Investigation of Retrofit Strategies to Extend the Service Life of Bridge Structures through Active Control
Publication: Journal of Bridge Engineering
Volume 30, Issue 2
Abstract
The durability of aging bridge structures has become a serious societal concern. It has been estimated that 40%–50% of the bridge stock in Europe and 36% in the US is approaching and exceeding the intended service life in some cases. Conventional retrofitting methods are generally effective under predetermined loading scenarios and can mitigate to some extent the effect of damage through strengthening and stiffening. However, typical retrofit measures involve the addition of components, which might cause unwanted stress accumulation, and in addition, they cannot perform adaptation after damage to recover functionality. Adaptive structural systems can modify the response under loading using sensors and mechanical actuators, instead of relying solely on the resistance offered through material and geometry. Previous work has shown that well-designed adaptive structures are effective in reducing peak responses under strong loading resulting in configurations that embody far fewer material and carbon resources than conventional passive systems. This work investigates retrofit strategies using active control through mechanical actuators integrated into the bridge's primary load path or as external systems. The objective is to extend the durability of most common bridge types including beam, tied-arch, and cable-stayed. Two active retrofit systems are considered: (1) an external adaptive tensioning (EAT) for beam bridges; (2) linear actuators placed in the hangers and stays of tied-arch and cable-stayed bridges. Depending on the failure mode (e.g., corrosion-, fatigue-induced), the effect of active control is simulated through a quasi-static controller based on least-squares minimization or through a linear quadratic regulator and explicit time-history analysis. Results shows that the stress reduction achieved by the EAT system retrofitted to a concrete bridge with corrosion-induced damage could extend service by approximately 12 years. In both cable-stayed and tied-arch bridges, the stress range amplitude caused by vehicular traffic is reduced below the constant amplitude fatigue limit, potentially extending service beyond 75 years.
Introduction
Previous Work
Aging bridge infrastructure is a cause of significant societal concern. It has been estimated that approximately 40%–50% of the Trans-European Transport Network (TEN-T) bridges are approaching or have exceeded a service life of 50 years (Brian 2004; Wu et al. 2011). A similar situation applies to other transport networks; for example, the American Road & Transportation Builders Association (ARTBA) reported that over 36% of the bridges in the United States require major repair work or replacement (ARTBA 2023). This motivates the need for effective retrofit strategies to address the emerging challenges caused by aging bridge infrastructure.
Global statistics on bridge failures categorized per bridge and traffic type are given in Zhang et al. (2022). Most failures have occurred for beam-type bridges since this is the most common configuration. Highway bridges feature the highest failure occurrence, accounting for 82%, followed by railway bridges at 18%. Within the subset of steel bridges, fatigue is the primary cause of damage, accounting for 67% of the reported failures (Boulent and Chryssanthopoulos 2010). Fatigue-related failure predominantly occurs when complex components comprising built-up structural members are adopted.
The conventional approach to mitigate damage involves the replacement or strengthening of critically affected structural members (Leland et al. 2002; Marzahn et al. 2009; Fawad et al. 2019; Connor and Lloyd 2017). Although conventional retrofit strategies can mitigate the effect of common damage types (e.g., differential settlement, cracks, construction defects, etc.) (Connor and Lloyd 2017; Fatemi et al. 2011), they are not effective under unpredicted load scenarios and their ability to prevent collapse highly depends on the timeliness with which damage is detected. In addition, conventional strengthening by adding or replacing structural components might cause unwanted stress accumulation that could worsen existing issues. For example, in Horgan (2021), it is shown that the propagation rate of existing fatigue cracks might have increased after retrofitting, leading to the failure of one of the longitudinal beams in the deck of a tied-arch bridge. This indicates that retrofitting should be implemented together with a digital twin and monitoring to have a reliable understanding of the bridge behavior.
Conventional civil structures are designed to meet strength and deformation requirements against strong loads that occur very rarely, and therefore, the structural capacity is often underutilized for most of the service life. Adaptive structures employ sensing and actuation to reduce the response under loading. Investigation of different structural types, including spatial frames (Li and Huang 2013; Preumont et al. 2008; Wang et al. 2021, 2020a) and bridges (Rodellar et al. 2002; Wang et al. 2020b), have shown that active control is effective in reducing the displacement response under diverse loading conditions including seismic and pedestrian traffic. Well-designed adaptive load-bearing structures satisfy typical limit states with much lower material usage and emissions generation than conventional passive structures (Sobek 2016; Senatore et al. 2019). Numerical and experimental testing has shown that deformation control is particularly beneficial to reducing material mass and associated emissions by more than 50% for stiffness-governed structures, including high-rise structures (Blandini et al. 2022; Steffen et al. 2022; Reksowardojo and Senatore 2023; Senatore et al. 2024, in press), floor systems (Nitzlader et al. 2022; Reksowardojo et al. 2024a), and long-span bridges (Reksowardojo et al. 2022; Senatore et al. 2017). Integrated structure-control methods that employ simultaneous synthesis of structural topology and actuator placement produce configurations that surpass known limits of material economy (Senatore and Wang 2024).
Previous simulation studies have demonstrated that active control based on external adaptive tensioning (EAT) (Reksowardojo et al. 2024b) can reduce material mass requirements by up to 30% for high-speed railway bridges and satisfy tight acceleration and displacement limits. Further studies using the EAT system with model predictive control (MPC) and state-dependent input constraints (Dakova et al. 2023; Zeller et al. 2023) have shown that up to a 50% increase in damping can be achieved, which could significantly reduce damage accumulated through fatigue. Generally, these studies show that the bridge response can be drastically reduced through a well-designed active control system without relying on additional stiffening components. Compared to conventional retrofit strategies, this can help avoid unwanted stress accumulation that could worsen the effect of existing damage. However, little attention has been given to the formulation and implementation of retrofit strategies that employ active systems to mitigate the effect of damage in aging structures.
New Contribution
A systematic quantification of potential service life extension through retrofit strategies employing active systems, in relation to typical failure modes, has not yet been carried out. This work aimed to fill this gap by investigating how active systems can be employed as key elements of retrofit strategies to improve the durability of different bridge types. Three case studies are investigated through analyses and numerical simulations: a prestressed concrete bridge with significant damage caused by corrosion and equipped with an EAT system, a tied-arch bridge equipped with actuators in the hangers, and a cable-stayed bridge equipped with actuators in the stays. All bridges are subjected to vehicular traffic. Failure modes are analyzed for each bridge type through case studies documented in the literature.
Two active control strategies are formulated: quasi-static deflection control through EAT and vibration control through linear actuators integrated into the main load path (e.g., hangers for the tied-arch bridge and stays for the cable-stayed bridge). The first strategy is implemented to increase the safety margins for all limit states, focusing on mitigating the stiffness loss due to the reduction of the cross-sectional area of the prestressing tendon caused by corrosion. The second strategy is implemented to reduce the vibration response (stress range) under dynamic loads to mitigate damage accumulation caused by fatigue. Depending on the failure mode, e.g., corrosion-induced or fatigue-induced, the effect of active control is simulated through a quasi-static controller based on least-squares minimization or through explicit time-history analysis involving the use of a liner quadratic regulator (LQR), respectively.
The EAT system studied in this work is similar to the under-deck cable-stayed configuration previously investigated in Madrazo-Aguirre et al. (2015), with the main difference being that the struts that deviate the cable from the bridge deck are active elements controlled to reduce the response under loading. The EAT system control formulation has been previously given in Reksowardojo et al. (2024b), and similar systems were investigated in Saad and Mabrok (2022) and Corral and Todisco (2022). Results showed that a significant reduction of the material and embodied carbon for new bridges equipped with EAT systems can be achieved under static (Corral and Todisco 2022) and dynamic (Reksowardojo et al. 2024b) loading. In Saad and Mabrok (2022), a similar strategy to a variable post-tension system with the actuator placed in the underdeck cable is considered. In this work, the actuators are placed in the struts (deviators), which allows for a significant reduction of the control forces and accurate response reduction since the active struts behave as intermediate supports. However, there has been no systematic quantification of the bridge service life extension by retrofitting with active systems. The loading conditions considered in this work account for realistic traffic statistics, including uncertainties such as traffic volume trends, to quantify the service life extension that can be achieved by active retrofitting.
Active Control System Design
Actuation Load
Two active control strategies are formulated: quasi-static deflection control through EAT and vibration control through linear actuators integrated into the main load path (e.g., hangers for the tied-arch bridge and stays for the cable-stayed bridge). A conceptual illustration is given in Fig. 1. In both strategies, the structure is discretized in nel beam elements. Depending on the retrofit strategy, nact elements are active; i.e., they house a linear actuator installed in series (Böhm et al. 2020). This means that the actuator and the housing element are subjected to the same force, i.e., force-serial. An actuator can change length, thereby, depending on the structural topology, modifying the stress and displacement response under loading. In the absence of external loads, assuming small strains and displacements, when an actuator changes its length, stress develops due to the resistance offered by the stiffness of the other elements to which it is connected. A way to model the effect of force-serial actuators is to map the length changes to equivalent actuation loads , where nd is the number of degrees of freedom. contains the force couples caused by the length change of the actuators
(1)

The input matrix contains the force couples produced by a unitary length change of the actuatorswhere = local-to-global transformation matrix for 2D beam elements. Assuming small displacements, linear-elastic equilibrium conditions can be expressed as follows:where = stiffness matrix; and = displacement compensation achieved through actuation
(2)
(3)
(4)
Note that the actuator length change is modeled as a fictitious elastic strain in the housing element. However, in practice, it is inelastic because it is performed by the linear actuator. Therefore, to compute the forces that develop because of the actuator length changes, the inelastic strain must be eliminated from the total strain for the elements housing the actuators, which is expressed as follows:where , , and = force, stiffness, and displacement vector extracted from the corresponding full vectors according to the degrees of freedom of the ith element, respectively.
(5)
Quasi-Static Control
The EAT system comprises a set of variable-length struts installed below the bridge deck and connected to a cable that is fixed at the bridge ends. Each active strut is equipped with a linear actuator that changes length to modify the tension in the cable. Since the cable is eccentric to the beam axis, the change of cable tension through actuation generates a moment that counteracts the external load. An example of an EAT system comprising nact = 3 actuators is shown in Fig. 1. The bridge is modeled through discretization using finite elements of the type beam. The active struts and cable are modeled using beam elements with end releases since they are assumed to be pin-jointed. Deflection control is carried out under quasi-static loading for the prestressed beam since failure is related to corrosion (ULS). The control method has been previously formulated in Reksowardojo et al. (2024b). The main steps are reported here for completeness. The linear-static equilibrium conditions are expressed as follows:where = displacement response caused by the combined action of the external load and actuation load defined in Eq. (2). The output (i.e., observed controlled displacements) is
(6)
(7)
The output matrix is obtained by extracting from an identity matrix the rows corresponding to the indices of the controlled degrees of freedom Sctr. The controlled displacement response can be decomposed in the share caused by the external loads and the compensation achieved through actuation
(8)
The actuator length changes (i.e., control commands) are determined to reduce as much as possible the output y ≈ 0 by minimizing the L2- norm of the difference between the response caused by the external load and the compensation achieved through actuation which can be solved directly using the pseudoinverse operator ()+
(10)
(11)
The cross-sectional dimensions of the cable and struts are determined through volume (i.e., material) optimization subjected to stress constraints taken from ANSI/AISC 360-22 (ANSI/AISC 2022). The optimization problem is stated as follows:where collates the cross-sectional area of cable segments and actuator housing elements. For simplicity, all cable segments and active structs have the same cross-sectional area, respectively. fn is the compression stress capacity (ANSI/AISC 360-22, ANSI/AISC 2022) using a modified form of the Euler buckling force. The active struts have a circular hollow section with a wall thickness set to 10% of the outer diameter. The problem stated in Eq. (12) is a nonlinear programming (NLP) problem that has been solved through the trust-region method implemented in the Python Scipy library (Van Rossum 2020).
(12)
Vibration Control
When dynamic effects are considered, the moving load (i.e., truck HS-20) is assumed to have constant speed. Since the structure is discretized with beam elements, the position of the load does not always coincide with that of the beam end nodes. Following the formulation in Reksowardojo et al. (2024b), a cubic shape function N is implemented to interpolate the load position between the nodeswhere nax = number of axles for each vehicle crossing the bridge. The shape functions arewhere li and ξ(t) ∈ ℝ = length and the location of the moving load in the local coordinate system of each element, respectively. Each axle of the truck is modeled as an external load
(13)
(14)
(15)
This study takes a conservative approach by not incorporating the effect of vehicle–bridge interaction (Reksowardojo et al. 2024b). The equation of motion under dynamic loading iswhere the terms M, C, and K are the mass, damping, and stiffness, respectively. The structure responses are denoted , , and for displacement, velocity, and acceleration. The external and actuation loads [ , ] are time-dependent. The steady state is derived using the state-space formulationwhere
(16)
(17)
(18)
The state vector collates displacement and velocity responses
(19)
The control gain matrix is obtained by implementing a linear quadratic regulator (LQR):
(20)
The LQR formulation is based on minimizing the performance index J, which quantifies the energy associated with the closed-loop system. Weighing matrices Q and R are employed to achieve a tradeoff between the required control performance and the effort (i.e., actuator forces) necessary to achieve it. When the entries in Q are greater than those in R, the control effort will be larger, resulting in a rapid response reduction, and vice versa. Bryson rules are adopted to set the weighting matrix coefficients. Normalization is applied to give equal importance to response reduction and control effort. In this work, the coefficients are set so that the uncontrolled peak displacement is expected to be reduced below serviceability limits (SLS) through actuator length changes smaller than 100 mm. Readers are referred to Reksowardojo et al. (2024b) for further details.
Case Studies
Loading Conditions
Loading conditions are set for all bridge types based on AASHTO (2020), as defined in Table 1. Limit states are evaluated using LRFD as recommended by applicable standards (ACI-318-19, ACI 2022). The dynamic amplification and multilane factor for live load are based on AASHTO (2020) and the load combination for all limit states are defined in Table 2.
Load type | — |
---|---|
Self-weight | pSW |
Superimposed dead load | pSIDL |
Prestress load | pt |
Camber load | pca |
Permanent load | pp = pSW + pSIDL + pt + pca |
UDL and tandem load | pLLa |
Total load | |
Fatigue load (HS-20) | pLLb |
Actuator load | pact |
Limit state | Load combination |
---|---|
ULS | 1.25pSW + 1.50pSIDL + pt + pca + 1.75pLLa + pact |
SLS | pSW + pSIDL + pt + pca + pLLa + pact |
FLS I | 1.75pLLb + pact |
FLS II | pLLb + pact recommended in Tallin and Petreshock (1990) when the load is modeled as a random process |
Mitigation of Corrosion-Induced Damage in Prestressed Beam Bridges
Petrulla Bridge is a three-lane multispan bridge built in the mid-1980s in Sicily (Italy). The cross-sectional dimensions and tendon profiles are given in Fig. 2 and further specifications are given in Table 3. According to Anania et al. (2018) and Bazzucchi et al. (2018), the primary cause for the formation of a plastic hinge in midspan was corrosion in the prestress tendons [Fig. 3(a)]. The investigation shows that the concrete in the I-girder cross section retained adequate mechanical properties and failure occurred due to the advanced state of corrosion in the prestress tendons caused by a noncorrect execution of the grout injection [Fig. 3(b)]. The spacing between adjacent prestress tendons did not meet code requirements (ACI-318-19, ACI 2022), causing an unwanted lack of concrete interface and resulting in a uniform rust distribution throughout the tendon area. Numerical analysis was carried out through nonlinear FEM (Anania et al. 2018). To simulate the effect of corrosion on the bridge response, the tendon cross-sectional area was progressively reduced. It was not possible to achieve convergence when the tendon cross-sectional area was reduced by 61%, indicating a drastic loss of stiffness and potential collapse (Anania et al. 2018). The residual cross-sectional area of 660 mm2, corresponding to an equivalent tendon diameter of 29 mm and an estimated adherence value of 16.5 t/mm per meter, agrees with the evidence gathered through inspections.

Fig. 2. I-girder cross section and prestress tendon arrangement.
[Reprinted from Engineering Structures, Vol. 162, L. Anania, A. Antonio Badalà, and G. D'Agata, “Damage and collapse mode of existing post tensioned precast concrete bridge: The case of Petrulla viaduct,” pp. 226–244, © 2018, with permission from Elsevier.]

Fig. 3. (a) Petrulla bridge collapse; and (b) corrosion in the prestress tendons.
[Reprinted from Bazzucchi et al. 2018, © Frattura ed Integrità Strutturale under CC-BY 4.0 (https://creativecommons.org/licenses/by/4.0/).]
Property | Petrulla | Toutle River | Fleher |
---|---|---|---|
Span | 38 m | 96 m | 368 m |
Bridge width | 11.3 m | 9 m* | 41 m |
Girder spacing | 2.825 m | — | — |
Number of lanes | 3 | 3 | 6 |
Year of construction | mid-1980s | 1969 | 1978 |
Year of inspection | Collapse occurred before inspection | 1998 | 2003 and 2019 |
Year of collapse | 2014 | — | — |
Traffic load | Traffic loads taken from AASHTO (ASSHTO 2020) | 33,000 crossings per day, 22% heavy trucks | 75,000 crossings per day |
Damage | Uniform rust distribution throughout the prestress tendon area | Fatigue cracks and perceivable vibration (Leland et al. 2002) under heavy trucks with speeds greater than 89 km/h | Stays were replaced due to corrosion (Marzahn et al. 2009); 1,158 fatigue cracks in the orthotropic steel deck (Straßen 2019) |
*Assumed value since no data was found in the literature.
Serviceability Limit State
The SLS for prestressed concrete beam bridges must be satisfied to ensure user comfort and prevent damage from cracks and deflections. Vibrational motion is not considered because typical concrete bridges have much higher inertia and damping than steel bridges. To determine whether the SLS is met in the original design (Anania et al. 2018), stress limits to avoid crack formation are taken, assuming full prestressing as recommended in ACI-318-19; (ACI 2022). Compression and tension stress limits arewhere fc = concrete compressive strength after 28 days of curing. The stress caused by the SLS load combination in the upper and lower outermost fibers of the cross section are denoted σSLS,u and σSLS,l, respectively. All terms are defined in the Appendix. This is a more conservative constraint than partial prestressing, which typically does not require the adoption of stress limits. The deformation is calculated based on Nawy (2009) using the response factors that are recommended by the Prestressed/Precast Concrete Institute (PCI) (PCI 2014). These factors account for long-term effects in prestressed concrete members. The deformation limit for simply supported beams is set to a fraction of the span length L/800, as recommended in AASHTO (2020). For clarity, since this analysis concerns an existing bridge, the erection stage is not considered. This means that the stress and deflection after the prestress transfer from the tendon to the concrete are not evaluated.
(21)
(22)
Ultimate Limit State
According to Anania et al. (2018), the bridge failed after 30 years in service upon fracture of the prestressed tendon in the midspan. Consequently, the structure fails in a nonelastic state, and thus failure is strength-related (ULS). Bending moment and shear under ULS load are evaluated as follows:where MR and VR = moment and shear resistance, respectively; MULS and VULS = moment and shear caused by the ULS load combination, respectively (Table 2); and ϕM and ϕV = resistance reduction factor for bending moment and shear, respectively (ACI-318-19, ACI 2022). The resistance factor ϕM for prestressed concrete subjected to bending moment is a function of the strain in the outermost reinforcement, which in this case is assumed to be the rebar as recommended in ACI-318-19 (ACI 2022) or equivalently by AASHTO (2020). The resistance factor ϕV for shear verification is 0.75, as stated in ACI-318-19 (ACI 2022). Considering the high corrosion developed in the prestressed tendons, the moment and shear capacity based on ACI-318-19 (ACI 2022) are reformulated as functions of the prestressed tendon area percentage reduction β. The moment capacity is based on strain compatibility between the cross section and the prestressed tendon by considering both elements fully bounded and having an equal strain. Since it is assumed that there is full compatibility between the tendon and the surrounding concrete, the strain is expressed by referring to the tendon. The total strain in each tendon is the summation of the initial strain , decompression strain , and final strain
(23)
(24)
The reader is referred to the Appendix for the definition of each strain component and to Nawy (2009) for further details. Once the total strain is obtained, the constitutive relation between stress and strain is employed to compute the tendon stress ft. However, since no information was available from the tendon supplier, the relationship was derived from the PCI (Precast Prestressed Concrete: Bridge Design Manual 2014):where Ct,A, Ct,B, and Ct,B = constants defined in the PCI (Precast Prestressed Concrete: Bridge Design Manual 2014). Once the tendon stress is obtained, the concrete compressive height cc(β) can be derived using force equilibrium, as illustrated in Fig. 4, and subsequently, the equivalent (simplified) height ac(β) (ACI-318-19, ACI 2022) is computed. The concrete force Fc is then obtained via integration over the compression area. The moment capacity MR(β) is obtained through equilibrium at the concrete neutral axis (plastic)
(25)
(26)

The shear capacity is calculated based on ACI-318-19 (ACI 2022). In reinforced concrete, the total shear capacity comprises that of the concrete VR,c(β) and transversal reinforcement VR,s (i.e., stirrups)
(27)
Two capacity types are considered: pure shear (web-shear) and flexure–shear interaction. The concrete shear contribution VR,c(β) is the minimum value of the governing shear capacity between pure shear Vcw(β) and flexural–shear interaction Vci(β). For simplicity, it is assumed that the shear capacity of the stirrup is not affected by corrosion since the effect on the prestressed tendon is dominant. Pure shear resistance is expressed as follows:where bw = web width. The compressive stress in the concrete cross section in the centroid caused by the prestress tendon force and the vertical component Vt(β) of the effective prestress force Pt,e(β) arewhere θt = prestress tendon slope, as shown in Fig. 5. Concerning flexure–shear interaction Vci(β), the capacity is taken as the minimum between Eqs. (31) and (32)where Vd = shear force caused by the unfactored dead load (pSW + pSIDL); Mmax,ULS = maximum moment caused by the ULS load; Vi = shear force occurring with Mmax,ULS; and Mcre is the bending moment causing flexural cracks, which is given as follows:where σt,e = concrete compressive stress caused by the effective prestress force Pt,e(β); and σd = concrete tensile stress caused by the unfactored dead load, both evaluated at the outermost lower fiber.
(28)
(29)
(30)
(31)
(32)
(33)

Corrosion Propagation over Time
When corrosion propagates in the prestressed tendon, it might cause a significant loss of cross-sectional area and therefore of structural capacity. A conceptual relationship between the deterioration level caused by corrosion and the structure service life is shown in Fig. 6. The initiation phase is the required time for carbon and/or chloride to go through the concrete cover and reach the reinforcement. An empirical expression of the corrosion rate for concrete reinforcement is taken from (Al-Mosawe et al. 2022; Val and Melchers 1997; Stewart and Rosowsky 1998)

(34)
(35)
(36)
(37)
(38)
The prestress tendon area percentage reduction β(t) is expressed as follows:where At,final and At,initial = final and initial cross-sectional tendon areas, respectively.
(39)
Response Control through EAT in the Prestressed Concrete Bridge (Petrulla)
An illustration of the Petrulla bridge retrofitted with two EAT systems, each comprising nact = 14 actuators for two adjacent I-girders, is given in Fig. 7. Following the investigation in Anania et al. (2018), key design parameters are identified. The minimum concrete cube compressive strength is set to 57.90 MPa ( is 49.21 MPa for concrete cylindrical strength), while the effective prestress force Pt,e to 1,150 MPa. The cross-sectional dimensions of the I-girder and prestressed tendon are given in Table 4. Concerning the EAT system, the yield stress for strut and cable elements fy are set to 350 and 1,470 MPa, respectively. The cross-sectional dimensions of the EAT system members, struts, and cable elements are given in Table 4. Uncontrolled and controlled displacement responses under characteristic loading conditions are illustrated in Fig. 8. The displacement caused by the permanent loads is assumed to be compensated by tensioning the EAT system cable through an external actuation system (e.g., jacking). The tension force required in the cable to reduce completely the displacements caused by the permanent loads is 7,222 kN. Cambering causes an initial compression state in the struts, which must be considered when sizing the actuator housing elements [Eq. (12)]. Concerning the state under vehicle load, the maximum control effort is required to compensate for the effect of a uniformly distributed load plus tandem load pLLa applied in the midspan, as shown in Fig. 8(c). In this case, the maximum required force for the actuator close to the midspan is 12,622 kN with a Δl of 38 mm. The control forces could be reduced by considering additional actuators. That being said, modern electrohydraulic actuators can apply forces up to 22,000 kN (VHT 2022).


Element | Cross section | Inertia (mm4) | Area (mm2) |
---|---|---|---|
Petrulla | |||
Beam bridge | I-beam | 4.62 × 1011 | 6 × 103 |
Prestress tendon diameter 37.4 mm | — | 1 × 103 | |
EAT system | Strut member, circular hollow section 92 × 9.2 mm | 3.51 × 106 | 6 × 103 |
Cable member diameter 110 mm | 7.12 × 106 | 9 × 103 | |
Toutle River | |||
Tie chords | I-beam | 1.52 × 1011 | 2 × 105 |
Arch | Rectangular hollow section | 5.08 × 1010 | 2 × 105 |
Hanger | Circular bar | 7.12 × 106 | 9 × 103 |
Fleher | |||
Steel orthotropic deck | See Fig. 25(c) | 7.38 × 1012 | 1 × 106 |
Concrete deck | See Fig. 25(d) | 8.46 × 1013 | 4 × 107 |
Cable stay, 6 per row | Diameter 110 mm, see Fig. 25(b) | 7.12 × 106 | 9 × 103 |
Pylon | Rectangular hollow section 5 m × 5 m × 1 m | 3.08 × 1013 | 5 × 106 |
Due to a lack of data concerning the load that caused the collapse, lower and upper bound estimates are evaluated based on SLS and ULS, respectively. Since full prestressing is considered, when the SLS limit exceeds, the concrete subjected to tensile stress during service will likely develop cracks. This might cause corrosion of the prestressed tendon, resulting in a reduction of the cross-sectional area, which could lead to collapse. The concrete lower fiber tensile stress σSLS,l is 4.7 MPa [Eq. (22)], which exceeds the limit recommended in ACI-318-19 (ACI 2022) (4.3 MPa). The concrete upper fiber compressive stress σSLS,u1 and σSLS,u2 are 15.3 and 25.2 MPa [Eq. (21)], which are lower than the limits recommended in ACI-318-19 (ACI 2022) (22.3 and 29.7 MPa, respectively). The exceedance in tensile stress is likely to cause and/or worsen corrosion propagation. The evaluation of the ULS for prestressed concrete for uncontrolled and controlled states is plotted against the corrosion rate in Fig. 9.

This evaluation is carried out by reducing the prestress tendon force iteratively, simulating the effect of corrosion. Results are compared with those of Anania et al. (2018) concerning the moment capacity. According to Anania et al. (2018), who adopted nonlinear FEM analysis, no convergence was possible when the tendon area is reduced by 61%. This is in good accordance with the simulation carried out in this work, which indicates failure when the prestress tendon area is reduced by 50% under ULS and by 71% under SLS. This variance is likely due to differences in model fidelity. The benchmark in (Anania et al. 2018) accounts for material nonlinearity allowing stress redistribution within the element as the concrete reaches a plastic state. In contrast, the assessment in this work keeps mechanical properties in the elastic state, resulting in a more conservative evaluation. When the EAT system is employed, the moment MULS and shear VULS under ULS are reduced by 61% and 42%, respectively. This response reduction enables a loss of up to 80% in prestress tendon cross-sectional area. Concerning SLS, a loss of up to 90% could be tolerated using the EAT system, as shown in Fig. 10. Assuming a tolerance of 80% loss in prestress tendon cross-sectional area, the service life could be extended by 12 years or 40% from the recorded time of failure. The service life extension quantification is carried out using the corrosion rate model stated in Eqs. (34)–(38). The start of corrosion propagation is obtained through backward calculations from Eqs. (34) to (38). Knowing that the bridge collapsed after 30 years in operation when the prestress cross-sectional area was reduced by 50%, it can be estimated that corrosion propagation tp initiates at year 18.

Mitigation of Fatigue-Induced Damage in Steel Tied-Arch and Cable-Stayed Bridges
Mitigation of fatigue-induced damage through active control is tested on two types of bridges: tied-arch and cable-stayed. Toutle River is a tied-arch highway bridge located in Washington, DC, with further specifications given in Table 3. Fig. 11 shows a perspective view of the bridge and cracks in the tie chords located at the fillet welds. Figs. 12(a and b) show a clear correspondence between the crack location and the truck position. Most cracks developed due to the truck crossing at one-quarter and three-quarters of the span, which are the positions of the antinodes (i.e., points of maximum amplitude) for the first vibration mode.


Fleher is a cable-stayed highway bridge located in Düsseldorf, Germany. Fig. 13 shows a photograph of the bridge, with further specifications given in Table 3. Details of fatigue cracks that developed in the orthotropic steel deck can be found in Straßen (2019). The causes of such extensive damage have been attributed to vehicle-induced vibrations, inadequate material quality, and complex connection details in the built-up orthotropic deck.

Fig. 13. Fleher bridge.
(Image courtesy of Wikimedia Commons/Christian Fu-Mueller.)
Serviceability Limit State
The SLS for cable-stayed and tied-arch bridges, as well as generally stiffness-governed structures, is based on deflections and vibration limits. Response acceleration limits for vehicular bridges are categorized as perceptible and unpleasant for the user. The perceptible vibration limit is a function of the natural frequency fv and is set to (Nassif et al. 2011). The unpleasant vibration limit is taken as 1.3 m/s2 (Nassif et al. 2011). The deformation limit is L/800, as in AASHTO (2020).
Ultimate Limit State
The ULS is not a critical limit state for stiffness-governed structures such as the cable-stayed and tied-arch bridges considered in this study. However, it was not possible to retrieve the structural element cross-sectional dimensions for both bridges from Roeder et al. (1998) and Straßen (2019). For this reason, the ULS criteria for bridge steel elements, as in AISC 360-22, are employed for element sizing and fatigue analysis. The ULS assessment includes evaluating the response in tension, compression, bending, and shear and the interaction between bending and axial forces. Further details are given in the Appendix.
Fatigue Limit State
Based on the damage assessment in Leland et al. (2002) and Straßen (2019), for both Fleher and Toutle bridges, the governing limit state is fatigue. Fatigue causes deterioration of the material mechanical properties induced by fluctuating stress. The repeated application of elastic stress below the yield value can cause significant deterioration through cracks, pitting, and deformations. The fatigue limit state is evaluated based on Stephens and Fuchs (2001), Russo et al. (2016), and Cui and Wang (2013). The dynamic response under loading is computed using time-history analysis. The fatigue assessment is based on the stress time-history domain. Rain-flow counting is employed to determine the governing stress range by identifying the closed cycle in the stress–strain hysteresis curve (ASTM 2017). The stress–life curve (S–N curve) is employed to determine the number of cycles for fatigue failure under cycling loading (AASHTO 2020). The fatigue life depends on several factors such as connection types and geometrical discontinuities of the structural members. In this work, the built-up members are considered in Category E, as described in Leland et al. (2002), for connections with fillet and groove welds that lack sufficiently smooth geometric transitions (ANSI/AISC 360-22, ANSI/AISC 2022).
Two fatigue limit states are considered: fatigue infinite service life (FLSI) and fatigue finite service life (FLSII). FLSI evaluates the bridge performance under the influence of a single truck per lane with a requirement to contain the stress below the constant amplitude fatigue limit (CAFL). FLSI is evaluated as follows:where σFLSI,rf = stress range obtained from rain-flow counting under the FLSI load combination and σCAFL is 31 MPa (Category E). FLSII involves quantifying the fatigue-induced damage accumulation, which is obtained using the Palmgren–Miner model (Cui and Wang 2013)where = number of cycles for the ith stress level obtained from rain-flow counting under the FLSII load combination; and = corresponding number of cycles to failure. If damage state D = 1, failure is considered to be caused by fatigue. The number of cycles to failure for the ith stress level is obtained as follows:where CFL = constant; and σFLSII,rf = stress range obtained from rain-flow counting under the FLSII load combination. For Category E, CFL is 3.61 × 1011 MPa3. Following Tallin and Petreshock (1990), the probability density of the normalized gross vehicle weight (GVW) with respect to the reference truck HS-20, denoted PDFGWV, is assumed to be a combined lognormal distribution (lognormal–lognormal) comprising heavy-weight and lightweight vehicleswhere pocr,L = occurrence probability of lightweight vehicles; and PDFL and PDFH = probability density functions for lightweight and heavy-weight vehicles, respectively. Table 5 gives the estimated traffic lognormal distribution parameters (Tallin and Petreshock 1990).
(40)
(41)
(42)
(43)
Log normal | Median (μPDF) | Lognormal standard deviation (δPDF) | Probability (pocr) |
---|---|---|---|
Lightweight vehicle PDFL | 0.39 | 0.255 | 0.598 |
Heavy-weight vehicle PDFH | 0.84 | 0.172 | 0.402 |
The distribution of both normalized vehicle weight and velocity is given in Fig. 14. Since no data were reported in Leland et al. (2002) and Roeder et al. (1998), all lanes are assumed to have equal probability distribution in terms of vehicle weight and velocity. The vehicle velocity is assumed normally distributed with a mean and standard deviation of 83.7 km/h and 11.26, respectively (FHWA 2019). The correlation between vehicle weight and velocity is not considered in this study due to the lack of available additional traffic statistics.

The estimated increasing vehicle volume per day is taken from Leland et al. (2002) and Roeder et al. (1998). Based on Leland et al. (2002), in 1969, an average of 20,000 vehicles crossed the bridge daily. A linear trend, with an annual increment of 420 vehicles, was inferred from monitoring data from Roeder et al. (1998). Concerning the cable-stayed bridge, the average daily number of crossings is 75,000 (Baukunst-nrw 2024). Since no traffic volume trend is available, the yearly increase is assumed to be the same as that for the tied-arch bridge.
Vibration Control through Active Hangers in Toutle River Tied-Arch Bridge
It is assumed that linear actuators are retrofitted in the hangers. The bridge dimensions are extracted from Leland et al. (2002) and Roeder et al. (2000) and illustrated in Fig. 15, where the actuators are indicated with thick lines. The cross-sectional properties of the bridge elements are given in Table 4. Since most data regarding cross-sectional dimensions of the element are incomplete, they are obtained to satisfy ULS and are given in Table 8. The deck consists of the main chords (tie chords) and cross beams. Based on Leland et al. (2002) and AASHTO (2020), the detail categories for the tie chords are of Type “E.” To simulate a condition comparable to that determined in the damage assessment (Leland et al. 2002), it is assumed that fatigue and serviceability states are not met since the stress in the tie chords exceeds the CAFL prescribed for Category E (AASHTO 2020) (31 MPa). An illustration of the uncontrolled and controlled displacement responses under characteristic loading conditions is given in Fig. 16. The natural frequency is estimated as 1.6 Hz (period of 0.6s), and the first vibration mode is illustrated in Fig. 17. Three critical points are identified for the tie-chord members: (1) quarter span, (2) midspan, and (3) third-quarter span (Fig. 12). The critical locations for the arch where the stress range is the highest are also indicated in Fig. 17.



The load combinations for the fatigue limit states are defined in Table 2 and reported in Fig. 16. It is assumed that one truck is crossing per lane with a 10 s interval between consecutive crossings, allowing the bridge response to dissipate. Concerning FLSI, the stress response for the upper and lower fibers of the cross section of the tie-chord members at the critical points is shown in Figs. 18 and 19, respectively. Both uncontrolled and controlled responses are given. Similarly, the fatigue limit state is evaluated for the arch element. The peak stress response plot for the arch member at the critical locations is given in Fig. 20. Table 6 gives the stress range obtained from rain-flow counting for the uncontrolled and controlled states. While the response can be generally reduced at the critical points through active control, it could also cause a relative stress increase in other locations compared to the uncontrolled state. Since the structural configuration is internally statically indeterminate, the effect of the actuator actions propagates through all elements. An average 26% reduction in stress in the tie chords generates a 1.6× increase compared to the stress range in the controlled state for the arch members, from 17.86 to 30.21 MPa. However, the actuator-induced stress range is kept lower than the CAFL (31 MPa). To achieve so, the actuator force must be limited to 1,126 kN. Further information concerning the actuation system is given in Table 7.



Location | Maximum rain-flow cycle stress range (MPa) | Reduction (%) | |
---|---|---|---|
Uncontrolled | Controlled | ||
Toutle River | |||
Tie chords, lower fiber | |||
¼ span | 39.01a | 28.34 | 26 |
½ span | 24.74 | 18.74 | 24 |
¾ span | 37.61a | 28.10 | 25 |
Tie chords, upper fiber | |||
¼ span | 38.90a | 26.97 | 31 |
½ span | 21.44 | 16.33 | 24 |
¾ span | 38.41a | 28.21 | 27 |
Arch elements | |||
Critical locations (Fig. 17) | 20.21 | 30.87 | −11 |
Fleher | |||
Lower fiber | 41.01a | 30.25 | 26.23 |
Upper fiber | 31.52a | 22.76 | 27.79 |
a
Exceeding the stress limit for FLSI (31 MPa), AASHTO “Category E: welded built-up members.”
Parameter | Toutle River | Fleher |
---|---|---|
Actuator force | 1,126 kN | Precamber: 5,579 kN |
Traffic load: 2,366 kN | ||
Total: 7,945 kN | ||
Stroke | 25 mm | Precamber:30 mm |
Traffic load: 24 mm | ||
Total: 54 mm | ||
Velocity | 18 mm/s | 46 mm/s |
Available electrohydraulic actuators | Max. 2,000 kN, velocity of 80 mm/s (VHT 2022) | Max. 8,000 kN, velocity of 40 mm/s (VHT 2022) |
Concerning FLSII, the traffic probability distribution parameters in Table 5 were taken, considering 30,000 crossings. Fig. 21 shows the stress response for a 300-vehicle crossing sample. All trucks are assumed to cross the bridge with a velocity of 80 km/h. Fig. 22 shows (1) the uncontrolled and controlled stress response at the critical locations for the arch and (2) the variation of the maximum required actuation force, which reaches 593 kN in compression.


The fatigue life at the critical points in the tie chords is calculated using the Palmgren–Miner model (Mouritz 2012) and Monte Carlo simulation. The fatigue damage accumulation curve plot is given in Fig. 23(a). After 30 years in operation, damage accumulation is estimated to be up to 44% . This figure matches well with that provided in the reference damage assessment (Roeder et al. 1998), which estimates it at 52%. Damage accumulation for the controlled state indicates a significant service life extension up to and above 75 years because most of the stress range obtained from rain-flow counting is reduced below the CAFL threshold of 31 MPa (Category E). As observed for FLSI, a greater damage accumulation occurs for the arch members since the stress at the critical locations increases in the controlled state [Fig. 23(b)]. However, the increased damage accumulation due to active control is not critical for the arch elements and the service life is still expected to be greater than 75 years. The peak response plots in service (SLS) in terms of acceleration and displacement for the uncontrolled and controlled states, considering 30,000 vehicles crossing, are given in Figs. 24(a and b), respectively. Peak values are indicated in Table 8. A significant mitigation of the response can be obtained through the active hangers. Referring to Figs. 24(c and d), in the uncontrolled state, the acceleration response is above the unpleasant and perceptible thresholds in the 98th and 93rd percentiles, respectively. Through active control, the controlled acceleration response is reduced by 95% from the uncontrolled state so that is always below the perceptible threshold. The maximum displacement is also reduced by 39% through active control.
Response | Uncontrolled | Controlled | Reduction (%) |
---|---|---|---|
Toutle River | |||
Displacement | 44 mm | 27 mm | 39 |
Acceleration | 2.4 m/s2a | 0.12 m/s2 | 95 |
Fleher | |||
Displacement | 668 mm | 457 mm | 32 |
Acceleration | 0.585 m/s2b | 0.188 m/s2 | 68 |
a
Exceeding unpleasant vibration limit of 1.3 m/s2.
b
Exceeding perceptible threshold of 0.37 m/s2.


Vibration Control through Active Stays in the Fleher Cable-Stayed Bridge
It is assumed that linear actuators are retrofitted in the stays. Unlike the previous case study, this bridge features an orthotropic deck with complex connections, as shown in Fig. 25. The data concerning sizing and reinforcement of the deck elements and pylon cannot be sourced. Element sizing is carried out to comply with ULS as recommended in ANSI/AISC 360-22 (ANSI/AISC 2022). The cross-sectional dimensions are given in Table 4. An illustration of uncontrolled and controlled displacement responses under characteristic loading conditions is given in Fig. 26. The natural frequency is estimated as 2.7 Hz (period of 0.37 s). The first vibration mode is illustrated in Fig. 27. The critical point for analysis is indicated with a circle mark. This is the location for the highest stress range occurrence under all load combinations.



Following AASHTO (2020), an additional load factor of 1.5 was considered to estimate the fatigue limit state for the orthotropic deck [Fig. 26(f)]. It is assumed that the primary welding connection for the orthotopic deck elements is Category type E as in the previous case study. The combination applied to the truck moving load for the fatigue limit state is detailed in Fig. 26. A vehicle speed of 100–120 km/h is considered as per the records in Gericht stoppt Fleher-Brücke-Blitzer (2014). The peak stress response plot in the uncontrolled and controlled states for FLSI is given in Fig. 28. Table 6 gives the stress range obtained from rain-flow counting for the uncontrolled and controlled states. Through active control, a reduction of up to 27% in stress for the orthotropic deck can be achieved. This effectively reduces upper and lower fiber stress ranges below the CAFL threshold, thereby providing a significant fatigue life extension. Actuator specifications are given in Table 7. Concerning FLSII, traffic parameters are based on those in Table 5 and Fig. 14. A crossing of 30,000 trucks per lane is considered, with the axle load for each truck and velocity determined using the probability density given in Table 5. Fig. 29 shows the stress response for a sample sequence of 300 trucks per lane with a uniform velocity of 100 km/h. Through active control, the peak stress in the upper fiber can be reduced by 30% and that in the lower fiber by 60%. Fig. 30 shows the plot of the fatigue service life. The orthotropic deck fatigue life is calculated using the Palmgren–Miner model (Mouritz 2012). A service life of 50 years is estimated without active control. Conversely, for the controlled state, the fatigue life can be extended beyond 75 years since most of the stress range obtained from rain-flow counting is reduced below the CAFL threshold (Category E).



The peak response plot in service (SLS) in terms of displacement and acceleration for the uncontrolled and controlled states considering 30,000 vehicles crossing is shown in Figs. 31(a and b), respectively. Peak values are indicated in Table 8. The response is significantly mitigated through active control via the stays. The displacement is reduced by 32% from the uncontrolled state. The acceleration in the uncontrolled state is above the perceptible threshold in the 99th percentile [Figs. 31(c and d)]. Through active control, the acceleration is reduced by 68% so that it is always below the perceptible threshold. The analysis done in this study employed a relatively low-fidelity model (2D beam elements 3dofs/node). Results showed that the upper fiber of the orthotropic deck is not as important as the lower fiber. However, crack propagation has also occurred at the upper fiber, according to the inspection documented in Baukunst-nrw (2024). This suggests that a more detailed analysis using shell and 3D beam elements to model more accurately connections and a transversal analysis within the orthotropic deck could be carried out to validate the findings of this study.

Discussion
Concerning prestressed concrete deck bridges, retrofitting using the EAT system described in this work has shown great potential to mitigate damage caused by corrosion. Results shows that a stiffness loss caused by up to 81% reduction in the prestress area tendon can be compensated and 61% of the moment response under ULS can be reduced, which results in an extension of the service life by 12 years. In the tied-arch bridge case study, the actuators are placed within the main load path, namely, the hangers. This retrofit strategy produced an average reduction of 26% in the mean rain-flow range stress at critical points in the tie-chord elements under FLSI. This brings the stress range below the CAFL threshold, ensuring a theoretically infinite cyclic resistance. Concerning FLSII, using Monte Carlo–based simulations, a 61-year service life is estimated for the uncontrolled system. Through the application of the active system, the structure lifespan can be extended beyond 75 years as most of the stress range is reduced below the CAFL threshold. Similar results apply to the cable-stayed bridge study, where the actuators are placed in the stays. Concerning FLSI, the stress range is reduced by 27% and below the CAFL amplitude threshold. Similar to the tied-arch bridge, it is important to ensure an adequate margin for the fatigue life of elements where the stress response increases due to the effect of actuation. An assessment of the finite lifetime under FLSII indicates a service life of 50 years for the uncontrolled system. Through the application of the active system, the structure lifespan can be extended beyond 75 years since most stress ranges remain below the CAFL threshold.
Conclusion
The new contribution of this work is to evaluate the efficacy of active retrofit measures for aging bridge structures. This conceptual exploration has considered different retrofit strategies, including external actuation systems for beam bridges and linear actuators placed in series with primary structural components of tied-arch and cable-stayed bridges. For all case studies, a pronounced reduction in the response under loading, which could lead to a significant service life extension, has been observed. The required control forces, although relatively high, remain within the limits reached by modern actuators, thus showing potential for feasible implementation. Careful consideration must be given to the placement of active elements within the structure when implemented as a retrofitting measure. When the actuators are placed in the main load path, an increase in stress response could occur in noncritically stressed elements. This is, in principle, a good strategy since the force flow is redirected from highly stressed regions to less critical regions, leading to better material utilization. However, it is important to ensure a reasonable safety margin to avoid fatigue-related issues due to the increased stress resulting from the actuator's actions. Additionally, incorporating the EAT component for the retrofit measure significantly reduces the bridge clearance. Therefore, further optimization is necessary to balance spatial and actuator force limits.
Future work could develop strategies involving mathematical optimization that account for these aspects to determine an optimal actuator placement for retrofitting as well as new designs. In this work, numerical analysis has been carried out using low-fidelity models (i.e., planar beam elements). Accurate stress localization is of key importance for the evaluation of fatigue-induced damage accumulation. Thus, a higher degree of fidelity that considers joint details, discontinuities in the cross-sectional members, and submodeling for transverse analysis will provide a more accurate estimate. Further analysis could also consider the uncertainty of key parameters, including material, connection detail, vehicle properties, active control fault scenarios, and road roughness, to carry out a quantification of the active system reliability.
Appendix. SLS Stress Derivation
The stresses caused by the SLS load combination in the upper and lower outermost fibers of the cross section are denoted σSLS,u and σSLS,l, respectively, and expressed in the elastic state
(44)
(45)
The effective prestress force, concrete area, and concrete cross-sectional second moment area are denoted Pt,e,Ac, and Ic, respectively. The external moment caused by self-weight, superimposed dead load, and live load are denoted MSW, MSIDL, and MLLa, respectively. Terms et, yu, and yl denote the tendon eccentricity and the distance of the cross-sectional centroid to the upper and lower outmost fibers, respectively. The index k indicates the tendon number (e.g., five tendons for the Petrulla Bridge).
ULS for Steel Members
The ULS design for the steel member is based on AISC 360-22. Axial, compression, bending, and shear are computed using Eqs. (46)–(50):where NR,t, Ag, and fy = tensile capacity, the gross cross-sectional area of the steel component, and the yield stress, respectively. The tensile force under the ULS load combination NULS,t must be lower than the factored tensile capacitywhere NR,c, fn, and fe = compression capacity, modified Euler buckling stress, and Euler buckling stress, respectively; Lc, Es, and r = buckling length, steel Young modulus, and cross-sectional radius gyration, respectively. The compression force under the ULS load combination NULS,c must be lower than the factored compression capacitywhere MR and Z = flexural strength and cross section plastic modulus, respectively. The moment under the ULS load combination MULS must be lower than the factored flexural strength MULSwhere VRand Av = shear strength and shear resisting area, respectively. The shear under the ULS load combination VULS must be lower than the factored shear strength. The interaction between the normal force and bending moment is considered under the ULS load combination by satisfying Eq. (50).
(46)
(47)
(48)
(49)
(50)
Strain Derivation of Prestressed Concrete under Flexure
The total strain for prestressed concrete comprises three parts: initial, decompression, and final. The initial strain is the compressive part when the cross section is subjected to the prestress forcewhere Et and At = prestress tendon Young modulus and cross-sectional area, respectively. The decompression strain is the effect of a load increment that brings the compressive state in the concrete surrounding the tendon to zero, which is expressed as follows:where and Ec = decompression strain and concrete Young modulus, respectively. The final strain is evaluated near collapse, when the concrete outermost compressive fiber reaches the ultimate limit, which is expressed as follows:where = concrete ultimate strain, which is assumed 0.003 (ACI-318-19, ACI 2022). The concrete cross-sectional compressive height and the distance of the prestress tendon centroid to the outermost compressive fiber are denoted cc and dt (Fig. 4), respectively.
(51)
(52)
(53)
Data Availability Statement
Some or all data, models, or codes that support the findings of this study are available from the corresponding author upon reasonable request.
Acknowledgments
The authors gratefully acknowledge the German Research Foundation (DFG), which provided core funding through the Collaborative Research Center CRC1244 “Adaptive Skins and Structures for the Built Environment of Tomorrow” (Grant No. 279064222). The authors thank Dr. Arka P. Reksowardojo of ILEK—University of Stuttgart for supporting the implementation of the numerical methods for active structural control employed in this work.
Notation
The following symbols are used in this paper:
- system matrix (state-space);
- Ac
- concrete cross-sectional area;
- Ar
- longitudinal rebar cross-sectional area;
- At
- prestress tendon cross-sectional area;
- Av
- steel cross-sectional shear area;
- ac
- concrete equivalent compressive height;
- control matrix;
- bw
- concrete cross-sectional web width;
- damping matrix;
- CFL
- fatigue constant;
- Ct,A, Ct,B, Ct,C
- constants to compute the prestress tendon strain;
- cc
- concrete compressive height;
- input matrix;
- DFLS
- fatigue damage accumulation;
- Dt
- prestress tendon diameter;
- Dt,0
- initial prestress tendon diameter;
- Dt,t
- remaining prestress tendon reduced diameter due to corrosion;
- displacement compensation achieved through actuation;
- displacement response caused by the external load;
- dt
- distance of the prestress tendon centroid to the outermost compressive fiber;
- Ec
- concrete Young modulus;
- et
- prestress tendon eccentricity with respect to the girder centroid;
- Fc
- concrete equivalent force (cross section);
- Fs
- longitudinal rebar force (cross section);
- Ft
- prestress tendon force (cross section);
- fc
- concrete cylindrical compressive strength after 28 days of curing;
- fv
- structural fundamental frequency;
- fy
- steel yield strength;
- identity matrix;
- Ic
- concrete second moment of area;
- icor
- corrosion current density;
- stiffness matrix;
- k
- prestress tendon index;
- mass matrix;
- Mcre
- bending moment causing flexural cracks;
- MR
- bending moment resistance;
- MULS
- bending moment caused by the ULS load combination;
- MULS,max
- maximum moment caused by external loads from ULS combination;
- shape function matrix;
- NR
- axial resistance;
- NULS
- axial force caused by the ULS load combination;
- nsn
- fatigue life expressed in the number of cycles obtained from the S–N curve;
- Pt,e
- prestress tendon effective force;
- PDFGVW
- probability density function for a gross weight vehicle;
- PDFH
- probability density function for a heavy-weight vehicle;
- PDFL
- probability density function for a lightweight vehicle;
- pact
- actuator load;
- pca
- camber load;
- pLLa
- live load UDL and tandem load;
- pLLb
- fatigue load, HS-20 (20-t three-axle highway semitrailer);
- pp
- permanent load;
- pSIDL
- superimposed-dead load;
- pSW
- self-weight load;
- pt
- tendon prestress load;
- ptot
- total load;
- response weighting matrix (LQR);
- actuation weighting matrix (LQR);
- transformation matrix;
- t
- time;
- Vcw
- pure shear resistance;
- Vi
- ULS factored shear caused by external loads corresponding to MULS,max;
- VR
- shear resistance;
- VR,c
- shear resistance from concrete cross-sectional contribution;
- VR,s
- shear resistance from transversal rebar contribution;
- VULS
- shear caused by the ULS load combination;
- w/c
- concrete water per cement ratio;
- ylow
- distance of the cross-sectional centroid to the lower outermost fiber;
- yup
- distance of the cross-sectional centroid to the upper outermost fiber;
- state vectors;
- β
- prestress tendon cross-sectional area percentage loss;
- δPDF
- lognormal standard deviation;
- prestress tendon decompression strain;
- prestress tendon final strain;
- prestress tendon initial strain;
- prestress tendon total strain;
- λ
- corrosion rate;
- μPDF
- lognormal median;
- σd
- concrete tensile stress caused by the unfactored dead load (pSW and pSIDL) at the outermost fiber;
- σFLS,rf
- cyclic stress range obtained from rain-flow counting under FLS load combination;
- σFLS,R
- fatigue stress amplitude limit;
- σSLS,l
- stress under SLS load combination in the lower outmost fiber;
- σSLS,u
- stress under SLS load combination in the upper outmost fiber;
- σt,e
- concrete compressive stress caused by the effective prestress force Pt,e at the outermost compressive fiber;
- ϕM
- bending moment resistance factor;
- ϕN
- axial resistance factor;
- ϕV
- shear resistance factor; and
- length changes of actuators.
References
AASHTO. 2020. LRFD bridge design specifications. 9th edn. Washington, DC: AASHTO.
ACI (American Concrete Institute). 2022. Building code requirements for structural concrete and commentary (reapproved 2022). ACI-318-19 2022. Farmington Hills, MI: ACI.
Al-Mosawe, D., L. Neves, and J. Owen. 2022. Reliability analysis of detoriarated post-tensioned concrete bridges: The case study of Ynsys-y-Gwas bridge in UK. Nottingham, UK: Univ. of Nottingham.
Anania, L., A. Badalà, and G. D’Agata. 2018. “Damage and collapse mode of existing post tensioned precast concrete bridge: The case of Petrulla viaduct.” Eng. Struct. 162: 226–244. https://doi.org/10.1016/j.engstruct.2018.02.039.
ANSI (American National Standards Institute)/AISC (American Institute of Steel Construction). 2022. Specification for structural steel buildings. ANSI/AISC 360-22. Chicago: ANSI/AISC.
ARTBA (American Road and Transportation Builders Association). 2023. ARTBA Bridge Report. https://artbabridgereport.org/.
ASTM. 2017. Practices for cycle counting in fatigue analysis. E08 Committee. West Conshohocken, PA: ASTM.
Baukunst-nrw. 2024. “Fleher Bridge Düsseldorf, Engineering [WWW Document].” Accessed September 18, 2024. https://www.baukunst-nrw.de/en/projects/Fleher-Bridge-533.htm.
Bazzucchi, F., L. Restuccia, and G. Ferro. 2018. “Considerations over the Italian road bridge infrastructure safety after the Polcevera viaduct collapse: Past errors and future perspectives.” Frattura ed Integrità Strutturale 12: 400–421. https://doi.org/10.3221/IGF-ESIS.46.37.
Blandini, L., W. Haase, S. Weidner, M. Böhm, T. Burghardt, D. Roth, O. Sawodny, and W. Sobek. 2022. “D1244: Design and construction of the first adaptive high-rise experimental building.” Front. Built Environ. 8: 814911. https://doi.org/10.3389/fbuil.2022.814911.
Böhm, M., J. L. Wagner, S. Steffen, J. Gade, F. Geiger, W. Sobek, M. Bischoff, and O. Sawodny. 2020. “Input modeling for active structural elements—Extending the established FE-work for modeling of adaptive structures.” In Proc., 2020 IEEE/ASME Int. Conf. on Advanced Intelligent Mechatronics (AIM), 1595–1600. Piscataway, NJ: Institute of Electrical and Electronics Engineers (IEEE).
Boulent, I., and M. K. Chryssanthopoulos. 2010. “A review of metallic bridge failure statistics.” In Bridge maintenance, safety, management and life-cycle optimization, edited by D. Frangopol, R. Sause, and C. Kusko, 635. Boca Raton, FL: CRC Press.
Brian, B. 2004. Sustainable bridges - assessment for future traffic demands and longer lives. Technical Rep. European Railway Bridge Demography: Sustainable Bridges SB-D1. 2.
Connor, R. J., and J. B. Lloyd. 2017. Maintenance actions to address fatigue cracking in steel bridge structures: Proposed guidelines and commentary. West Lafayette, IN: Purdue University.
Corral, M. d. M., and L. Todisco. 2022. “Structural design of responsive under-deck cable-stayed footbridges.” Structures 41: 1501–1514. https://doi.org/10.1016/j.istruc.2022.05.077.
Cui, W., and F. Wang. 2013. “Damage accumulation.” In Encyclopedia of tribology, edited by Q. J. Wang and Y.-W. Chung, 706–709. Boston, MA: Springer US.
Dakova, S., A. Zeller, A. P. Reksowardojo, G. Senatore, M. Böhm, L. Blandini, and O. Sawodny. 2023. “A model predictive control strategy for adaptive railway bridges.” IFAC-PapersOnLine 56 (2): 7686–7691. https://doi.org/10.1016/j.ifacol.2023.10.1170.
Fatemi, A. A., Z. Tabrizian, G. G. Amiri, and M. H. A. Beigi. 2011. “Steel bridges retrofitting techniques.” Int. J. Earth Sci. Eng. 4: 695–698.
Fawad, M., K. Kalman, R. A. Khushnood, and M. Usman. 2019. “Retrofitting of damaged reinforced concrete bridge structure.” Procedia Struct. Integrity 18: 189–197. https://doi.org/10.1016/j.prostr.2019.08.153.
FHWA (Federal Highway Administration). 2019. Chap. 5 in Speed concepts: Information guide. Washington, DC: FHWA.
Gericht stoppt Fleher-Brücke-Blitzer. 2014. WDR Lokalzeit.
Horgan, R. 2021. “Is seismic retrofit to blame for ‘unusual’ failure in Mississippi River bridge.” New Civil Engineer. Accessed January 30, 2024. https://www.newcivilengineer.com/latest/is-seismic-retrofit-to-blame-for-unusual-failure-in-mississippi-river-bridge-14-05-2021/.
Leland, A. C., N. S. Brown, and J. A. Van Lund. 2002. “Rehabilitation study of Toutle river tied-arch bridges.” Transp. Res. Rec. 1814: 37–43. https://doi.org/10.3141/1814-05.
Li, W. P., and H. Huang. 2013. “Integrated optimization of actuator placement and vibration control for piezoelectric adaptive trusses.” J. Sound Vib. 332: 17–32. https://doi.org/10.1016/j.jsv.2012.08.005.
Madrazo-Aguirre, F., A. M. Ruiz-Teran, and M. A. Wadee. 2015. “Design criteria of under-deck cable-stayed composite bridges for short and medium spans.” Struct. Eng. Int. 25: 125–133. https://doi.org/10.2749/101686615X14210663188574.
Marzahn, G. A., S. Sieberth, and S. Gurtmann. 2009. “Replacing stay-cables of the Rhine river bridge Rheinbrücke Flehe without traffic interruption.” In Structures congress 2009, 1–10. Austin, TX: ASCE.
Mouritz, A. P. 2012. “Fatigue of aerospace materials.” In Introduction to aerospace materials, 469–497. Sawston, UK: Woodhead.
Nassif, H., M. Liu, D. Su, and M. Gindy. 2011. “Vibration versus deflection control for bridges with high-performance steel girders.” Transp. Res. Rec. 2251: 24–33. https://doi.org/10.3141/2251-03.
Nawy, E. G. 2009. Prestressed concrete: A fundamental approach. 5th ed. Prentice-Hall international series in civil engineering and engineering mechanics. Hoboken, NJ: Prentice Hall.
Nitzlader, M., S. Steffen, M. J. Bosch, H. Binz, M. Kreimeyer, and L. Blandini. 2022. “Designing actuation concepts for adaptive slabs with integrated fluidic actuators using influence matrices.” Civ. Eng. 3: 809–830. https://doi.org/10.3390/civileng3030047.
PCI (Precast/Prestressed Concrete Institute). 2014. Precast prestressed concrete bridge design manual. 3rd ed. Second Release. Chicago, ILL: Precast / Prestressed Concrete Institute.
Preumont, A., B. De Marneffe, A. Deraemaeker, and F. Bossens. 2008. “The damping of a truss structure with a piezoelectric transducer.” Comput. Struct. 86: 227–239. https://doi.org/10.1016/j.compstruc.2007.01.038.
Reksowardojo, A. P., and G. Senatore. 2023. “Design of ultra-lightweight and energy-efficient civil structures through shape morphing.” Comput. Struct. 289: 107149. https://doi.org/10.1016/j.compstruc.2023.107149.
Reksowardojo, A. P., G. Senatore, M. Bischoff, and L. Blandini. 2024a. “Design and control benchmark of rib-stiffened concrete slabs equipped with an adaptive tensioning system.” J. Struct. Eng. 150: 04023200. https://doi.org/10.1061/JSENDH.STENG-12320.
Reksowardojo, A. P., G. Senatore, M. Bischoff, and L. Blandini. 2024b. “Design and control of high-speed railway bridges equipped with an under-deck adaptive tensioning system.” J. Sound Vib. 579: 118362. https://doi.org/10.1016/j.jsv.2024.118362.
Reksowardojo, A. P., G. Senatore, A. Srivastava, C. Carroll, and I. F. C. Smith. 2022. “Design and testing of a low-energy and -carbon prototype structure that adapts to loading through shape morphing.” Int. J. Solids Struct. 252: 111629. https://doi.org/10.1016/j.ijsolstr.2022.111629.
Rodellar, J., V. Mañosa, and C. Monroy. 2002. “An active tendon control scheme for cable-stayed bridges with model uncertainties and seismic excitation.” J. Struct. Control 9: 75–94. https://doi.org/10.1002/stc.4.
Roeder, C., G. A. MacRae, K. Arima, P. N. Crocker, and S. D. Wong. 1998. Fatigue cracking of riveted steel tied arch and truss bridges (no. WA-RD 447.1). Seattle, WA: Washington State Transportation Center (TRAC).
Roeder, C. W., G. MacRae, P. Crocker, K. Arima, and S. Wong. 2000. “Dynamic response and fatigue of steel tied-arch bridge.” J. Bridge Eng. 5: 14–21. https://doi.org/10.1061/(ASCE)1084-0702(2000)5:1(14).
Russo, F., D. R. Mertz, K. H. Frank, and K. E. Wilson. 2016. Design and evaluation of steel bridges for fatigue and fracture—Reference manual (no. FHWA-NHI-16-016). Vienna, VA: National Highway Institute (US).
Saad, A. S., and M. A. Mabrok. 2022. “Retrofitting of bridge superstructures using negative imaginary control theory.” Int. J. Control 95: 1970–1978. https://doi.org/10.1080/00207179.2021.1889032.
Senatore, G., P. Duffour, and P. Winslow. 2019. “Synthesis of minimum energy adaptive structures.” Struct. Multidiscip. Optim. 60: 849–877. https://doi.org/10.1007/s00158-019-02224-8.
Senatore, G., P. Duffour, P. Winslow, and C. Wise. 2017. “Shape control and whole-life energy assessment of an ‘infinitely stiff’ prototype adaptive structure.” Smart Mater. Struct. 27: 015022. https://doi.org/10.1088/1361-665X/aa8cb8.
Senatore, G., F. Virgili, and L. Blandini. 2024, in press. “Global Optimal Actuator Placement for Adaptive Structures: New Formulation and Benchmarking,” J. Intell. Mater. Syst. Struct. https://doi.org/10.1177/1045389X241293861.
Senatore, G., and Y. Wang. 2024. “Topology optimization of adaptive structures: New limits of material economy.” Comput. Methods Appl. Mech. Eng. 422: 116710. https://doi.org/10.1016/j.cma.2023.116710.
Sobek, W. 2016. “Ultra-lightweight construction.” Int. J. Space Struct. 31: 74–80. https://doi.org/10.1177/0266351116643246.
Steffen, S., A. Zeller, M. Böhm, O. Sawodny, L. Blandini, and W. Sobek. 2022. “Actuation concepts for adaptive high-rise structures subjected to static wind loading.” Eng. Struct. 267: 114670. https://doi.org/10.1016/j.engstruct.2022.114670.
Stephens, R. I., and H. O. Fuchs. 2001. Metal fatigue in engineering. 2nd ed. New York: Wiley.
Stewart, M. G., and D. V. Rosowsky. 1998. “Structural safety and serviceability of concrete bridges subject to corrosion.” J. Infrastruct. Syst. 4: 146–155. https://doi.org/10.1061/(ASCE)1076-0342(1998)4:4(146).
Straßen, N. R. W. 2019. “Sachstand Baumaßnahme Fleher Brücke Auswirkungen auf den Verkehrsfluss der B9 und das übrige Stadtgebiet.”
Tallin, A. G., and T. Petreshock. 1990. “Modeling fatigue loads for steel bridges.” Transp. Res. Rec. 1275: 23–26.
Val, D. V., and R. E. Melchers. 1997. “Reliability of deteriorating RC slab bridges.” J. Struct. Eng. 123: 1638–1644. https://doi.org/10.1061/(ASCE)0733-9445(1997)123:12(1638).
Van Rossum, G. 2020. The python library reference, release 3.8.2. Wilmington, DE: Python Software Foundation.
VHT (Van Halteren Technologies). 2022. “Electro Hydraulic Actuator [WWW Document].” Accessed March 1, 2024. https://vanhalteren.com/application/files/2016/7948/1768/Datasheet_Electro_Hydraulic_Actuator.pdf.
Wang, Q., G. Senatore, K. Jansen, A. Habraken, and P. Teuffel. 2020a. “Design and characterization of variable stiffness structural joints.” Mater. Des. 187: 108353. https://doi.org/10.1016/j.matdes.2019.108353.
Wang, Q., G. Senatore, K. Jansen, A. Habraken, and P. Teuffel. 2020b. “Vibration suppression through variable stiffness and damping structural joints.” Front. Built Environ. 6: 550864. https://doi.org/10.3389/fbuil.2020.550864.
Wang, Q., G. Senatore, K. Jansen, A. Habraken, and P. Teuffel. 2021. “Seismic control performance of a three-story frame prototype equipped with semi-active variable stiffness and damping structural joints.” Earthquake Eng. Struct. Dyn. 50: 3379–3402. https://doi.org/10.1002/eqe.3514.
Wu, H. J., Y. Q. Kang, and P. Lu. 2011. “Service life of European and American bridges.” Adv. Mater. Res. 255–260: 916–920. https://doi.org/10.4028/www.scientific.net/AMR.255-260.916.
Zeller, A., S. Dakova, C. Stein, M. Böhm, G. Senatore, A. P. Reksowardojo, L. Blandini, O. Sawodny, and C. Tarín. 2023. “Bridge state and average train axle mass estimation for adaptive railway bridges.” IEEE/ASME Trans. Mechatron. 28: 1880–1889. https://doi.org/10.1109/TMECH.2023.3277317.
Zhang, G., Y. Liu, J. Liu, S. Lan, and J. Yang. 2022. “Causes and statistical characteristics of bridge failures: A review.” J. Traffic Transp. Eng. 9: 388–406. https://doi.org/10.1016/j.jtte.2021.12.003.
Information & Authors
Information
Published In
Copyright
This work is made available under the terms of the Creative Commons Attribution 4.0 International license, https://creativecommons.org/licenses/by/4.0/.
History
Received: Mar 8, 2024
Accepted: Aug 21, 2024
Published online: Nov 26, 2024
Published in print: Feb 1, 2025
Discussion open until: Apr 26, 2025
Authors
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.
Cited by
- Jun Zou, Zhiyin Xu, Jinyu Lu, Haichen Zhang, Zhicheng Sha, Na Li, Amplification mechanism of double rhombic struts in adaptive beam string structures, Structures, 10.1016/j.istruc.2025.108910, 76, (108910), (2025).
- Arka P. Reksowardojo, Gennaro Senatore, Lucio Blandini, Ian F. C. Smith, Application of Self‐Diagnosis and Self‐Repair on a Truss Prototype That Adapts to Loading Through Shape Morphing, Structural Control and Health Monitoring, 10.1155/stc/8827609, 2025, 1, (2025).
- Liju Elias, B.I. Bijimol, C.V. Geethanjali, Anaswara Anil, T.C. Bhagya, S.M.A. Shibli, Fine-recycled concrete aggregate based cementitious polymer composite coating for corrosion and biocorrosion protection of concrete reinforcing steel bar, Construction and Building Materials, 10.1016/j.conbuildmat.2025.140877, 472, (140877), (2025).