Fact-checked by Grok 2 weeks ago

Vortex shedding

Vortex shedding is a fundamental phenomenon in where a fluid flow past a bluff body, such as a or , causes alternating vortices to detach periodically from each side of the body, forming a pattern of swirling structures in the wake known as a . This unsteady process arises from due to adverse pressure gradients on the body's surface, resulting in oscillating pressure fields and forces on the structure. Vortex shedding typically initiates when the , defined as Re = \frac{UD}{\nu} (where U is the , D is the body's , and \nu is the ), exceeds approximately 47 for Newtonian fluids. The shedding frequency f is quantified by the dimensionless St = \frac{f D}{U}, which remains nearly constant at about 0.2 for circular cylinders across a broad range of Reynolds numbers (roughly $40 < Re < 10^5). This periodicity leads to a regular vortex street with counter-rotating vortices spaced in a manner first analyzed by Theodore von Kármán in the early 20th century. In practical terms, the phenomenon manifests in both gaseous and liquid flows, influencing drag, lift fluctuations, and acoustic emissions such as Aeolian tones—whistling sounds produced by the periodic pressure variations. Vortex shedding holds critical engineering importance due to its role in vortex-induced vibrations (VIV), where the shedding frequency synchronizes with a structure's natural frequency (lock-in), amplifying oscillations that can cause fatigue, structural failure, or enhanced heat transfer. Notable applications include the design of bridges, offshore risers, chimneys, and heat exchangers, where uncontrolled shedding contributed to events like self-excited motions in slender structures; mitigation often involves surface modifications like helical strakes or fairings to disrupt vortex formation. Conversely, controlled vortex shedding is harnessed in energy harvesting devices and aerodynamic studies to optimize performance in flows ranging from aeronautics to marine engineering.

Physical Principles

Definition and Phenomenon

Vortex shedding is an unsteady phenomenon in fluid dynamics where a fluid flowing past a bluff body generates periodic, alternating vortices that detach from the body's trailing surfaces, resulting in oscillating low-pressure regions in the wake. This process disrupts the steady flow, creating a dynamic pattern of fluid motion that exerts periodic forces on the body. The occurrence of vortex shedding requires specific flow conditions around bluff bodies, which are non-aerodynamic shapes—such as cylinders, spheres, or flat plates—that induce substantial flow separation due to their geometry. In these cases, the boundary layer—the thin layer of fluid adjacent to the surface—separates from the body because of an adverse pressure gradient, where pressure increases in the flow direction, causing the fluid to detach and form free shear layers. By contrast, streamlined bodies, resembling fish shapes or airfoils at low angles of attack, promote attached boundary layers with gradual flow divergence, minimizing separation and suppressing vortex formation. Visually, the phenomenon appears as pairs of counter-rotating vortices that alternately peel away from the upper and lower sides of the bluff body, propagating downstream and forming an unsteady wake characterized by regions of high and low velocity. This alternating detachment creates an asymmetric pressure distribution around the body at each instant. Vortex shedding applies to both gaseous and liquid fluids, with air flows past structures like chimneys and water flows around submerged obstacles serving as primary illustrative examples.

Formation of Kármán Vortex Street

Theodore von Kármán provided the foundational theoretical analysis of the in his 1911 paper, where he examined the stability of a two-dimensional array of vortices forming in the wake of a bluff body immersed in an inviscid fluid flow. This theoretical analysis built upon earlier experimental observations of the vortex pattern by the French physicist in 1908. Building on this, von Kármán and Gustav Rubach extended the work in 1912 through experimental validation, confirming the existence of a stable, periodic vortex pattern characterized by alternating rows of counter-rotating vortices. This configuration arises as a self-sustaining structure that minimizes drag and maintains equilibrium against perturbations in the flow. The mechanism underlying the formation of this vortex street is the , a global wake instability that emerges when the flow separates from the bluff body, creating two opposing shear layers. These shear layers roll up due to local instabilities, but their mutual interaction—coupled with adverse pressure gradients in the near wake—induces a feedback loop that causes vortices to shed alternately from each side of the body. The pressure differences across the wake amplify displacements in the shear layers, leading to the periodic release of coherent vortices in an anti-symmetric fashion, which propagates downstream as a coherent street. Geometrically, the vortices in the Kármán street arrange in a staggered, double-row pattern, with the transverse distance h between the rows and the longitudinal distance a between consecutive vortices in the same row satisfying the ratio h/a \approx 0.28 in von Kármán's inviscid stability analysis. This spacing ensures marginal stability of the array, where the induced velocities from neighboring vortices balance to prevent collapse or dispersion of the structure. Experimental observations align closely with this theoretical ratio, though viscous effects introduce slight deviations. The Kármán vortex street remains stable in the laminar regime for Reynolds numbers approximately between 40 and 150, where the shedding maintains a regular, two-dimensional pattern without significant three-dimensional perturbations. Beyond Re = 150, the pattern enters a transitional regime with emerging instabilities, yet the fundamental staggered vortex array persists into turbulent flows at much higher Reynolds numbers, up to around 10^5, albeit with increased disorder and secondary structures.

Mathematical Modeling

Strouhal Number and Frequency Prediction

The Strouhal number serves as a fundamental dimensionless parameter for predicting the frequency of vortex shedding behind bluff bodies in fluid flows. It is defined mathematically as St = \frac{f D}{V}, where f is the vortex shedding frequency, D is the characteristic dimension of the body (such as the diameter for a circular cylinder), and V is the free-stream velocity. This formulation allows engineers to estimate f = St \frac{V}{D}, facilitating the anticipation of periodic wake instabilities. The Strouhal number originates from dimensional analysis applied to the governing variables of the shedding process. Assuming the frequency f depends primarily on the convective scales V and D, the Buckingham Pi theorem yields the dimensionless group St, which encapsulates the ratio of oscillatory to convective time scales in the flow. This approach highlights how shedding periodicity scales inversely with body size and directly with flow speed, independent of absolute magnitudes. For canonical bluff bodies like circular cylinders, experimental measurements yield typical Strouhal numbers in the range of 0.18 to 0.22, reflecting near-constant behavior under steady uniform inflow conditions. These values enable reliable frequency predictions for structures such as chimneys or cables, where D is the cross-sectional dimension. A critical application of the Strouhal number arises in resonance scenarios, where the predicted shedding frequency f coincides with the natural frequency of the structure, resulting in lock-in and significantly amplified vibrations. This condition underscores the parameter's role in linking fluid dynamics to structural response, though detailed structural modeling lies beyond frequency prediction alone. The Strouhal number is validated and determined empirically through controlled experiments on diverse bluff body geometries, using sensors to capture periodic velocity fluctuations in the wake. Techniques such as hot-wire anemometry in wind tunnels measure f directly, allowing computation of St for shapes ranging from circular to rectangular prisms, where values vary systematically with geometry—for instance, increasing for sharper-edged forms due to altered separation points. Seminal wind tunnel studies, employing oscilloscopes and spectral analysis, have established these shape-specific correlations, ensuring predictive accuracy across applications.

Influence of Reynolds Number

The Reynolds number, defined as \operatorname{Re} = \frac{\rho U D}{\mu}, where \rho is the fluid density, U the free-stream velocity, D the diameter of the circular cylinder, and \mu the dynamic viscosity, characterizes the ratio of inertial to viscous forces and delineates distinct flow regimes that govern the onset, stability, and pattern of vortex shedding behind bluff bodies. In low-Re flows (Re < 47), no vortex shedding occurs, with attached steady recirculation bubbles forming instead; shedding initiates in the laminar regime around Re ≈ 47–300, producing unsteady but two-dimensional vortices. However, stable periodic shedding, forming the characteristic , emerges prominently in higher regimes, where transitions in boundary layer and wake turbulence profoundly influence shedding dynamics. Flow regime transitions markedly alter vortex shedding characteristics. In the subcritical regime (approximately 300 < Re < 3 × 10^5), the on the cylinder remains laminar up to separation, while the wake is turbulent, supporting dominant periodic shedding with a well-organized, essentially two-dimensional Kármán street. The critical regime (roughly 3 × 10^5 < Re < 10^6) introduces a drag crisis, where the undergoes laminar-to-turbulent transition before separation, leading to delayed separation, a sharp drop in drag coefficient, and temporary disruption or irregularity in the vortex shedding pattern. Beyond this, in the supercritical regime (Re > 10^6), a fully turbulent re-establishes earlier separation, restoring vortex shedding albeit with greater irregularity and three-dimensionality due to enhanced . The Strouhal number, = f D / U (where f is the shedding frequency), exhibits regime-dependent variations that reflect these transitions. In the subcritical regime, remains nearly constant at approximately 0.20, indicating consistent shedding frequency scaling with flow velocity, though it shows a slight increase from about 0.18 at Re ≈ 300 to 0.21 at Re ≈ 10^5. During the critical regime, fluctuates and often increases to around 0.27-0.30 due to the altered separation and changes in vortex coherence amid the drag crisis. In the supercritical regime, stabilizes at higher values near 0.25 but with broader scatter, underscoring the more chaotic, less periodic shedding influenced by turbulent structures. Mathematical models of vortex shedding, often based on two-dimensional Navier-Stokes assumptions, accurately capture periodic behavior in the subcritical regime but encounter limitations at higher Re where three-dimensional instabilities dominate. For instance, oblique shedding modes and spanwise variations emerge around Re ≈ 1000, invalidating 2D approximations; in critical and supercritical flows, turbulent boundary layers and 3D wake interactions further necessitate advanced three-dimensional simulations such as large-eddy simulation (LES) or direct numerical simulation (DNS) for reliable predictions. These models highlight the need for regime-specific adjustments to ensure applicability beyond laminar-dominated conditions.

Engineering Impacts

Structural Vibrations and Resonance

Vortex shedding generates fluctuating aerodynamic forces on structures immersed in flows, primarily through periodic variations that induce oscillatory and . These forces manifest as transverse (perpendicular to the ) due to alternating fluctuations and inline (parallel to the ) from variations, with the component often dominating in bluff bodies like cylinders or bridge decks. The lock-in phenomenon occurs when the vortex shedding frequency, often characterized by the , synchronizes with the structure's , leading to amplified vibration amplitudes through negative aerodynamic . This , typically within a reduced velocity range where the frequency ratio falls between approximately 0.75 and 1.05, results in a where the shedding pattern adjusts to match the structural motion, such as transitioning between 2S (two single vortices per cycle) and 2P (two pairs per cycle) modes. Such vibrations impose cyclic loading on structures, accumulating stress over time and risking material failure in components like girders, towers, and cables, particularly in low mass-damping systems where amplitudes can reach up to one . Prolonged exposure exacerbates crack initiation and under these resonant conditions, compromising structural integrity without immediate catastrophic collapse. To quantify these effects, accelerometers are employed to measure structural and , capturing and responses during testing, while sensors distributed along the surface detect fluctuating loads to compute integrated and forces. These instruments enable phase-averaged analysis to correlate flow-induced pressures with patterns, validating models in both and settings.

Notable Historical Examples

One of the most iconic incidents involving vortex shedding is the collapse of the Tacoma Narrows Bridge on November 7, 1940, in Washington State, United States. While the primary mechanism of failure was aeroelastic flutter leading to torsional oscillations, vortex shedding contributed to the initial vertical and lateral motions by generating alternating aerodynamic forces on the slender bridge deck as wind passed through the narrow strait. Investigations confirmed that the shedding frequency locked into the bridge's natural modes, amplifying early vibrations before flutter dominated. Vortex shedding has historically induced aeolian vibrations in overhead power lines and cables, causing persistent low-amplitude oscillations that produce audible humming tones and lead to conductor fatigue over time. These vibrations, driven by wind speeds typically between 1 and 7 m/s, prompted the invention of Stockbridge dampers in the 1920s by engineer George H. Stockbridge at to dissipate energy and prevent failures in transmission infrastructure. Patented in 1928, these tuned mass dampers remain a standard solution for suppressing such vortex-induced effects in electrical grids worldwide. In industrial applications, vortex shedding has led to notable structural failures in flare stacks at oil and gas facilities, where slender cylindrical booms experience resonant vibrations that cause cracks and collapses. Wind tunnel studies on flare boom members have highlighted how these vibrations amplify under specific Reynolds numbers, contributing to historical incidents in refineries. A prominent example occurred in when one of the three 81-meter towers of the VertiGo slingshot ride at amusement park in collapsed during the off-season; without the ride capsules providing stabilization, wind-induced vortex shedding caused oscillatory in the unsupported . Since the early 2000s, vortex shedding has been positively exploited for energy harvesting via vortex-induced vibrations, transforming a destructive phenomenon into a renewable power source. Pioneering work, such as the VIVACE (Vortex Induced Vibration Aquatic Clean Energy) system developed by researchers at the University of Michigan, demonstrated in 2008 the conversion of low-velocity water flows into mechanical energy using oscillating cylinders, achieving efficiencies up to 30% in controlled tests. Subsequent advancements include piezoelectric harvesters for wind applications, where a 2019 study showed a miniature MEMS device generating power in the nanowatt range at wind speeds around 4-6 m/s by leveraging VIV lock-in on a bluff body. As of 2024, studies have explored tandem square bluff bodies to enhance energy harvesting efficiency through optimized spacing ratios in VIV setups. These innovations have expanded to cable-based systems, enabling scalable energy capture from ocean currents and urban winds without traditional turbines.

Mitigation Techniques

Flow Alteration Methods

Flow alteration methods aim to disrupt the coherent formation of vortices in the wake of bluff bodies, such as circular cylinders, by modifying the incoming or separating fluid flow to prevent the organized Kármán vortex street. These passive techniques target the upstream or boundary layer flow characteristics to reduce vortex shedding intensity without relying on active energy input, thereby mitigating associated structural vibrations in engineering applications like chimneys, bridges, and offshore risers. By breaking flow symmetry or delaying separation, these methods can significantly lower drag and vibration amplitudes, often achieving suppression rates of 70-90% under relevant Reynolds numbers. Helical strakes, also known as Scruton strakes, consist of triangular fins helically wrapped around the structure's to induce three-dimensional perturbations that disrupt spanwise of vortex shedding. Typically, these strakes have a height of approximately 10% of the and a pitch of five times the , promoting continuous rotation of the wake and preventing coherent vortex pairing. Their effectiveness in breaking symmetry was first demonstrated in 1957, where they reduced wind-induced oscillations on circular-section structures by altering the shear layer roll-up process. Modern implementations confirm that such configurations suppress vortex-induced vibrations (VIV) by up to 90% while increasing drag by 20-50%, making them suitable for long-span applications. Fairings and spoilers streamline the body geometry or introduce localized disruptions to minimize and weaken the that drives vortex formation. Fairings, often in bullet or teardrop shapes, enclose the ends or sections with a tapered profile that aligns the flow, reducing the wake width and suppressing shedding by maintaining attached s over longer distances. Bullet fairings, with a streamlined and blunt tail, are particularly effective for marine risers, achieving near-complete VIV elimination in cross-flow while significantly reducing by up to 60% compared to bare . Teardrop fairings, featuring a more elongated taper, further delay separation by optimizing the recovery, though they may introduce minor spanwise instabilities if not properly flanged. Spoilers, as fixed protrusions like plates or wedges, actively trip the to reattach flow or deflect vortices, providing targeted suppression in high-velocity regimes. Surface modifications, such as dimples or controlled roughness elements, alter the from laminar to turbulent, energizing the near-wall to delay separation and disrupt the coherent vortex shedding pattern. Dimples, analogous to those on balls, create localized low-pressure zones that promote earlier , reducing the wake deficit and suppressing VIV amplitudes at moderate Reynolds numbers. Roughness coatings, applied uniformly or in strips, increase and induce small-scale that scatters vortex formation, effectively shortening the correlation length of shed structures. These modifications are lightweight and cost-effective for , though optimal roughness height (typically 1-5% of ) depends on conditions to avoid excessive penalties. Perforated shrouds, essentially porous screens encircling the body at a small standoff distance, diffuse the oncoming through openings to weaken incoming and attenuate wake unsteadiness. These cylindrical meshes, with ratios of 40-60%, act as a diffuser that broadens the profile and reduces gradients, suppressing vortex shedding by desynchronizing the layers and diminishing fluctuations by up to 80%. The pattern influences effectiveness; uniform holes promote even diffusion, while non-uniform designs can target specific shedding modes. Shrouds are advantageous for flexible structures like risers, as they maintain suppression across varying angles of attack without significant .

Damping and Control Strategies

Damping and control strategies for vortex shedding focus on dissipating the vibrational energy transferred to structures after vortices form, thereby preventing and without directly modifying the flow field. These methods typically involve mechanical or material-based absorbers that target the structural response, converting into or counteracting oscillations through tuned dynamics. Such approaches are essential for slender structures like chimneys, bridges, and cables, where vortex-induced vibrations (VIV) can amplify displacements significantly during lock-in conditions. Tuned mass dampers (TMDs) consist of a attached to the primary structure via a spring and damper, tuned to the structure's to counteract VIV-induced oscillations. In , TMDs have been effectively deployed to mitigate from vortex shedding; for instance, a TMD installed in a 183 m industrial reduced excessive responses during coexistence with a taller adjacent structure, limiting accelerations to safe levels. For long-span bridges, multiple TMDs optimized for multi-mode VIV control can suppress deck displacements by up to 70% under wind loads, as demonstrated in analytical studies comparing TMDs to inerter-enhanced variants. These passive devices excel in damping but require precise tuning based on to achieve optimal energy dissipation. Stockbridge dampers, featuring a flexible messenger cable clamped to the structure with counterweights at each end, are widely used to suppress aeolian vibrations in overhead power lines caused by at low wind speeds (typically 1-7 m/s). Invented by George H. Stockbridge in 1925 and patented in 1928, these dampers operate on the principle of tuned absorption, where the masses oscillate out of with the to dissipate through hysteretic in the cable. Design principles emphasize impedance matching between the damper and to maximize power dissipation across resonant frequencies (3-150 Hz), with optimal placement at 70-80% of the span length; experimental validations show they reduce amplitudes by 50-90% in critical modes. Modern variants incorporate asymmetric masses for broader frequency coverage, maintaining effectiveness in preventing fatigue failures at suspension clamps. Active control systems employ from to drive actuators that apply counter-forces opposing VIV displacements, offering adaptability for varying conditions. Emerging prominently after 2010, these methods use proportional-integral (PI) controllers with windward-suction-leeward-blowing actuation on cylinders at low Reynolds numbers (e.g., Re=100), achieving complete suppression of oscillations by altering local gradients based on history. In flexible structures, linear and nonlinear controllers reduce by over 90% through opposing transverse forces, with improved by optimal integration and constraints. Such systems, while computationally intensive, provide superior in dynamic environments compared to passive alternatives, though they demand reliable power and sensing . Recent hybrid approaches combining with optimization have shown promise for enhanced efficiency as of 2023. Viscoelastic coatings and mounts enhance structural by incorporating materials with both and viscous properties, which convert vibrational into without influencing the surrounding . Applied as thin layers on cylinders or supports, these materials suppress VIV in viscoelastically mounted rigid circular cylinders by increasing the system's effective ratio, reducing peak amplitudes by 40-60% across a range of reduced velocities. For risers and pipelines, constrained viscoelastic layers mitigate seismic- and vortex-induced responses by shifting natural frequencies and dissipating through deformation, as validated in numerical models showing significant life extension. These coatings are particularly valued for their simplicity and maintenance-free operation in harsh environments.

References

  1. [1]
    Bluff Body Flows – Introduction to Aerospace Flight Vehicles
    In cross-section, a bridge deck can cause vortex shedding and oscillatory aerodynamic forces. The twisting of the bridge deck increases flow separation, ...
  2. [2]
    [PDF] FUNDAMENTALS OF VORTEX-INDUCED VIBRATION
    o To understand VIV, you must first understand the physics of a steady approach flow past a nonvibrating circular cylinder. • How and why does VIV occur? • What ...Missing: dynamics | Show results with:dynamics
  3. [3]
    [PDF] Journal of Non-Newtonian Fluid Mechanics
    Dec 12, 2023 · In fact for a Newtonian fluid, the critical Reynolds number for vortex shedding, Recrit = 47, is significantly smaller than the value observed ...
  4. [4]
    Vortex Shedding - Fedele's GATECH Website
    Vortex shedding is a phenomenon that occurs when the current flow of a body of water is disrupted by an obstruction, in this case a bluff body. When placed in a ...
  5. [5]
    Separated Viscous Flows & Vortex Induced Vibrations - MIT
    Vortices shed from bluff bodies are typically shed alternately from top and bottom. • Each time a vortex is shed there is a resultant force on the body. • ...
  6. [6]
    Drag of Blunt Bodies and Streamlined Bodies
    A streamlined body looks like a fish, or an airfoil at small angles of attack, whereas a bluff body looks like a brick, a cylinder, or an airfoil at large ...
  7. [7]
    Vortex Shedding in Water
    Water moving around a cylindrical pendulum makes the cylinder swing at the vortex shedding frequency.Missing: definition | Show results with:definition
  8. [8]
    Ueber den Mechanismus des Widerstandes, den ein bewegter ...
    Kármán, Th. von. "Ueber den Mechanismus des Widerstandes, den ein bewegter Körper in einer Flüssigkeit erfährt." Nachrichten von der Gesellschaft der ...Missing: PDF | Show results with:PDF
  9. [9]
    On the phenomenon of vortex street breakdown | Journal of Fluid ...
    Mar 29, 2006 · A von Kármán vortex street generated in the usual way was subjected to a deceleration, thereby changing the ratio of longitudinal to lateral ...
  10. [10]
    Bénard-von Kármán instability: transient and forced regimes
    Apr 21, 2006 · The wake of a circular cylinder is investigated near the oscillation threshold by means of a laser probe. Above the threshold the transient regime is studied.
  11. [11]
    [PDF] Simulation and Instability Investigation of the Flow around a Cylinder ...
    The formation of Karman vortex street originates from the combinations of the interaction of two shear layers at two lateral sides of the cylinder and the ...
  12. [12]
    Self-Consistent Mean Flow Description of the Nonlinear Saturation ...
    Aug 20, 2014 · The Bénard–von Kármán vortex shedding instability in the wake of a cylinder is perhaps the best known example of a supercritical Hopf ...
  13. [13]
    [PDF] a model for the von karman vortex street
    As von Kármán (1911) pointed out, it is sufficient to consider a parallel array of vortices extending to infinity in both upstream and downstream directions.Missing: Theodore | Show results with:Theodore
  14. [14]
    An Observational Study of Vortex Spacing in Island Wake Vortex ...
    Vortex streets behind bluff bodies have been studied theoretically using inviscid stability analysis (von Kármán and Rubach 1912) and direct experimentation ( ...
  15. [15]
    [PDF] on the development of turbulent wakes from vortex streets
    The Reynolds number range of periodic vortex shedding -is divided into two distinct subranges. At R=40 to 150, called the stable range, regular vortex ...
  16. [16]
    [PDF] Vortex Shedding and Vortex Induced Vibrations - MIT
    Frequency of vortex shedding f = ω/2π is given by a non-dimensional number. where f is the Strouhal frequency, D is the body diameter and S is the Strouhal ...
  17. [17]
    Strouhal-Reynolds Number Relationship for Vortex Streets
    Aug 18, 2004 · Abstract. A rationale for the empirically observed Strouhal-Reynolds number relation for vortex shedding in the wake of a cylinder is provided.
  18. [18]
    Vortex Dynamics in the Cylinder Wake - ResearchGate
    Aug 6, 2025 · This aligns with the findings of Williamson, 43 where higher Re regimes (Re ¼ 150) trigger secondary instabilities that disrupt span-wise ...
  19. [19]
    Fluctuating lift on a circular cylinder: review and new measurements
    The present work will focus on the unsteady cross-stream force, the fluctuating lift, acting on a single circular cylinder in cross-flow.
  20. [20]
    On the flow past a circular cylinder from critical to super-critical ...
    Large-eddy simulations (LES) of the flow past a circular cylinder are used to investigate the flow topology and the vortex shedding process at Reynolds numbers.
  21. [21]
    Simulation of the flow past a circular cylinder from sub-critical to ...
    In this work, we investigate the prediction of the flow around a circular cylinder from sub-critical to super-critical Reynolds numbers using a hybrid approach.<|control11|><|separator|>
  22. [22]
    [PDF] Effect of Reynolds number and a low-intensity freestream turbulence ...
    Fig. 3a Strouhal number vs Reynolds number for Tu = 0.1%. Freestream velocity for D = 120 mm corrected for blockage (5.8%).
  23. [23]
    None
    Summary of each segment:
  24. [24]
    An overview of vortex-induced vibration (VIV) of bridge decks
    A critical literature review of VIV of bridge decks that highlights physical mechanisms central to VIV from a renewed perspective is provided.
  25. [25]
    Field monitoring and validation of vortex-induced vibrations of a long ...
    Aug 6, 2025 · Although VIV does not lead to catastrophic damage to bridges similar to flutter, prolonged VIV endangers pedestrian comfort, traffic safety, and ...
  26. [26]
    Tacoma Narrows Bridge history - Bridge - Lessons from failure
    3. Also contributing to the torsional motion of the bridge deck was "vortex shedding." In brief, vortex shedding occurred in the Narrows Bridge as follows: (1) ...First investigations - Partial... · Blind spot" - Design lessons of... · End of an era
  27. [27]
    [PDF] Resonance, Tacoma Narrows bridge failure, and undergraduate ...
    It can be concluded that natural vortex shedding was not the cause of the collapse. This rules out one type of periodic exciting force implied by a few of ...
  28. [28]
    Vertical and torsional vibrations before the collapse of the Tacoma ...
    The vertical vibration occurs from the frequency lock-in with the vortex shedding, and its wavelength and frequency agree well with the recorded data in 1940.
  29. [29]
    [PDF] Modelling and assessment of aeolian vibrations of overhead ...
    Aeolian vibration can be controlled by dampers. The most popular type of transmission line damping device is the Stockbridge damper, patented in 1925 by George ...
  30. [30]
  31. [31]
    Materials Give Roller Coaster Enthusiasts a Reason to Scream
    The cause of the crack has been found to be vortex shedding. The VertiGo was a slingshot ride that consisted of three 81 meter columns, made of carbon steel, ...
  32. [32]
    Cedar Point removes thrill ride that collapsed - Toledo Blade
    Mar 7, 2002 · Two months after part of Cedar Point's VertiGo thrill ride collapsed, the park's parent company has decided to remove the giant triple tower rather than repair ...
  33. [33]
    Vortex-induced vibration wind energy harvesting by piezoelectric ...
    Dec 31, 2019 · Here, we investigate the use of vortex-induced vibration (VIV) for miniature wind energy harvesting. In particular, we consider the use of ...
  34. [34]
    Optimal energy harvesting from vortex-induced vibrations of cables
    Nov 1, 2016 · Vortex-induced vibrations (VIV) of flexible cables are an example of flow-induced vibrations that can act as energy harvesting systems by ...Missing: seminal | Show results with:seminal
  35. [35]
    Performance of two- and three-start helical strakes in suppressing ...
    Oct 15, 2018 · Helical strakes can suppress VIV well when its pitch ratio is greater than 10 and height is greater than 0.2 times cylinder diameter. If ...<|control11|><|separator|>
  36. [36]
    Getting Twisted about Helical Strakes | Article | Meca Enterprises
    Jun 25, 2019 · Can I deviate from the Standard? The standard for fabricating strakes is as follows: Strake Width = 0.1 * Stack Diameter Strake Pitch = 5 * ...Missing: dimensions | Show results with:dimensions
  37. [37]
    [PDF] Suppression of vortex-induced vibration of a cylinder with helical ...
    The pitch of helical strakes has less impact on the suppression efficiency;. 3. The height of the helical strakes has a greater influence in the VIV response.
  38. [38]
    [PDF] Fairing Shape: The Pros and Cons of Popular Fairing Geometries
    Aug 21, 2018 · U-shaped fairings sometimes have flanges at each end to reduce spanwise correlation of vortex shedding that can occur due to the wide wake.Missing: bullet | Show results with:bullet
  39. [39]
    Suppression of Vortex-Induced Vibration by Fairings on Marine Risers
    Jan 24, 2020 · Vortex-induced vibration of the offshore risers could be effectively suppressed by fairing devices. In this paper, a newly developed vortex- ...
  40. [40]
    Spoiler plate effects on the suppression of vortex-induced motions of ...
    Aug 15, 2020 · The results demonstrated the occurrence of a desynchronization of the vortex shedding, which led to a significant reduction of the in-line and ...
  41. [41]
    Passive control of the onset of vortex shedding in flow past a circular ...
    Jan 10, 2020 · Surface modification with roughness, dimple, helical wire ... modifications are made to control vortex shedding and suppress VIVs.
  42. [42]
    [PDF] Wake Stabilization and Force Modulation via Surface Dimples on an ...
    Jul 31, 2025 · 6, indicating that dimples influ- ence vortex shedding by strengthening the vortex cores and delaying vortex breakdown in the wake. To ...
  43. [43]
    [PDF] Suppression of vortex shedding around a square cylinder using ...
    A variety of passive methods that use the surface modifications are roughness, dimples. (Choi 2006), helical strakes (Lee & Kim 1997), longitudinal grooves ...
  44. [44]
    Effect of Surface Roughness on Aerodynamic Loads of Bluff Body in ...
    Mar 29, 2024 · In fluid–structure interactions, surface roughness (passive technique) can control the dynamics of vortex shedding by inducing suppression. The ...
  45. [45]
    Suppression of vortex shedding and vortex-induced vibration of a ...
    The screen shrouds can significantly suppress VIV amplitudes, delay the onset of the VIV and narrow the lock-on region.
  46. [46]
    Vortex shedding control of circular cylinder by perforated shroud in ...
    May 12, 2017 · The aim of the present study is to control the vortex shedding downstream of a circular cylinder (inner cylinder) by the existence of outer ...
  47. [47]
    On the Effectiveness And Mechanism of Vortex-induced Vibration ...
    Jun 17, 2012 · It was shown that the screen shroud can suppress VIV by about 50%. In exploring the mechanism on VIV reduction, two sets of experiments using ...
  48. [48]