1. Introduction
Bridges are critical in modern infrastructure, facilitating transportation and connecting communities [
1]. However, they are subject to various environmental forces, including wind, which can induce complex aerodynamic phenomena, leading to structural vibrations and fatigue [
2]. Understanding and mitigating these effects are paramount to ensuring bridges’ safety, durability, and performance.
In this comprehensive examination, we delve into the intricate realm of bridge aerodynamics, focusing on critical aspects such as wind tunnel modeling, aerodynamic mitigation devices, open-jet testing, and integrating aerodynamic mitigation with green energy solutions. We aim to elucidate the challenges and opportunities inherent in mitigating aerodynamic effects on bridges through meticulous analysis and empirical studies.
Firstly, we explore the advantages and limitations of wind tunnel modeling in bridge aerodynamics. Wind tunnels offer controlled environments for simulating wind-induced forces on bridge structures, allowing engineers to assess aerodynamic performance and refine design parameters [
3]. However, scale effects and boundary interference pose significant challenges, necessitating careful interpretation of results and consideration of alternative testing methodologies [
4,
5,
6,
7]. Subsequently, we delve into aerodynamic mitigation devices, crucial in reducing vortex-induced vibrations and enhancing bridge safety. Devices such as handrails, windshields, guide vanes, spoilers, and fairings can act as aerodynamic mitigation devices [
8,
9,
10,
11,
12,
13,
14].
Section 3 provides a detailed review of aerodynamic mitigation devices.
Next, we focus on open-jet testing, a specialized approach for simulating atmospheric boundary layer flow and studying bridge aerodynamics in realistic conditions. By analyzing the validation of flow properties and model placement in open-jet facilities, we gain insights into the advantages of large-scale testing and its implications for understanding bridge behavior under dynamic wind conditions. Finally, we explore the emerging trend of integrating aerodynamic mitigation with green energy solutions, such as installing solar panels on long-span bridges. While previous studies have analyzed wind loading on PV arrays or utilized small wind turbines for energy generation, this work focuses on utilizing PV panel arrangements as a passive aerodynamic mitigation device that simultaneously generates energy. This innovative approach enhances bridge safety and contributes to sustainable infrastructure development by harnessing renewable energy sources. Specifically, the strategic utilization of PV geometry to actively re-profile pressure distributions and induce favorable flow re-attachment distinguishes this approach from prior studies focused merely on the structural loading of PV systems. By integrating solar panel arrangements, we seek to balance energy production and aerodynamic performance, paving the way for resilient and eco-friendly bridge infrastructure.
2. Wind Tunnel Modeling in Aerodynamics: Advantages and Limitations
Traditionally, most analyses of deck performance have been conducted in wind tunnels [
15]. The AASHTO LRFD Bridge Design Specifications state that wind tunnel testing is an appropriate tool for designing long-span bridges [
16]. These tunnels can be closed-circuit or open-circuit. Wind tunnel tests are typically done on scaled models of the desired bridge. Scaling factors usually range from 1:50 to 1:200 [
15]. Over the years, multiple researchers have provided insight into the drawbacks of wind tunnel testing for bridge decks. One study found that the small scales used in wind tunnel tests made it difficult to accurately model mitigation devices like guide vanes, as they became too small to correctly model, relative to the bridge [
17]. Wind tunnel tests are usually performed at lower Reynolds numbers because high velocities are needed in the tunnel to scale the Reynolds number to match full-scale conditions correctly. A comparative study was performed on a streamlined single box girder at 1:50 scale in a wind tunnel and at full scale using CFD. The results showed that under small-scale conditions, aerodynamic forces are more sensitive to Reynolds number changes, leading to inaccurate estimations of vibration amplitude [
18].
However, another type of wind testing called open-jet testing can alleviate some of the issues inherent in typical wind tunnel testing. This is particularly crucial for bridge decks, where open-jet testing enabled the use of a 1:20 scale model to achieve a Reynolds number of 1 million, thus enhancing the fidelity of simulated turbulent flow interactions. Open-jet facilities allow for larger-scale testing than typically possible in other types of wind tunnels. Because of the larger flow domain, interference from wall boundaries does not affect the size of the vortices produced in the flow. One study in an open-jet facility at Louisiana State University compared small-scale wind tunnel tests of low-rise buildings to a larger-scale open-jet test. The results showed that in the smaller-scale tests, the vortices were not allowed to reach peak size due to the constriction of the tunnel, resulting in an underestimation of peak pressure values [
19]. Furthermore, another study on high-rise building cladding pressures, produced in the same open-jet facility, corroborated the previous study’s findings. The results showed that higher pressures were achieved using larger-scale open-jet testing than typical wind tunnel testing [
20].
In conjunction with wind tunnel and open-jet testing, computational fluid dynamics (CFD) is also used to analyze aerodynamic loads. Fundamentally, CFD uses the Navier–Stokes equations to solve for velocity and pressure at desired points within the flow. CFD analysis also provides solutions to the issues with wind tunnel tests. It allows for larger-scale testing, which enables Reynolds numbers to be maintained, and it allows for data collection at all points on the deck section with no intrusion of sensors. One drawback of CFD analysis is computational cost. Depending on the quality of the mesh used in the model and the method used to solve it, computational cost can be significant [
15]. Typically, wind tunnel testing is used in combination with CFD analysis in order to validate the results [
13]
All aerodynamic analyses aim to produce Reynolds numbers that match those measured at full scale. The scaling of the Reynolds number is significant because the prediction of different fluid flow attributes depends on its size. Under-estimation or over-estimation could lead to inaccurate results. Notably, in the case of most scaled analyses, the Reynolds number is underestimated [
21]. A few modifications can be made to the wind tunnel to increase the Reynolds number and match the high Reynolds numbers measured in the field. These modifications include increasing the flow velocity, the characteristic length of flow, pressurizing the wind tunnel, and lowering the wind tunnel temperature [
22].
The Reynolds number Re is a dimensionless number defined by the relationship between inertial and viscous forces in a fluid. When viscous forces dominate, the flow is considered laminar; when inertial forces dominate, the flow is turbulent. The relationship is expressed symbolically as [
21]:
In the preceding equation,
is fluid density,
V is flow velocity,
L is the length scale of flow, and
μ is dynamic viscosity [
21]. It is crucial to develop the correct Reynolds number for the type of flow being modeled. Internal flows are laminar for Reynolds numbers up to 2300 and turbulent for Reynolds numbers larger than 4000. In contrast, external flows are turbulent at Reynolds numbers greater than 3 × 10
5 [
23].
The changes in fluid behavior are highly dependent on
. For instance, observing flow over a cylinder (
Figure 1a), increasing
transitions the flow from laminar with no separation (
< 5) to the formation of stable vortices, to eventual vortex shedding instability (
40), and fully turbulent shedding (300 <
< 3 × 10
5). At very high Reynolds numbers (
3.5 × 10
6), the wake narrows, the shear layer becomes turbulent, and vortex shedding dissipates [
24].
Reynolds number dependence and scaling are critical in the turbulent boundary layer. This layer is developed by the interaction of the shear force at the face of the wall with the flow along the wall. The result of this interaction is a logarithmic velocity profile [
21]. In this region, the vorticity is non-zero. In other words, for high Reynolds number cases, the boundary layer is composed of turbulent eddies [
25]. Two significant effects result from increasing the Reynolds number. First, the velocity profile’s curvature increases when everything is held constant and the velocity increases. Second, when everything is held constant, and the characteristic length is increased, the extent to which velocity and vorticity govern the overall domain is reduced because the shear layer becomes a smaller fraction of the overall domain. The second effect is the most important for wind tunnel tests and designing scaled models [
21]. One study compared flow with Reynolds numbers of 3.62 × 10
5 and 1.8 × 10
4 over a deck modified with guide vanes. There was a more significant flow blockage at the lower Reynolds number at the guide vane due to the larger shear layer at low Reynolds numbers [
17].
Another critical component of the turbulent boundary layer is the friction velocity, which is tied to the shear force at the boundary wall. It is a function of the Reynolds number and is essential for predictive modeling of eddy formation in the boundary layer [
26]. One significant aspect of the turbulence generated by wind tunnel tests is that it has a large scaling range. The difference between the smallest and largest eddy in the flow is significant. These large scaling ranges can only be achieved at high Reynolds numbers and are comparable to those measured at full scale in the field, which was shown by a study that used a wind tunnel with an active grid that allowed the characteristic length of flow to be altered in order to achieve a very high Reynolds number, Re = 2.2 × 10
7 [
27].
Similarly, a different study employed an active grid to increase the Reynolds number. It found a measurable change in the quality of the turbulence produced as the Reynolds number increased. An extensive scaling range of eddies was produced at the high Reynolds numbers. The range was similar to full-scale atmospheric conditions [
28]. Another study performed a wind tunnel test on a flat plate to test different Reynolds number scaling attributes. A logarithmic velocity profile was found to be an acceptable assumption for Reynolds numbers between Re = 1.4 × 10
3 and 3.1 × 10
4 [
29].
Specifically, the Reynolds number is vital for modeling and designing long-span bridges due to its relationship with the Strouhal number and vortex-induced vibrations [
10]. Vortex-induced vibrations became prominent due to the Tacoma Narrows Bridge collapse in 1940. These vibrations are created by wind flowing over a bluff deck, creating alternating vortices at the top and bottom of the deck [
30]. The vortices apply a force normal to the flow direction and, as a result, induce vibrations in the deck. These alternating vortices are called von Karman vortex shedding. The bluff bodies’ Strouhal number (
S) defines the frequency at which the shedding occurs.
N,
D, and
U are shedding frequency, defining dimension, and velocity, respectively [
31].
The Strouhal number is based on the geometry of the deck and, under most conditions, remains constant (along with the defining dimension,
D). To maintain a constant Strouhal number, shedding frequency increases linearly with wind speed [
31]. However, when the wind speed increases to a certain velocity (critical velocity), an aerodynamic phenomenon called lock-in occurs. This happens when the wind velocity causes vortex shedding that produces a shedding frequency similar to the bridge’s natural frequency. As a result, the wind velocity no longer controls the shedding frequency, and they do not maintain their linear relationship; instead, the natural frequency of the bridge controls. During lock-in, the velocity keeps increasing, but the shedding frequency remains constant with the bridge’s natural frequency. The most prominent deck vibrations occur when the bridge is locked into this zone [
32].
The Strouhal number is sensitive to changes in the Reynolds number. One study showed this by analyzing fluid flow over a sphere. The results showed that the Strouhal number increased linearly with the Reynolds number until the Reynolds number was approximately 3 × 10
5, after which the Strouhal number began to decrease and levels out around a Reynolds number of 1 × 10
5. Higher Strouhal numbers are representative of higher frequency vortex shedding occurring in the wake [
33], which was corroborated by a study that also found the Strouhal number increased linearly with the Reynolds number [
34]. Another study found that the attachment of mitigation devices like handrails and windshields reduced the Reynolds number sensitivity of the deck, producing a lower Strouhal number that did not fluctuate with changing Reynolds numbers. Comparatively, the bare deck was shown to have significant Reynolds number sensitivity, resulting in a Strouhal number that fluctuated with changes in Reynolds number [
10]. The angle of attack can also affect the Reynolds number sensitivity of the Strouhal number of a long-span bridge deck, which was determined by a study that used a 9 m × 9 m wind tunnel to test a 1:10 scale model. The Strouhal number became more sensitive to the Reynolds number at larger angles of attack, such as + or −5 degrees [
35].
Pressure distribution is also affected by the Reynolds number. Analysis of the pressure coefficients
Cp determines the flow behavior around the bluff body. For bridge decks, positive pressure coefficients indicate no flow separation around the deck. Conversely, negative pressure coefficients indicate that suction or flow separation is occurring. Pressure coefficients vary with the angle of attack. With negative angles of attack, positive pressure coefficients are observed at the leading edge of the deck. However, the pressure coefficients become negative as the angle decreases to zero and then increase back into the positive range. This suggests that positive angles of attack lead to more flow separation at the leading edge [
36]. In the following equation, p,
, and V represent pressure, fluid density, and velocity.
These coefficients can be plotted over the bridge deck’s width to understand better the flow characteristics around the deck [
37]. One study used an open-jet facility for testing. It compared the wind pressures measured at low (Re = 6 × 10
4) versus high (Re = 0.1 × 10
7) Reynolds number tests. The results showed that wind pressures are underestimated at lower Reynolds numbers [
19]. Another study performed in the NRCC wind tunnel tested a 1:10 scale long-span bridge model. It was shown that there is sensitivity of pressure distribution to changes in Reynolds number, especially at larger angles of attack [
35].
Furthermore, testing a 1:50 scale high-rise building in an open-jet facility produced higher peak pressure coefficients than the model tested at a smaller scale in a typical wind tunnel. The open-jet model had a Reynolds number of 0.1 × 10
7 while the wind tunnel model’s Reynolds number was approximately 6 × 10
4. The study concluded that the higher peak pressures resulted from larger-scale turbulence forming at higher Reynolds numbers [
20].
Producing a CFD model with a Reynolds number that is too low will lead to inaccurate modeling of the vortices in the flow. At lower Reynolds numbers, the flow modeled is laminar. It will not accurately represent the turbulent action occurring in the wake because viscous shear stresses dominate at low Reynolds numbers [
23]. As the Reynolds number increases, the flow transitions into a flow with 2D oscillations [
38]. One study compared the results from a wind tunnel and CFD model. The model scale used for both was 1:20, considered large-scale, to ensure the smallest possible Reynolds number effects from the wind tunnel model. The vortex-induced vibration response showed discrepancies where a model 2.5 times smaller was used [
39]. It was concluded that Reynolds number scaling was the cause [
13]. Another comparison of wind tunnel versus CFD was performed by comparing the results of the Re = 2000 wind tunnel model versus the Re = 20,000 CFD model. It was found that the wind tunnel underestimated the vortex-induced vibration amplitude, leading to the conclusion that vortex-induced vibrations are highly dependent on Reynolds number [
37], which was supported by another study that found a correlation between increase in Reynolds number and vibration amplitude by analyzing a CFD model at different Reynolds numbers [
34].
3. Aerodynamic Mitigation Devices: Principles and Applications
Aerodynamic mitigation devices are integral to proactively safeguarding the structural integrity and serviceability of long-span bridges by counteracting detrimental wind-induced vibrations. These devices, either integrated into initial bridge designs or strategically retrofitted, serve to fundamentally modify the airflow patterns around the bridge deck, thereby suppressing the formation and intensity of vortices that induce vibrations [
15]. Their implementation represents a critical advancement in modern bridge engineering, ensuring enhanced safety and performance under dynamic wind loads.
The fundamental principle governing these devices is the alteration of deck geometry to disrupt organized vortex shedding, primarily impacting von Karman and Kelvin-Helmholtz (K-H) vortices (
Figure 1). von Karman vortices arise from the interaction of shear layers in the wake, while K-H vortices result from shear layer unsteadiness around the deck. Through engineered flow separation and wake widening, mitigation devices effectively reduce both the frequency and amplitude of these vortices, thereby diminishing vortex-induced vibrations. By increasing the distance between the upper and lower shear layers at the trailing edge, mitigation devices achieve larger-scale, lower-frequency vortex shedding, a condition less likely to coincide with the bridge’s natural frequencies [
40]. Comprehensive analysis using wind tunnel testing and computational fluid dynamics (CFD) is essential for predicting and optimizing these flow alterations.
Below, we detail the application, mechanisms, and performance of several key aerodynamic mitigation devices:
3.1. Handrails
Handrails, while primarily fulfilling essential pedestrian and driver safety functions on long-span bridges (
Figure 2a), simultaneously provide significant aerodynamic mitigation benefits [
41]. Research indicates that handrails can induce positive pressure on the top surface of the deck by slowing down the airflow, which in turn reduces localized vortex shedding. Furthermore, handrails increase flow separation at the leading edge of the deck, contributing to a reduction in the vortex shedding frequency. Studies on design parameters, such as railing bar placement, have shown differential impacts on vibration modes; for instance, a mid-railing height placement is effective for vertical vibrations but less effective for torsional vibrations. Crucially, the inclination angle of handrails is directly correlated with vibration suppression, with an inclination of 20 degrees determined to be optimal for reducing streamwise velocity in the wake and providing superior vibration control compared to a perpendicular alignment [
38,
42]. These findings underscore the importance of integrating aerodynamic considerations into the multi-functional design of handrails.
3.2. Windshields
Windshields (
Figure 2b [
11]) are employed on long-span bridges to create areas of reduced wind exposure, enhancing driver comfort and safety [
10]. Their design, characterized by reduced porosity, directly contributes to a reduction in mean wind velocity over the bridge deck [
43]. From an aerodynamic perspective, lower porosity barriers generally exert a greater influence on the surrounding airflow. Windshields promote significant flow separation at the leading edge of the bridge deck, leading to the formation of large-scale, low-frequency vortex shedding in the wake. While this effectively reduces vertical vortex-induced vibrations, it can paradoxically increase torsional vibrations, potentially escalating the bridge’s vulnerability to torsional flutter [
44]. This dual aerodynamic impact necessitates meticulous design optimization and, often, the strategic combination with other mitigation devices to address the increased torsional response and maintain overall stability.
3.3. Guide Vanes
Guide vanes (
Figure 2c) are typically positioned beneath the bridge deck, frequently alongside maintenance rails, which themselves can induce vortex-induced vibrations. Their primary function is to redirect airflow and reduce pressure fluctuations at the base of the deck, thereby suppressing vortex formation [
13]. Studies on various placements have demonstrated that internal guide vanes, particularly on the leeward side, are effective in mitigating vortex formation and suppressing vibrations, whereas external placement may have limited effect [
45]. A notable real-world success story includes the Storbælt suspension bridge, where guide vanes installed at the bottom deck’s windward and leeward edges effectively suppressed significant post-erection vortex-induced vibrations [
12]. While guide vanes can influence pressure distribution by increasing negative pressure (suction) at the leading edge, their principal benefit lies in their capacity to actively control flow separation beneath the deck, preventing detrimental vortex formation and enhancing stability [
40].
3.4. Spoilers
Spoilers (
Figure 2d) are typically affixed to the top of barriers or handrails and can be inclined at various angles. These devices have demonstrated remarkable efficacy in suppressing vortex-induced vibrations [
13]. Research indicates that spoilers can completely eliminate vertical vortex-induced vibrations and effectively suppress the pressure fluctuations that lead to them, showing minimal dependency on inclination angle for vibration response. Furthermore, spoilers can significantly reduce torsional vibration amplitude by generating turbulent flow that actively counteracts these vibrations [
40]. A compelling case study is the Humen Suspension Bridge in China, where the installation of spoilers successfully reduced vertical vortex-induced vibration amplitude by half, addressing a critical low damping ratio [
46]. Spoilers thus represent a highly effective and robust mitigation strategy for improving the aerodynamic stability of bridge decks.
3.5. Fairings
Fairings are specifically designed aerodynamic mitigation devices that streamline the bluff edges of a box girder (
Figure 2e), fundamentally altering the flow patterns around bridge decks and reducing aerodynamic instabilities. Their primary mechanism involves modifying the geometry to achieve a more favorable aerodynamic shape, which can lead to reduced vortex formation and associated vibrations.
However, the effectiveness of fairings is highly dependent on precise design parameters, particularly their inclination and geometry. For instance, one study revealed that sub-optimal placement or design, such as a slight bottom angle (around 10 degrees), can paradoxically escalate vortex formation at the top of the deck, leading to an
increase in vibration amplitude [
47]. This emphasizes that not all fairing configurations will yield positive mitigation outcomes.
Further research investigating the influence of different angles highlighted this sensitivity: while shallow bottom angles were found to increase vortex shedding, fairings with bottom angles ranging from 15 to 25 degrees effectively angled the nose upward, resulting in a demonstrable decrease in aerodynamic response [
48]. This critical insight underscores the necessity of optimized design and extensive aerodynamic analysis to ensure fairings achieve their intended mitigation.
Real-world applications further illustrate these complexities. For example, the Trans-Tokyo Bay Bridge experienced significant vortex-induced vibrations, reaching up to 50 cm under transverse winds. Although wind tunnel tests on a bridge model incorporating fairings alongside double flaps and a skirt successfully reduced the vibration amplitude to 18 cm, this still did not meet the stringent allowable amplitude of 10 cm [
49]. This case study demonstrates that while fairings can contribute to vibration reduction, they may require integration with other mitigation strategies or more refined optimization to fully address severe aerodynamic challenges on long-span bridges.
Overall, fairings offer a viable pathway for aerodynamic mitigation by streamlining bridge deck geometries. Nevertheless, their successful implementation is contingent upon rigorous aerodynamic design and validation, with particular attention to factors like inclination angles, to ensure they effectively reduce rather than exacerbate wind-induced vibrations.
3.6. Summary of Aerodynamic Mitigation Devices
The diverse array of aerodynamic mitigation devices—including handrails, windshields, guide vanes, spoilers, and fairings—offers multifaceted solutions for enhancing bridge stability against wind-induced vibrations. While each device operates on distinct aerodynamic principles, ranging from shape modification and flow streamlining to targeted vortex suppression and reattachment control, their individual efficacy is often contingent on precise design, optimal placement, and site-specific environmental conditions. Critical insights gleaned from both advanced laboratory experiments (such as wind tunnels and open-jet facilities) and real-world case studies consistently highlight the imperative of comprehensive aerodynamic analysis to accurately predict and optimize their performance. Furthermore, the inherent complexities of long-span bridge aerodynamics frequently necessitate the integrated application of multiple mitigation strategies to achieve stringent vibration reduction targets, thereby contributing to the development of more resilient and durable infrastructure. This necessity for integrated solutions provides the foundational rationale for exploring multi-functional geometric modifications, such as the proposed integration of photovoltaic systems for simultaneous aerodynamic control and energy production. Future research efforts should continue to explore novel designs, synergistic combinations, and adaptive control mechanisms to address increasingly intricate aerodynamic challenges in contemporary bridge engineering.
7. Conclusions
This study presents a comprehensive exploration into advancing bridge aerodynamics, notably by integrating green energy solutions with traditional mitigation strategies to foster resilient and sustainable infrastructure. The research specifically addressed critical aspects of wind tunnel modeling, the principles of aerodynamic mitigation devices, and the efficacy of open-jet testing, culminating in an innovative investigation into solar panels as dual-purpose aerodynamic mitigation devices.
Firstly, our analysis affirmed wind tunnel modeling as a valuable tool for understanding complex aerodynamic phenomena, while simultaneously highlighting its inherent limitations, such as scale effects and boundary interference, which necessitate complementary advanced methodologies. This underscored the importance of alternative approaches, particularly large-scale open-jet testing, which was demonstrated to accurately replicate atmospheric boundary layer (ABL) flow conditions and produce more realistic peak pressure coefficients on bridge deck models compared to smaller-scale wind tunnel tests. The Open-Jet facility at the Windstorm Impact, Science, and Engineering (WISE) research facility provided crucial insights by enabling tests at higher Reynolds numbers and allowing for the development of larger-scale turbulence, thus enhancing the fidelity of flow measurements relevant to full-scale bridge behavior. It is important to note that while this research focused on a streamlined single box girder, the principle that higher-Reynolds-number testing more accurately captures turbulent flow characteristics (vortex size, pressure coefficients) applies broadly to bluff body aerodynamics, suggesting similar benefits for non-streamlined bridge deck sections. However, the specific geometry optimization developed here (Arrangement 4) is currently tailored for the streamlined single box girder, suggesting that customized PV configurations would be necessary to achieve comparable mitigation performance on inherently bluff sections.
Furthermore, while conventional aerodynamic mitigation devices like handrails, windshields, guide vanes, spoilers, and fairings effectively modify airflow patterns to suppress vortex-induced vibrations, this research introduced a paradigm shift by demonstrating the aerodynamic efficacy of integrated solar panel systems. Our experimental validation of Arrangement 4 in the Open-Jet facility yielded significant quantitative improvements in aerodynamic performance:
The integration of solar panels resulted in a remarkable 54% reduction in the peak negative pressure coefficient at the windward top face of the bridge deck, decreasing from −2.296 (bare-deck) to −1.065. This substantial reduction directly translates to a diminished magnitude of wind-induced forces, thereby enhancing overall aerodynamic stability and reducing susceptibility to vibration. This result favorably compares the geometric mitigation efficacy of the integrated PV system to single-purpose traditional mitigation devices like spoilers, while simultaneously providing renewable energy production. This level of peak pressure reduction (54%) is highly competitive with reported performance metrics of high-impact, single-purpose devices, such as spoilers, which have been shown to reduce vertical VIV amplitude by approximately 50%.
Crucially, complete flow reattachment at the deck top was observed with Arrangement 4, evidenced by a 71% increase in the peak positive pressure coefficient (0.669 at tap 28) compared to the bare-deck model. This reattachment capability is vital for suppressing detrimental vortex formation and mitigating vortex-induced vibrations.
Overall, the addition of the solar panels in Arrangement 4 led to a general decrease in the magnitude of pressure coefficients across the deck, effectively re-profiling the pressure distribution into a more aerodynamically favorable configuration that mitigated adverse wind effects. This solar panel integration offers a sustainable alternative, suitable especially for long-span bridges where energy autonomy is desired and where initial aerodynamic stability challenges allow mitigation through geometric modification. Conversely, highly complex or extreme aeroelastic problems (e.g., flutter) might still necessitate the application of optimized traditional devices (spoilers, guide vanes) or mechanical solutions (TMDs).
These findings definitively conclude that strategically integrated solar panels serve as highly effective aerodynamic mitigation devices, simultaneously producing renewable energy and substantially enhancing the safety and performance of long-span bridges. This interdisciplinary approach, combining advanced aerodynamic analysis with sustainable design principles, paves the way for the development of more resilient, efficient, and eco-friendly bridge infrastructure capable of meeting future environmental and energy demands.
Future research should prioritize optimizing solar panel parameters (e.g., specific tilt angles and spacing) to maximize simultaneous energy generation and aerodynamic stability across varying wind conditions. Specifically, comprehensive wind tunnel or open-jet testing comparing Arrangement 4 performance against optimized traditional mitigation devices (e.g., highly inclined spoilers or fairings) is required to definitively position this system within the existing hierarchy of mitigation solutions. Furthermore, addressing the scale disparity observed in Open-Jet testing requires investigation using models with larger panel dimensions or employing computationally intensive methods such as Large Eddy Simulation (LES) to accurately capture complex turbulence interactions near the panel geometry, potentially yielding higher peak pressures consistent with experimental findings.