. This figure illustrates that, in both cases, the structural response is dominated by a single peak. For the case without HSI, this peak is centered around the natural frequency of the footbridge (2 Hz). When HSI is accounted for, the peak is centered around the effective natural frequency of the coupled crowd–structure system (Van Nimmen et al. 2017a), which is, in this case, located at 1.93 Hz.
Fig. 3. (a) PSD of structural response at midspan induced by pedestrian density d = 0.5 pedestrians/m2; and (b) enlargement of the dominant frequency range. Black = without HSI; gray = with HSI.

Optimal Solution

In this example, the PSD of the structural response at midspan (frequency range 0–10 Hz) is considered as the input for the objective function [Eq. (5)]. The objective function is now evaluated for a realistic range of the two decision variables (Fig. 4): the natural frequency 1 ≤ fh1 ≤ 6 Hz and the modal damping ratio 0.1 ≤ ξh1 ≤ 0.8. This figure shows that the objective function, and thus the impact of HSI, is mainly sensitive to changes in the natural frequency fh1 of the MSD interaction model. Fig. 4 illustrates a generally low sensitivity of the objective function to the modal damping ratio ξh1 of the MSD interaction model. The sensitivity to ξh1 (slightly) increases for increasing values of fh1.
Fig. 4. Evaluation of objective function in terms of natural frequency fh1 and modal damping ratio ξh1 of MSD interaction model. Solid circle = solution of optimization procedure.
For different feasible sets of initial values, the optimization procedure is found to successfully converge toward the true minimum: fh1 = 3.0 Hz and ξh1 = 0.3.

Influence of Modeling Errors

To gain insight into the influence of the modeling errors on the solution of the parameter estimation problem, a parameter study was conducted to consider a reasonable range of footbridge parameters and pedestrian densities.
1.
Footbridge parameters. The previously introduced structure of a simply supported Euler–Bernoulli beam with a span of 50 m and a width of 3 m was used. Only the fundamental vertical bending mode with a sinusoidal mode shape was considered. The following ranges of modal parameters were considered: 1 ≤ f1 ≤ 6 Hz in steps of 0.5 Hz, and two values of the structural damping ratio ξ1 = [0.5, 2.0]%.
2.
Pedestrian parameters. Instead of considering different pedestrian densities, three crowd-to-structure mass ratios (μm) were considered: μm=[2.5,5.0,10.0]%. The crowd-to-structure mass ratio was defined as the ratio between the modal mass added by the crowd as a mass per unit area (m1,add) [see also Van Nimmen et al. (2021a)] and the modal mass of the empty structure (m1).
In addition to the physical constraints formulated in Eq. (5), upper bounds of 10 Hz and 100% were set for the natural frequency and the damping ratio of the MSD interaction model, respectively.
The modeling errors are related to the inputs that the parameter estimation procedure assumes to be known (Section “Parameter Estimation Procedure”). Given the high accuracy with which the structural response can be measured, and the ease with which the weight of the pedestrians can be determined, the impact of these related errors are not investigated here. The modeling errors (±) considered in this study are related to the following.
1.
The dynamic behavior of the empty structure characterized by its modal parameters. The parameter estimation procedure presented in this paper assumes that the dynamic behavior of the empty structure is known. When this procedure is applied to full-scale data involving pedestrian-induced vibrations, it is common practice that an operational or experimental modal analysis is performed on the empty footbridge. Given the uncertainty bounds on the modal parameters identified using state-of-the-art system identification techniques (Peeters and De Roeck 1999; Reynders and De Roeck 2008), it is reasonable to assume errors of 1% and 10% on the experimentally identified natural frequencies and modal damping ratios, respectively. The corresponding modal masses can be calculated using a (preferably calibrated) FE model of the structure. For the calculated modal mass, an error of 2% is considered reasonable. Less satisfactory results may be characterized by errors of 5%, 30%, and 5% on the natural frequency, modal damping ratio, and modal mass, respectively.
2.
The distribution of step frequencies in the crowd. Assuming that the distribution of step frequencies follows a Gaussian distribution, the impact of an error of 1% and 2.5% on the mean value and standard deviation of the step frequencies was investigated.
As the impact of modeling errors on the parameter estimation problem will always be case-specific, the objective of the analysis presented here is not to provide an exact quantification of the resulting estimation errors. The objective is to qualitatively assess in which conditions (i.e., mass ratio and modal parameters of a structural mode) it is feasible to get a reliable estimate of the MSD interaction model parameters, given the reasonably expected modeling errors.
Fig. 5 shows the ratio of the estimated and true values of the MSD interaction model parameters resulting from modeling errors on the structural modal parameters. When these ratios are close to unity, the impact of the modeling errors is low, and vice versa. When analyzing these results, it is important to keep in mind some key results from previous research involving the impact of HSI (Van Nimmen et al. 2017a). The impact of HSI on a structural vibration mode is mainly reflected in a (considerable) increase of the effective damping ratio of the coupled crowd–structure system in relation to the modal damping ratio of the empty structure. The impact of HSI increases as: (1) the frequency of the MSD interaction model is closer to the optimal tuned mass damper (TMD) frequency for accelerations (Asami et al. 2002) (i.e., for fh1 smaller than but close to f1, in the following expressed as fh1f1); (2) the crowd-to-structure mass ratio increases; and (3) the damping ratio of the footbridge decreases. The impact of HSI is most sensitive to the frequency ratio fh1/f1, where the largest impact of HSI is found for f1 larger than but close to fh1, in the following expressed as f13 Hz with fh1 = 3 Hz. The following observations can now be made from Fig. 5, for the considered range of modeling errors:
1.
The impact of the structural modeling errors decreases as the impact of HSI increases, thus as: (1) f13 Hz; (2) the crowd-to-structure mass ratio increases; and (3) the damping ratio of the footbridge decreases.
2.
The accuracy of the estimates of the natural frequency and the modal damping ratio generally go hand in hand: if the estimate of the natural frequency is close to its true value, so is the estimate of the damping ratio.
3.
As the natural frequency of the structural mode decreases from 2.5 Hz onward, the impact of the modeling errors increases rapidly, in particular for low mass ratios. This is because the impact of HSI decreases rapidly as f1 decreases below 2.5 Hz. Conversely, as the natural frequency of the structural mode (f1) increases from 3 Hz onward, the impact of these modeling errors increases slowly as the impact of HSI decreases slowly with increasing f1.
Fig. 5. Impact of error on modal mass (black, εm = 2%; gray, εm = 5%); damping ratio (black, ϵξ=10%; gray, ϵξ=30%) and natural frequency (black, εf = 1%; gray, εf = 2%) of footbridge, on ratio of estimated and true value of natural frequency (f^h1/fh1) and damping ratio (ξ^h1/ξh1) of the MSD interaction model, in terms of natural frequency of footbridge (f1), for three damping ratios of the footbridge (open circle, ξ1 = 0.5%; multiplication sign, ξ1 = 1.0%; plus sign, ξ1 = 5%) and three mass ratios (μm={2.5%,5.0%,10.0%}). Solid circle: constraint is active for optimal solution.
In practical terms, it is advisable to minimize the influence of structural modes with a natural frequency below ≈2 Hz in the parameter estimation problem, because of the adverse effect on the accuracy of the obtained results, even for small structural modeling errors. This can, for example, be enforced by only considering the frequency range >2 Hz in the objective function. For modes with a natural frequency of 3 Hz and higher, the impact of the structural modeling errors increases as the natural frequency increases. This is particularly relevant for modeling errors related to the natural frequency. More generally, the higher the mass ratio, the more robust the obtained result is against structural modeling errors.
Fig. 6 shows the ratio of the estimated and true values of the MSD interaction model parameters resulting from modeling errors on the mean value and the standard deviation of the distribution of step frequencies. The following observations are made for the considered range of modeling errors.
1.
The impact of the modeling errors decreases for increasing mass ratio.
2.
The results are mainly sensitive to errors on the mean value of the distribution of step frequencies. The errors introduced by the standard deviation of the step frequencies stay (well) below 10% (except for the case where the mass ratio is lower than 5%).
3.
The impact of the error on the mean value of the distribution of step frequencies is as follows:
a.
On the estimated natural frequency of the MSD interaction model, for structural modes with a frequency equal to or higher than 2.5 Hz and μm5%, the estimation error is limited to 15%. For structural modes with a frequency lower than 2.5 Hz, the estimation error can increase up to 200%.
b.
On the estimated damping ratio of the MSD interaction model, for structural modes with a frequency equal to or higher than 2.5 Hz, the estimation error is limited to 50% (μm=5%) and 25% (μm=10%). For structural modes with a frequency close to 1.5 Hz (and 4.5 Hz for μm=2.5%), the estimation error can increase up to 200% or more.
Fig. 6. Impact of error on mean value (black, ϵμfs=1%; gray, ϵμfs=2.5%) and standard deviation (black, ϵσfs=1%; gray, ϵσfs=2.5%) of distribution of step frequencies on ratio of estimated and true value of natural frequency (f^h1/fh1) and damping ratio (ξ^h1/ξh1) of MSD interaction model, in terms of natural frequency of footbridge (f1), for three damping ratios of footbridge (open circle, ξ1 = 0.5%; multiplication sign, ξ1 = 1.0%; plus sign, ξ1 = 5%) and three mass ratios (μm={2.5%,5.0%,10.0%}). Solid circle: constraint is active for optimal solution.
For the considered range of modeling errors, it is shown that the results of the parameter estimation problem are mostly sensitive to errors related to the structural natural frequency and the mean value of the step frequencies. Furthermore, the impact of the structural modeling errors decreases as the impact of HSI increases. In other words, the impact of these modeling errors decreases as the sensitivity of the objective function to the decision variables increases.
The conclusions made here and in the numerical verification example (Section “Parameter Estimation Problem”), are case-specific and depend primarily on the dynamic behavior of the empty structure and the pedestrian density. For example, it can be expected that when a number of structural modes have a relevant contribution to the structural response, the optimization problem will be more robust against modeling errors: the insensitivity of a subset of structural modes to a decision variable can be countered by the sensitivity of another subset of structural modes. This is under the assumption that the structural modal parameters are estimated with a reasonable level of uncertainty and the natural frequencies are well spread over the interval 1–6 Hz. It is therefore advised for each case study (1) to investigate the duration of the experimental observations required to arrive at a reasonable level of convergence of the PSD of the structural response; and (2) to properly assess the impact of modeling errors at hand on the obtained optimal solution. Based on the insights gained in the latter, a greater weight can be assigned in the objective function [Eq. (6)] to specific frequency ranges and DOFs, such that the impact of modeling errors is minimized.

Case Study: Eeklo Footbridge

The case study and experimental data used in this section are fully documented in the Eeklo footbridge open access benchmark dataset (Van Nimmen et al. 2021b).

Structure

The Eeklo footbridge has three spans, a main central span of 42 m and two side spans of 27 m. The dynamic behavior of the footbridge is characterized by 14 modes, with a frequency up to 12 Hz (Fig. 7). An excellent agreement is found between the experimentally identified modal characteristics and the ones calculated by the calibrated FE model. For more information related to the footbridge and the model calibration, the reader is referred to Van Nimmen et al. (2014). The corresponding digital twin, describing the dynamic behavior of the footbridge, is part of the open access benchmark dataset (Van Nimmen et al. 2021b).
Fig. 7. Top and side views of first eight modes of Eeklo footbridge (compare with digital twin, calculated using ANSYS).

Full-Scale Observations

The bridge and pedestrian motion were registered simultaneously using wireless triaxial accelerometers on the deck and at the lower back of the pedestrians. Four data blocks were collected involving two pedestrian densities, 0.25 and 0.50 persons/m2, representing a total of more than one hour’s worth of data for each pedestrian density. Van Nimmen et al. (2021b) show that it takes 1,000 s to limit the convergence error of the PSD of the structural response to less than 10%. These results indicate that the dataset collected for each pedestrian density, with a total length of more than 3,600 s, can be considered representative of the involved load case. The converged PSDs are shown in Fig. 8 for a pedestrian density of 0.50 persons/m2. The PSD shows that the vertical structural accelerations are dominated by the contributions of the first (1.69 Hz) and second (2.97 Hz) modes of the footbridge. At the side spans, an important contribution of Modes 5 (5.68 Hz) and 8 (6.42 Hz) is also observed: two lightly damped vertical bending modes that have an antinode at the center of the side span. Similar observations are made for a pedestrian density of 0.25 persons/m2.
Fig. 8. PSD of (black) vertical and (gray) lateral acceleration levels at center of (a) central; and (b) side span of Eeklo footbridge for pedestrian density of 0.50 persons/m2.
Using the procedure detailed in Van Nimmen et al. (2018b), the step frequency of all pedestrians is reported in Van Nimmen et al. (2021b). The results show that the distribution of step frequencies can be well fitted by a Gaussian distribution. For pedestrian densities of 0.25 and 0.50 persons/m2, the distribution of step frequencies is characterized by N(1.77,0.13) Hz (χ2 test, 94%) and N(1.66,0.19) Hz (χ2 test, 87%), respectively.
For the tests involving a pedestrian density of 0.25 and 0.50 persons/m2, the average body mass is 74.9 and 73.2 kg, respectively. The corresponding crowd-to-structure (modal) mass ratios (μm) are given in Table 1.
Table 1. Crowd-to-structure (modal) mass ratios (μj,m %) for Mode j of Eeklo footbridge for pedestrian density of 0.25 and 0.50 persons/m2
 μj,m (%)
Mode0.25 persons/m20.50 persons/m2
Mode 10.30.7
Mode 25.310.5
Mode 31.32.7
Mode 40.20.4
Mode 55.611.2
Mode 61.63.3
Mode 71.22.4
Mode 85.711.4
Mode 91.22.4
Mode 100.91.7
Mode 116.513.0
Mode 120.51.1
Mode 131.12.2
Mode 140.30.6

Parameter Estimation

In a preliminary step, the digital twin of the Eeklo footbridge and an MSD interaction model with parameters corresponding to recent findings in the literature (natural frequency fh1 = 3 Hz and modal damping ratio ξh1=30%), were used to get a first idea of the impact of HSI for the Eeklo footbridge. Fig. 9 shows the FRF of the Eeklo footbridge, empty and with HSI for a pedestrian density of 0.50 persons/m2. This figure shows that the impact of HSI is high for Modes 2 (f2 = 2.97 Hz), 5 (f5 = 5.68 Hz), and 8 (f8 = 6.42 Hz). These modes are bending modes characterized by a low modal mass (and thus high crowd-to-structure modal mass ratio: ≈10% for a pedestrian density of 0.50 persons/m2) and a low damping ratio (0.2%–0.6%). The highest impact is found for Mode 2, as the natural frequency of the MSD interaction model is, in this case, close to the optimal TMD tuning frequency for accelerations (Asami et al. 2002). Fig. 9 furthermore shows that HSI has a negligible impact on the lateral structural response. The latter results from using a vertical MSD interaction model, as well as from the fact that the modes for which a high impact of HSI is found do not display significant lateral modal displacements.
Fig. 9. FRF for (black) vertical and (gray) lateral input and output acceleration at center of (a) central; and (b) side span of Eeklo footbridge. Dashed line = empty; solid line = with HSI, pedestrian density = 0.50 persons/m2.
Next, the inputs for the parameter estimation procedure were set to match the conditions of the in-field observations involving pedestrian densities of 0.25 and 0.50 persons/m2, respectively. The average body masses of the pedestrians were 74.9 and 73.2 kg, respectively. The distributions of step frequencies in the crowd were fsN(1.77,0.13) Hz and fs=N(1.66,0.19)Hz, respectively. The digital twin of the Eeklo footbridge was used to describe the dynamic behavior of the empty footbridge. The structural response measured at the center of the main and the side span was used (close to the parapet). As the preliminary results show that HSI is most relevant for the vertical output, only the acceleration levels registered in the vertical direction were used in the parameter estimation problem. The PSD of the structural response was calculated, considering consecutive time windows of length T = 25 s, with 50% overlap. Given that the highest impact of HSI is expected for Modes 2, 5, and 8, the frequency range [2.5, 8.0] Hz is taken into account for the parameter estimation problem. Following the results of Section “Influence of Modeling Errors,” the high impact of HSI on the selected frequency range will limit the impact of modeling errors.
In Fig. 10, the objective function defined in Eq. (5) is evaluated for a realistic range of the two decision variables characterizing the MSD system: the natural frequency 1 ≤ fh1 ≤ 6 Hz and the modal damping ratio 0.15 ≤ ξh1 ≤ 0.80. This figure shows that the objective function, and thus the impact of HSI, is mainly sensitive to changes in the natural frequency fh1 of the MSD interaction model. These observations agree with the findings of the numerical verification example.
Fig. 10. Evaluation of objective function in terms of natural frequency fh1 and damping ratio ξh1 of MSD interaction model, and (dashed lines) indicative uncertainty bounds on (solid circle) optimal solution, for pedestrian density of (a) 0.25 persons/m2; and (b) 0.50 persons/m2.
For a pedestrian density of 0.25 persons/m2 (0.50 persons/m2), an optimal solution of fh1=2.87 Hz, ξh1=36% (fh1=3.08 Hz, ξh1=34%) is found.

Influence of Modeling Errors

In a last step, the impact of a reasonable range of modeling errors on the results of the parameter estimation problem was investigated. Given the high accuracy with which the structural response can be measured, and the ease with which the total mass of the pedestrians can be determined, the impact of related errors was not investigated. Also, it is assumed that the impact of (correlated) measurement noise, e.g., related to the sensor placement, is negligible. The modeling errors considered in this study are related to the following.
1.
The dynamic behavior of the empty structure; thus, the structural modal parameters. The natural frequencies and modal damping were identified using state-of-the-art system identification techniques (Peeters and De Roeck 1999; Reynders and De Roeck 2008). Errors of ϵf=1% and ϵξ=10% on the experimentally identified natural frequencies and modal damping ratios, respectively, were considered reasonable (Section “Influence of Modeling Errors”). The corresponding modal masses were calculated using the calibrated FE model of the structure. A corresponding error of ϵm=2% is considered reasonable. Less satisfactory results may be characterized by errors of 5%, 30%, and 5% (upper bound error) on the natural frequencies, modal damping ratios, and modal masses, respectively.
2.
The distribution of step frequencies in the crowd. Assuming that the distribution of step frequencies follows a Gaussian distribution, the impacts of errors of 2% and 5% on the mean value (ϵμfs) and standard deviation (ϵσfs) of the step frequencies was investigated.
This study does not consider the impact of other modeling errors, such as lateral HSI, which is neglected here, or errors in the load model used to describe the autonomous force term. Fig. 11 presents the individual and joint impacts of the modeling errors on the solution of the parameter estimation problem (estimation error). The estimation error (Δ) is expressed as a percentage of the optimal solution of the parameter estimation problem given in Section “Parameter Estimation.” The following observations are made.
1.
The natural frequency of the MSD interaction model fh1 is most sensitive to an error in the mean value of the distribution of step frequencies. This corresponds to the observations made in Section “Influence of Modeling Errors”: owing to the high impact of HSI for the Eeklo footbridge, the impact of (small) structural modeling errors is very low.
For a pedestrian density of 0.25 persons/m2 [Fig. 11(a)], an estimation error of 4% (8%) is found for a modeling error of 2% (5%) on the mean value of the step frequencies. The joint impact of the modeling errors is limited to an estimation error of 10%.
For a pedestrian density of 0.50 persons/m2 [Fig. 11(c)], an estimation error of 1.5% (4%) is found for a modeling error of 2% (5%) on the mean value of the step frequencies. The joint impact of the modeling errors is limited to an estimation error of 5%.
2.
Figs. 11(b and d) show that for a pedestrian density of 0.25 persons/m2 (0.50 persons/m2), the damping ratio of the MSD interaction model fh1 is most sensitive to an error in the mean value of the step frequencies (an error in the modal damping ratios of the structure). The joint impact of the modeling errors is limited to an estimation error of 25% (0.25 persons/m2) and 20% (0.50 persons/m2).
Fig. 11. Estimation error (Δ) of (a and c) natural frequency; and (b and d) damping ratio of MSD interaction model due to modeling errors: εf (1% black, 5% gray); εξ (10% black, 30% gray); εm (2% black, 5% gray); ϵμfs=ϵσfs (2% black, 5% gray). Pedestrian density = (a and b) 0.25 persons/m2; and (c and d) 0.50 persons/m2.
These results show that the reasonable range of modeling errors that can be expected for this case have a small impact on the estimated natural frequency of the MSD interaction model (5%–10%). The impact on the estimated damping ratio is higher (20%–25%), but is still reasonable, given that the structural response is less sensitive to changes in this parameter. These uncertainty bounds are also indicated in Fig. 10.
It is noted that the low impact of the modeling errors, in particular those related to the modal parameters of the footbridge, on the estimated parameters is, in this case, due to the high impact of HSI on the higher modes of the footbridge (Modes 2, 5, and 8). This is, in turn, due to the involved low modal masses, low damping ratios, and natural frequencies of these modes.

Results

The nominal results of Section “Parameter Estimation” are combined with the uncertainty bounds determined in Section “Influence of Modeling Errors”:
1.
0.25 persons/m2: fh1=2.87±0.29 Hz; ξh1=36±9%
2.
0.50 persons/m2: fh1=3.08±0.15 Hz; ξh1=34±7%.
Fig. 12 finally compares the PSD of the measured output with that calculated for the footbridge with and without HSI, whereby HSI is modeled with the fitted MSD interaction model (nominal values) for a pedestrian density of 0.50 persons/m2. This figure shows that the fitted MSD interaction model allows a good fit with the measurements to be obtained. The residual is well below 1% of the error when HSI is disregarded. Similar observations are made for a pedestrian density of 0.25 persons/m2.
Fig. 12. PSD of vertical acceleration levels at center of (a) central; and (b) side span of Eeklo footbridge for pedestrian density of 0.50 persons/m2. (black, solid) measurements, and numerically simulated (gray, solid) without HSI and (black, dashed) with HSI, based on fitted MSD interaction model.
Furthermore, it is noted that the identified natural frequency (fh1 ≈ 3.0 Hz) and damping ratio (ξh135%) of the MSD interaction model are in agreement with recent findings in the literature (Shahabpoor et al. 2016b; Van Nimmen et al. 2017a)

Case Study: Waalse Krook South Footbridge

The case study and experimental data used in this section were first introduced by Van Nimmen et al. (2018a).

Structure

The Waalse Krook (WK) South footbridge is a steel pedestrian bridge with a single span of 25 m and a width of 6.9 m (Fig. 13). The bridge deck is composed of a two-way ribbed steel–concrete composite plate. The bridge deck is supported by two neoprene bearings, on one side, and two steel columns, on the other side. An operational modal analysis was performed for the empty footbridge (Van Nimmen et al. 2018a). Two modes with a frequency up to 5.5 Hz were identified (see Fig. 14): a vertical bending mode and a torsional mode. A detailed FE model was developed according to the architectural drawings using the FE software ANSYS. By tuning the translational stiffness of the neoprene supports, an optimal agreement was obtained with the experimentally identified natural frequencies and mode shapes: for both modes, the relative frequency deviation is limited to 2% and the corresponding modal assurance criterion (MAC) values are high (>0.95).
Fig. 13. WK South footbridge: (a) photograph (image by authors); and (b) cross section.
Fig. 14. Operational modal parameters of two modes identified for empty footbridge: (a) 3.30 Hz, 0.86%; and (b) 4.14 Hz, 0.25%.
(Reprinted with permission from Van Nimmen et al. 2018a.)
As reported in the following, numerical simulations were performed based on the experimentally identified natural frequencies and modal damping ratios, and the mass-normalized mode shapes of the calibrated FE model. All modes with a natural frequency below 20 Hz are taken into account. Unless stated otherwise, the simulated and measured structural responses were both low-pass filtered with a cut-off frequency of 10 Hz.

Full-Scale Observations

A full-scale measurement campaign was carried out, considering a pedestrian density of 1 person/m2 (free walking). The average weight of the pedestrians is 72 kg. This corresponds to a crowd-to-structure (modal) mass ratio (μm) of 5.8% and 5.7% for Modes 1 and 2, respectively. During these tests, the vertical structural response was registered at the locations corresponding to the antinodes of the first and second modes of the footbridge (Fig. 14).
A time frame of 8 min, reflecting the stationary response of the structure, was selected. The structural response was low-pass filtered with a cut-off frequency of 10 Hz. The PSD of the structural response was calculated, considering nT = 8 consecutive time windows of length T = 60 s, with 50% overlap. As shown in Fig. 15, the convergence of the PSD was first analyzed by evaluation for an increasing number kT of consecutive time windows of length T. The convergence error (ΔGu¨,kT) is defined as the discrepancy between the PSD calculated considering kT (Gu¨,kT) and the total of nT = 8 (Gu¨,nT) consecutive time windows. Fig. 15 shows that the discrepancy between the PSDs calculated for nT = 7 and nT = 8 is approximately 10%. Based on this observation, the remaining convergence error on the PSD considering the full, yet relative short, available time window is estimated at 10%.
Fig. 15. (a) PSD of structural response induced by pedestrian density d = 1.0 pedestrians/m2 measured at antinode of first mode of WK footbridge, calculated for increasing number kT of consecutive time windows of length T = 60 s, for (palest gray) kT = 1 to (black) kT = 8; and (b) corresponding convergence error ΔG~u¨.
(Reprinted with permission from Van Nimmen et al. 2018a.)
The dominant contributions around 3.3 and 4.1 Hz in Fig. 16 show that the pedestrians mainly excite the fundamental and second modes of the footbridge. In addition, the structural response is characterized by clear contributions around 1.8 Hz. These contributions reflect the nonresonant excitation by the fundamental harmonic of the walking load. The aim next was to use this observation to derive information on the distribution of step frequencies in the crowd. To do this, the structural response was first low-pass filtered with a cut-off frequency of 2.3 Hz, such that the influence of the dynamic modes of the structure was minimized. Fig. 16 shows that, in this way, the contribution of the fundamental harmonic of the walking load to the PSD of the structural response is isolated. Furthermore, this figure shows that this part of the PSD was barely influenced by the dynamic contribution of the modes of the footbridge. Second, it is assumed that this contribution is negligibly influenced by HSI, an assumption that is reasonable, based on the observations made by Van Nimmen et al. (2017a). Based on this assumption and by minimizing the difference between the PSD of the measured and simulated structural responses, low-pass-filtered at 2.3 Hz, the following distribution of step frequencies in the crowd is fitted: N(1.68,0.13) Hz. At this point, simulations were performed without HSI. Fig. 17 shows that, based on this fitted distribution of step frequencies, a good agreement is found between the PSD of the measured and simulated structural responses for frequencies below 2.3 Hz. Fig. 17 also shows, for frequencies higher than 2.3 Hz, and in particular for frequencies close to the fundamental and second natural frequencies of the footbridge, significant differences between the PSDs of the measured and simulated structural responses. It is reasonable to assume that these differences are largely due to HSI, which was disregarded in these simulations.
Fig. 16. PSD of structural response induced by pedestrian density d = 1.0 pedestrians/m2 measured at antinodes of Modes 1 and 2 of WK footbridge, low-pass filtered with cut-off frequencies of (dashed) 10 Hz and (solid) 2.3 Hz.
Fig. 17. PSD of structural response induced by pedestrian density d = 1.0 pedestrians/m2 at antinodes of Modes 1 and 2 of WK footbridge, low-pass filtered with cut-off frequency of (dashed) 10 Hz and (solid) 2.3 Hz. (black) Measured; and (gray) simulated without HSI.

Parameter Estimation

In a preliminary step, an MSD interaction model with parameters corresponding to recent findings in the literature (natural frequency fh1 = 3 Hz and modal damping ratio ξh1=30%), was used to get a first idea of the impact of HSI for the WK footbridge. The dynamic behavior of the footbridge was simulated using the experimentally identified natural frequencies and modal damping ratios, and the mass-normalized mode shapes of the calibrated FE model. Fig. 18 visualizes the FRF of the WK footbridge, empty and with HSI for a pedestrian density of 1.0 persons/m2. This figure shows that the impact of HSI is high for Modes 1 (f1 = 3.30 Hz) and 2 (f2 = 4.14 Hz). This figure furthermore shows that HSI has a negligible influence on the PSD of the structural acceleration levels for frequencies up to 2.5 Hz.
Fig. 18. FRF for vertical input and output acceleration at antinode of Mode 1 of WK footbridge: (dashed) empty and (solid) with HSI, for a pedestrian density of 1.0 persons/m2.
Next, the inputs for the parameter estimation procedure were set to match the conditions of the in-field observations: 1.0 person/m2, fsN(1.68,0.13) Hz, average body mass 72 kg. As the lateral and longitudinal modal displacements of the relevant modes (Fig. 14) are negligible, the parameter estimation procedure only used the PSD (2–6 Hz) of the vertical acceleration levels. Following the results of Section “Influence of Modeling Errors,” the high impact of HSI on the selected frequency range will limit the impact of modeling errors. The following optimal solution was obtained: fh1=2.8 Hz and ξh1=28%.

Influence of Modeling Errors

Next, the impact of modeling errors on the solution of the parameter estimation problem of the WK footbridge was investigated. The modeling errors considered here are identical to those considered for the Eeklo footbridge (Section “Influence of Modeling Errors”):
1.
Errors on the structural modal parameters: ϵf={1%,5%}, ϵξ={10%,30%}, and ϵm={2%,5%} on the experimentally identified natural frequencies, modal damping ratios, and calculated modal masses, respectively;
2.
Errors on the distribution of step frequencies in the crowd: 2% and 5% on the mean value (ϵμfs) and standard deviation (ϵσfs) of the Gaussian distribution of step frequencies.
Fig. 19 presents the individual and joint impact of the modeling errors on the solution of the parameter estimation problem (estimation error). The estimation error (Δ) is expressed in terms of the percentage of difference with the nominal optimal solution of the parameter estimation problem (Section “Parameter Estimation”). The following observations are made.
1.
The natural frequency of the MSD interaction model fh1 is most sensitive to an error in the natural frequency of the structural modes. This corresponds to the observations made in Section “Influence of Modeling Errors” for structural modes with frequencies in the interval 3.3–4.4 Hz (compare with Modes 1 and 2 of the WK South Footbridge).
An estimation error of 2% (9%) is found for a modeling error of 1% (5%) on the natural frequency of the structural modes. The joint impact of the modeling errors is limited to an estimation error of 15%.
2.
The damping ratio of the MSD interaction model fh1 is most sensitive to errors in the mean value of the step frequencies and the natural frequency of the structural modes. The joint impact of the modeling errors is limited to an estimation error of 35%.
Fig. 19. Estimation error (Δ) of (a) natural frequency and (b) damping ratio of MSD interaction model due to modeling errors: εf (1% black, 5% gray); εξ (10% black, 30% gray); εm (2% black, 5% gray); ϵμfs=ϵσfs (2% black, 5% gray); for a pedestrian density of 1.0 persons/m2.
These results show that the reasonable range of modeling errors that can be expected for this case has a small impact on the estimated natural frequency of the MSD interaction model (15%). The impact on the estimated damping ratio is higher (35%) but still reasonable, given that the structural response is less sensitive to changes in this parameter.
Section “Full-Scale Observations” shows that, as a result of the relative short measurement duration of 8 min, a convergence error of approximately 10% remains. Therefore, the impact of the level of convergence of the PSD of the measured structural response on the estimation error was also investigated. To this end, the parameter estimation problem was solved for G~u¨ equal to the PSD evaluated for four and seven consecutive time windows of length T = 60 s, corresponding to a convergence error of ±20% and ±10%, respectively. The results show that, for this case, an estimation error of 3% (1%) on the natural frequency and 10% (8%) on the damping ratio of the MSD interaction model can be expected for a convergence error of 20% (10%).

Results

The nominal results of Section “Parameter Estimation” are combined with the uncertainty bounds determined in Section “Influence Modeling Errors”:
1.
fh1=2.8±0.4 Hz;
2.
ξh1=28±8%.
Fig. 20 shows that a reasonable fit is obtained between the simulated and measured PSD of the structural response. The agreement has significantly improved around the contribution of the fundamental and second modes of the footbridge. The residual is 2% of the error when HSI is disregarded. Furthermore, these results corroborate recent findings in the literature (Shahabpoor et al. 2016a; Van Nimmen et al. 2017a), as well as the results for the Eeklo footbridge (Section “Case Study Eeklo Footbridge”).
Fig. 20. PSD of structural response induced by pedestrian density d = 1.0 pedestrian/m2 at antinodes of Modes 1 and 2 of WK footbridge: (black) measured and (gray) simulated with HSI using fitted MSD interaction model.

Conclusions

A parameter estimation procedure is proposed to estimate the parameters that govern HSI. An MSD model is used to describe the interaction between each pedestrian and the structure. The dynamic characteristics of the MSD interaction model are estimated by minimizing the discrepancy between the observed and simulated PSDs of the structural response. The procedure assumes that the dynamic behavior of the empty structure, the average weight, and the distribution of step frequencies in the crowd are known. The proposed approach is verified using numerical simulations and the influence of modeling errors is investigated. The results show that, as footbridges and the human body are by nature lightly (2%) and strongly (≈30%) damped, respectively, the structural response is most sensitive to small variations in the natural frequency of the MSD interaction model. The results furthermore show that the parameter estimation problem is mostly sensitive to errors related to the mean value of the distribution of step frequencies and the structural natural frequency and modal mass. The results are less sensitive to modeling errors related to the standard deviation of the distribution of step frequencies and the structural modal damping ratio. The impact of the structural modeling errors decreases as the impact of HSI increases. The impact of modeling errors is case-specific, but can be investigated numerically. Based on such an analysis, the impact of modeling errors can be minimized, for example, by carefully selecting the frequency range and output locations in the parameter estimation problem.
Next, the approach is applied to two real footbridges, where the walking behavior and the structural response induced by high pedestrian densities are observed. The influence of the reasonably expected modeling errors is translated into an expected level of uncertainty of the obtained solution of the parameter estimation procedure. For both case studies, a relatively low impact of the modeling errors is observed, which is due to the high impact of HSI on a number of structural modes. The smallest impact is found on the estimated natural frequency of the MSD interaction model (≈10%). The impact on the estimated damping ratio is higher (≈25%), but is considered still reasonable, given that the structural response is less sensitive to changes in this parameter. The final results show that an estimate of the natural frequency (≈3.0 Hz) and damping ratio (≈34%) of the MSD interaction model is obtained that is in accordance with recent findings in the literature. These estimates are, however, for the first time ever, it is believed, based on full-scale observations involving high pedestrian densities.

Data Availability Statement

Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request. The data related to the Eeklo footbridge and the WK footbridge are available at the KU Leuven Research Data Repository: Replication data for Eeklo Footbridge: Benchmark Dataset on Pedestrian-Induced Vibrations, (Van Nimmen et al. 2021b) and Replication Data for Using Full-Scale Observations on Footbridges to Estimate the Parameters Governing Human–Structure Interaction, (Van Nimmen et al. 2018a).

Acknowledgments

The first author is a postdoctoral fellow of the Research Foundation Flanders (FWO). Financial support is gratefully acknowledged. This work uses experimental data collected on two footbridges: the contributions of everyone involved in organizing these comprehensive measurement campaigns are gratefully acknowledged.

References

AFGC (Association Française de Génie Civil). 2006. Sétra: Evaluation du comportement vibratoire des passerelles piétonnes sous l’action des piétons (Assessment of vibrational behaviour of footbridges under pedestrian loading). AFGC-SÉTRA.
Agu, E., and M. Kasperski. 2011. “Influence of the random dynamic parameters of the human body on the dynamic characteristics of the coupled system structure-crowd.” J. Sound Vib. 330 (3): 431–444. https://doi.org/10.1016/j.jsv.2010.06.029.
Ahmadi, E., C. Caprani, S. Zivanovic, N. Evans, and A. Heidarpour. 2018. “A framework for quantification of human-structure interaction in vertical direction.” J. Sound Vib. 432: 351–372. https://doi.org/10.1016/j.jsv.2018.06.054.
ANSYS. 2007. Basic analysis guide, ANSYS release 11.0. Canonsburg, PA: ANSYS.
Asami, T., O. Nishihara, and A. Baz. 2002. “Analytical solutions to h-infinity and h-2 optimization of dynamic vibration absorbers attached to damped linear systems.” J. Vib. Acoust. 124 (2): 284–295. https://doi.org/10.1115/1.1456458.
Bocian, M., J. H. G. Macdonald, and J. F. Burn. 2013. “Biomechanically-inspired modelling of pedestrian-induced vertical self-excited forces.” J. Bridge Eng. 18 (12): 1336–1346. https://doi.org/10.1061/(ASCE)BE.1943-5592.0000490.
Brownjohn, J., and T. Fu. 2005. “Vibration excitation and control of a pedestrian walkway by individuals and crowds.” Shock Vib. 12 (5): 333–347. https://doi.org/10.1155/2005/857247.
Brownjohn, J. M. W. 2001. “Energy dissipation from vibration floor slabs due to human-structure interaction.” J. Shock Vib. 8 (6): 315–323. https://doi.org/10.1155/2001/454139.
Brownjohn, J. M. W., and A. Pavić. 2007. “Experimental methods for estimating modal mass in footbridges using human-induced dynamic excitation.” Eng. Struct. 29 (11): 2833–2843. https://doi.org/10.1016/j.engstruct.2007.01.025.
Bruno, L., and F. Venuti. 2009. “Crowd-structure interaction in footbridges: Modelling, application to real case-study and sensitivity analysis.” J. Sound Vib. 323 (1–2): 475–493. https://doi.org/10.1016/j.jsv.2008.12.015.
Cappellini, A., S. Manzoni, M. Vanali, and A. Cigada. 2016. “Evaluation of the dynamic behaviour of steel staircases damped by the presence of people.” Eng. Struct. 115: 165–178. https://doi.org/10.1016/j.engstruct.2016.02.028.
Caprani, C., and E. Ahmadi. 2016. “Formulation of human–structure interaction system models for vertical vibration.” J. Sound Vib. 377: 346–367. https://doi.org/10.1016/j.jsv.2016.05.015.
Dang, H., and S. Živanović. 2013. “Modelling pedestrian interaction with perceptibly vibrating footbridges.” FME Trans. 41 (4): 271–278.
Dang, H., and S. Živanović. 2016. “Influence of low-frequency vertical vibration on walking locomotion.” J. Struct. Eng. 142 (12): 04016120. https://doi.org/10.1061/(ASCE)ST.1943-541X.0001599.
Dong, W., M. Kasperski, and G. Shiqiao. 2011. “Change of the dynamic characteristics of a pedestrian bridge during a mass event.” In Proc., 8th Int. Conf. on Structural Dynamics of EURODYN, 1002–1008. https://bwk.kuleuven.be/bwm.
Georgakis, C. T., and E. Ingólfsson. 2014. “Recent advances in our understanding of vertical and lateral footbridge vibrations.” In Proc., 5th Int. Footbridge Conf. http://www.footbridge2014.com/.
Heinemeyer, C., et al. 2009. Design of lightweight footbridges for human induced vibrations - background document in support to the implementation, harmonization and further development of the Eurocodes. JRC-ECCS 2009. Accessed January 5, 2022. http://www.stb.rwth-aachen.de/projekte/2007/HIVOSS/download.php.
Helbing, D., and P. Molnar. 1995. “Social force model for pedestrian dynamics.” Phys. Rev. 51 (5): 4282–4286.
ISO. 1981. Vibration and shock - mechanical driving point impedance of the human body. ISO 5982:1981. Geneva: ISO.
Jimenez-Alonso, J., and A. Saez. 2014. “A direct pedestrian-structure interaction model to characterize the human induced vibrations on slender footbridges.” Inf. Constr. 66 (Extra-1): m007. https://doi.org/10.3989/ic.2014.v66.iExtra-1.
Maca, J., and M. Valasek. 2011. “Interaction of human gait and footbridges.” In Proc., 8th Int. Conf. on Structural Dynamics of EURODYN, 1083–1089. https://bwk.kuleuven.be/bwm.
MATLAB. 2018. 9.7.0.1190202 (R2019b). Natick, MA: The MathWorks Inc.
Matsumoto, Y., and M. J. Griffin. 2003. “Mathematical models for the apparent masses of standing subjects exposed to vertical whole-body vibration.” J. Sound Vib. 260 (3): 431–451. https://doi.org/10.1016/S0022-460X(02)00941-0.
NBN EN 1991-2 ANB. 2013. Eurocode 1: Actions on structures - Part 2: Traffic loads on bridges. Belgian National annex October 2011. Brussel: Bureau voor Normalisatie.
Nocedal, J., and S. Wright. 2006. Numerical optimization. 2nd ed. Berlin: Springer.
Peeters, B., and G. De Roeck. 1999. “Reference-based stochastic subspace identification for output-only modal analysis.” Mech. Syst. Signal Process. 13 (6): 855–878. https://doi.org/10.1006/mssp.1999.1249.
Pimentel, R., and P. Waldron. 1996. “Validation of the numerical analysis of a pedestrian bridge for vibration serviceability applications.” In Proc., Int. Conf. on Identification in Engineering Systems, edited by L. Davis, 648–657. https://michael.friswell.com/ies96.html.
Qin, J. W., S. S. Law, Q. S. Yang, and N. Yang. 2013. “Pedestrian-bridge dynamic interaction, including human participation.” J. Sound Vib. 332 (4): 1107–1124. https://doi.org/10.1016/j.jsv.2012.09.021.
Qin, J. W., S. S. Law, Q. S. Yang, and N. Yang. 2014. “Finite element analysis of pedestrian-bridge dynamic interaction.” J. Appl. Mech. 81 (4): 041001. https://doi.org/10.1115/1.4024991.
Reynders, E., and G. De Roeck. 2008. “Reference-based combined deterministic-stochastic subspace identification for experimental and operational modal analysis.” Mech. Syst. Signal Process. 22 (3): 617–637. https://doi.org/10.1016/j.ymssp.2007.09.004.
Salyards, K., and Y. Hua. 2015. “Assessment of dynamic properties of a crowd model for human-structure interaction modelling.” Eng. Struct. 89: 103–110. https://doi.org/10.1016/j.engstruct.2015.01.016.
Shahabpoor, E., A. Pavić, and V. Racić. 2016a. “Identification of mass-spring-damper model of walking humans.” Structures 5: 233–246. https://doi.org/10.1016/j.istruc.2015.12.001.
Shahabpoor, E., A. Pavić, and V. Racić. 2016b. “Interaction between walking humans and structures in vertical direction: A literature review.” Shock Vib. 2016: 12–17.
Shahabpoor, E., A. Pavić, and V. Racić. 2017. “Structural vibration serviceability: New design framework featuring human-structure interaction.” Eng. Struct. 136: 295–311. https://doi.org/10.1016/j.engstruct.2017.01.030.
Silva, F., and R. Pimentel. 2011. “Biodynamic walking model for vibration serviceability of footbridges in vertical direction.” In Proc., 8th Int. Conf. on Structural Dynamics of EURODYN, 1090–1096. https://bwk.kuleuven.be/bwm.
Van Nimmen, K., G. Lombaert, G. De Roeck, and P. Van den Broeck. 2017a. “The impact of vertical human-structure interaction on the response of footbridges to pedestrian excitation.” J. Sound Vib. 402: 104–121. https://doi.org/10.1016/j.jsv.2017.05.017.
Van Nimmen, K., G. Lombaert, I. Jonkers, G. De Roeck, and P. Van den Broeck. 2014. “Characterisation of walking loads by 3D inertial motion tracking.” J. Sound Vib. 333 (20): 5212–5226. https://doi.org/10.1016/j.jsv.2014.05.022.
Van Nimmen, K., A. Pavic, and P. Van den Broeck. 2021a. “A simplified method to account for vertical human-structure interaction.” Structures 32: 2004–2019. https://doi.org/10.1016/j.istruc.2021.03.090.
Van Nimmen, K., P. Van den Broeck, and G. Lombaert. 2017b. “Inverse identification of the pedestrian characteristics governing human-structure interaction.” In Proc., 10th Int. Conf. on Structural Dynamics, EURODYN 2017, edited by F. Vestroni, F. Romeo and V. Gattulli, 2889–2894. Amsterdam, The Netherlands: Elsevier Procedia.
Van Nimmen, K., P. Van den Broeck, and G. Lombaert. 2018a. “Using full-scale observations to estimate the parameters governing human-structure interaction.” In Proc., 9th Int. Conf. on Bridge Maintenance, Safety and Management. Melbourne, Australia: Taylor and Francis Group, CRC Press.
Van Nimmen, K., P. Van den Broeck, G. Lombaert, and F. Tubino. 2020. “Pedestrian-induced vibrations of footbridges: An extended spectral approach.” J. Bridge Eng. 25 (8): 04020058. https://doi.org/10.1061/(ASCE)BE.1943-5592.0001582.
Van Nimmen, K., J. Van Hauwermeiren, and P. Van den Broeck. 2021b. “Eeklo footbridge: Benchmark dataset on pedestrian-induced vibrations.” J. Bridge Eng. 26 (7): 05021007. https://doi.org/10.1061/(ASCE)BE.1943-5592.0001707.
Van Nimmen, K., G. Zhao, A. Seyfarth, and P. Van den Broeck. 2018b. “A robust methodology for the reconstruction of the vertical pedestrian-induced load from the registered body motion.” Vibration 1 (2): 250–268. https://doi.org/10.3390/vibration1020018.
Venuti, F., V. Racic, and A. Corbetta. 2016. “Modelling framework for dynamic interaction between multiple pedestrians and vertical vibrations of footbridges.” J. Sound Vib. 379: 245–263. https://doi.org/10.1016/j.jsv.2016.05.047.
Zang, M., C. Georgakis, and J. Chen. 2016. “Biomechanically excited SMD model of a walking pedestrian.” J. Bridge Eng. 21 (8): C4016003. https://doi.org/10.1061/(ASCE)BE.1943-5592.0000910.
Zhang, M., C. T. Georgakis, W. Qu, and J. Chen. 2015. “Smd model parameters of pedestrians for vertical human-structure interaction.” In Vol. 2 of Dynamics of civil structures, edited by J. Caicedo and S. Pakzad, 311–317. Cham: Springer.
Živanović, S. 2012. “Benchmark footbridge for vibration serviceability assessment under vertical component of pedestrian load.” J. Struct. Eng. 138 (10): 1193–1202. https://doi.org/10.1061/(ASCE)ST.1943-541X.0000571.
Živanović, S., A. Pavić, and E. Ingólfsson. 2010. “Modelling spatially unrestricted pedestrian traffic on footbridges.” J. Struct. Eng. 136 (10): 1296–1308. https://doi.org/10.1061/(ASCE)ST.1943-541X.0000226.
Živanović, S., A. Pavić, and P. Reynolds. 2005. “Vibration serviceability of footbridges under human-induced excitation: A literature review.” J. Sound Vib. 279 (1–2): 1–74.
Živanović, S., A. Pavić, and P. Reynolds. 2007. “Probability-based prediction of multi-mode vibration response to walking excitation.” Eng. Struct. 29 (6): 942–954. https://doi.org/10.1016/j.engstruct.2006.07.004.

Information & Authors

Information

Published In

Go to Journal of Bridge Engineering
Journal of Bridge Engineering
Volume 28Issue 1January 2023

History

Received: Mar 29, 2022
Accepted: Aug 26, 2022
Published online: Nov 9, 2022
Published in print: Jan 1, 2023
Discussion open until: Apr 9, 2023

ASCE Technical Topics:

Authors

Affiliations

Dept. of Civil Engineering, Structural Mechanics, KU Leuven, Ghent 9000, Belgium (corresponding author). ORCID: https://orcid.org/0000-0002-8188-1297. Email: [email protected]
P. Van den Broeck
Dept. of Civil Engineering, Structural Mechanics, KU Leuven, Ghent 9000, Belgium

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share