You are on page 1of 42

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Vortex Shedding Induced Loads on Free Standing Structures


Author: I. Giosan, P.Eng. Introduction Estimating the wind-induced response of structures is a main topic in the area of the wind engineering research over the last 40 years [1]. Analyses are usually carried out through theoretical formulations, numerical algorithms, wind tunnel tests, full scale experiments and code provisions. A dominant property of modern wind engineering is that accurate analyses of complex problems are seldom possible by using one of these methods and instead they require the joint application of different techniques For instance, wind tunnel and full-scale experiments frequently provide the input for theoretical solutions, numerical simulations and code provisions; theoretical and numerical methods are often the bases for organizing developing and processing measurements. The growing importance of wind effects on structures, joined with theoretical, computational and instrumental advances, has given great impulse to these techniques; in the meanwhile, the evolution of each technique is linked and even more inspired by the growth of complimentary methods. All these powerful analytical and simulation tools are required when investigating complex wind induced phenomena like vortex shedding. Vortex Shedding (Fig. I.1) is the instance where alternating low pressure zones (blue colors) are generated on the downwind side of the stack, as shown in this figure.

Fig. I.1 Vortex Shedding phenomenon induced by wind flowing over a cylinder. These alternating low pressure zones cause the stack to move towards the low pressure zone, causing movement perpendicular to the direction of the wind. When the critical wind speed of the stack is reached, these forces can cause the stack to resonate where large forces and deflections are experienced. Every stack has a critical wind speed at which vortex shedding occurs. Vortex shedding can be observed also at larger scale as a meteorological phenomenon. In [Fig. I.2] is shown an image taken from space of the vortex shedding phenomenon behind the island Juan Fernandez along the Chilean rim.

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Fig. I.2 Vortex Shedding as a meteorological phenomenon. Vortex shedding is a complex physical phenomenon, especially when it degenerates into lockin condition. As large vibrations may occur at moderate and frequent wind velocities, structures may undergo a great number of stress cycles that lead to damage accumulation and may determine structural failure without exceeding the ultimate limit stress. Considering the potential vortex shedding fatigue induced damage it is very important the design procedures to account in a realistic manner for the vortex shedding dynamic induced loads. Although an immense analytical and experimental effort has been made during the last fifty years to improve the analytical vortex shedding prediction models, the design standards are still lacking in presenting concise and easy to use analytical methodologies [3-13]. Because of the complexity of vortex shedding phenomenon at this moment there is no general analytical method available to calculate the response of the structure to the vortex shedding dynamic induced loads [2]. Avery detailed comparison on wind induced response of major codes and standards was made and presented by Working Group E Dynamic Response set up at the First International Codification Workshop held in Bochum, Germany on September 15th 2000 [11]. A comprehensive comparison of the along wind loads and their effects on structures was conducted on the major international codes and standards: the US Standard ASCE 7-98, 200 [8], the Australian Standard AS1170-2, 1989 [9], the National Building Code of Canada NBCC, 1995 [3], the Arhitectural Institute of Japan AIJ-RLB, 1993 [10] and the European Standard Eurocode ENV1991-2-4 [6]. The comparison study concluded that there are considerable scatter in predictions among these standards and is very important to understand the underlying differences in order to develop unified international codes. An historical overview in the research and simulation areas of vortex shedding phenomenon is presented by Giovanni Solari [1].

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Some of the most important pioneering research contributions in the area of wind dynamics have come from the University of Western Ontario (UWO) in Canada and the Technical University Aachen (RWTH) in Germany. The Canadian research team was lead by Dr. Alan G. Davenport since 1965 when he founded the Boundary Layer Wind Tunnel Laboratory. This laboratory has carried out innovative design studies for major structures. For example, of the 40 tallest buildings in the world, roughly two thirds have been studied at Western Ontario Boundary Layer Wind Tunnel Laboratory. Dr. Davenport has pioneered in the application of boundary layer wind tunnels to the design of wind sensitive structures. The mathematical spectral model proposed by the Canadian research group is used as the basis for the Canadian National Building Code [3]. It gives accurate results for relatively stiff structures, such as concrete chimneys, in which the aeroelastic effects are relatively small or moderate. For more flexible structures like steel tubular towers, antenna towers, transmission or highmast towers some of the Canadian research team results are incorporated into Canadian Bridge Design Code [4,5]. The German team research is focused on accurate modeling of large aeroelastic effects and it is incorporated into Eurocode [6]. As mentioned the study of dynamic response of vertical structures to vortex shedding begun in the 60s thanks to the pioneering contribution of Davenport. In his studies [16,17] he defined the main parameters necessary to analyze the structural wind actions: mean static wind pressure, mean alongwind static displacement, gust response factor and equivalent static force. Studies developed in the 70s by Vickery [18] and Simiu [19] perfected the method introduced by Davenport especially with respect to wind and aerodynamic modeling. On the other hand ESDU [20] and ECCS[21] introduced procedures to determine the maximum values of structural effects by using influence function techniques. Starting from Simius formulation, at the beginning of 80s, Solari derived closed for solutions for alongwind response of structures. He also developed the Equivalent Spectrum Technique [22], a method that schematizes the wind as an equivalent velocity field perfectly coherent in space. Research carried out in 90s followed two distinct lines. The first was aimed at determining the maximum effects due to alongwind response. The second extends original methods from alongwind response to crosswind and torsional responses. The first research line derives from the observation that Equivalent Static Force (ESF), as conceived by Davenport and used by most subsequent authors, usually produces correct mean maximum displacements but may give place to other effects like bending moments and shear forces. Efforts were made by Kaspersky [23] and Holmes [24,25] do develop new analytical methods for defining the ESF in a more realistic and simplified [26] manner. The second research line was aimed at determining the three-dimensional wind induced response of structures. A wide research program was carried out in Japan [27] to derive the maximum alongwind, crosswind and torsional response of structures and their related ESF, by fitting the results of wind-tunnel tests. A study focused on slender structures and structural elements was initiated and coordinated by Giovanni Solari at Department of Structural and Geotechnical Engineering from University of

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Genoa in Italy. This study extended the concepts defined by Davenport for alongwinds to crosswinds and torsional responses [28,29]. Alongwind, crosswind and torsional actions were defined as a linear combination of the longitudinal, lateral and vertical turbulence components by quasi-steady theory; crosswind forces and torsional moments due to vortex shedding were superposed by considering these as independent of turbulence actions [30]. Alongwind, crosswind and torsional responses, analyzed as uncoupled and only dependent on the related fundamental modes of vibration, were determined in closed forms [31,32] by the generalized spectrum technique [33] . This solution was generalized to slender vertical structures with localized masses and linearized aeroelastic terms [33]. In the recent years the researchers were focused on refining the analytical models to predict accurate results for vortex shedding at high Reynolds numbers [34], to account for mode shape corrections [35] or to account for fatigue induced by vortex shedding phenomenon [36]. The most extensive and important experimental full scale tests to investigate the response of structures to cyclic loads were done at University of Minnesota, Minneapolis, USA [37], and the experimental program was focused on fatigue-resistant design on cantilevered signal sign and light supports. Experimental observations and numerical modeling of the vortex shedding phenomenon linked to structural fatigue were also the focus of the research program lead by dr. Christopher Foley at University of Marquette, Wisconsin [38]. Experimental observations, full scale tests and advanced nonlinear numerical modeling of the dynamic induced loads on free standing structures are also a part of the research program initiated and conducted by Ioan Giosan at West Coast Engineering Group Ltd., Vancouver, Canada. The research program is aimed to develop standard design methodologies and standard numerical modeling procedures accounting for dynamic induced loads. The sources of the dynamic induced loads are simulated realistically, as a result of the impact between the structure and the surrounding environment [39,40].

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

1. Vortex Shedding Phenomenon When the wind blows across a slender structural member, vortices are shed alternately from one side and then the other, giving rise to a fluctuating force acting at right angles to the wind direction. This organized pattern of vortices is referred to as von Karman vortex street. A structural member may be considered slender in this context if the aspect ratio exceeds 20 [3]. For lightly damped members, which are free to oscillate, large amplitude vibrations in the plane normal to the wind may develop when the vortex shedding is in resonance with one of the natural frequencies of vibration. Although this is most likely to occur for the lower modes of vibration, vortex shedding induced effects for very flexible members may also be important for higher modes. The character of the vortex shedding forces for circular cylinders depends on the Reynolds number, VD Re =

where: V= wind speed [m/s]; D= structure diameter [m]; = kinematic viscosity [m2/s ]. The shedding tends to be organized at sub-critical and trans-critical Reynolds numbers [2,3,4,5]. Sub critical Range : Re < 3 x10 5

Critical Range : 3x10 5 Re 3x10 6 Trans critical Range : Re > 3x10 6 In the critical range, namely for: 3x105 Re 3x106 vortex shedding tends to be irregular unless the structural motion is sufficiently large to organize the fluctuating flow around the body. This phenomenon, referred to as locking in, becomes important for lightly damped members. When the lock-in phenomenon occurs, in some cases, severe vibration has persisted long enough to cause fatigue cracking or structural failure (Fig.1.1).

Fig. 1.1 Antenna Tower Fatigue FailurePictures provided by Morrison Hershfield consulting company.

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

In free standing tubular structures (transmission, telecommunication, highmast or street light poles), cracks are usually observed at the welded connection between the vertical shaft and base plate. If there are stiffeners or gussets reinforcing the pole to base plate connection, then the cracks will typically form at the tops of the stiffeners. If there are hand holes, cracks may appear around the perimeter of these. Cracked anchor rods have also occurred. Vortex shedding tends to occur with steady continuous winds at a critical velocity. The velocity need not be very high, but it has been found that significant vibration does not occur unless the velocity is greater than 5 m/s. The periodic frequency of the vortex shedding can lock in on the natural frequency of the pole, resulting in very large alternating forces acting transverse to the wind flow direction. Occasionally, the vibration is so severe that fatigue cracks will appear. Although vortex shedding can "lock in" and continue as the velocity increases or decreases slightly, if the velocity changes by more than 20 percent, the vortex shedding will stop. Gusty variable winds, such as might occur in a severe storm typically will not cause vortex shedding. In fact, if the wind velocity is greater than 15 m/s (35 mph), the wind is generally too turbulent for vortex shedding to occur. In summary, the winds that are dangerous for vortex shedding are steady winds in the velocity range 5 to 15 m/s [37]. Recent studies [38] have verified that vortex shedding can occur in tapered as well as prismatic circular poles with almost any diameter. Although nearly periodic in a smooth air stream, vortex shedding in turbulent boundary layer flow conditions, which is characteristic of natural wind, tends to become less regular, with energy distributed over a band of frequencies around e. e = frequency at which vortex shedding occurs at a specific location [Hz]. The presence of turbulence effectively reduces the extent of the member over which the vortex shedding forces remain correlated [3,4,5]. A reduction of the aspect ratio has somewhat similar effects [5]. Several measures may be considered should vortex shedding induced effects prove to be excessive. these include: - a strengthening or stiffening of the member, or both; - increasing the mass; - increasing the damping; and - changing the aerodynamic characteristics for example, by increasing the taper, or by the addition of aerodynamic spoilers. Of these alternatives, increasing the damping of the member is the most desirable solution. The effective damping can be increased using visco-elastic materials or special dynamic absorbers [5].

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

2. Vortex Shedding Induced Loads - Analytical Models.

The excitation due to vortex shedding is treated as a time-varying load of frequency: (2.1) SV e = D where: S= the Strouhal Number for the cross-sectional geometry (Table 2.1).Resonant vibrations are assumed to occur when the frequency of vortex shedding coincides with a natural frequency of the member. The evaluation of vortex shedding-induced effects requires a dynamic analysis of the member to determine its natural frequencies and associated mode shapes of vibration. All modes of vibration for which vortex shedding-induced resonant vibrations occur at wind speeds equal or less than that corresponding to the design mean hourly reference wind pressure V, must be considered [4,5].
V 1.24(qC e )
0.5

(2.2)

Where: q=hourly mean reference wind pressure for the design return period [Pa]; Ce=wind exposure coefficient (Table 2.3). In the case of a member with a constant cross-section, resonant vibrations for a particular mode of vibration with natural frequency, ni, occur at a specific or critical wind speed: (2.3) D i Vcri = S In the case of a tapered member, the frequency of vortex shedding at a particular wind speed varies over the length of the member. As the wind speed increases, resonant excitation occurs first at the smaller diameter portion of the member and then shifts to portions with larger diameter. Consequently, vortex shedding effects associated with a particular mode of vibration with frequency, i, must be examined for a range of critical wind speeds. Defining Dmin and Dmax as the minimum and maximum cross-sectional diameters, respectively, this range is expressed as: (2.4) i Dmin D Vcr i max i S S Considering an harmonic sinusoidal model, the root mean square (RMS) time-varying vortex shedding induced load acting at a particular location, x, along a member is expressed as:
FS ( x, t ) = where:

V 2 C L D( x) sin[2 e ( x)t ]
2

(2.5)

= air mass density, taken as 1.29 [kg/m3]; V= the mean wind speed at location x, [m/s]; CL= RMS lift (across-wind) force coefficient for the cross-sectional geometry a specified in Table: 2.1; x= coordinate describing length along the member, [m]; D(x)= the diameter of frontal width of a member at location x, [m]; e(x)= frequency at which vortex shedding occurs at location x, [Hz]; t= time, [s].

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

For members with constant diameter or frontal width, the magnitude of the excitation is taken to be invariant with x and proportional to the velocity pressure at the top of the member [4,5]. This is a conservative assumption since V approaches zero at the ground level. It is also conservative to treat FS(x,t) as a spatially coherent excitation, that is, acting in phase along the entire length of the member (Fig. 2.1 a). In reality, this begins to occur only at large amplitudes of motion.

a) Non-tapered structure

b) Structure with taper, p

Fig. 2.1 Typical Vortex Shedding-Induced Responses in Two Modes of Vibration The variation of the wind speed with height, the turbulent flow functions normally experienced and the presence of signs and other accessories all tend to disrupt the spatial correlation of the excitation. For structures with varying diameter or frontal width, the magnitude of the excitation will vary along the length of the member. The vortex shedding excitation at location x1, is taken to remain in phase over the portion of the member for which the diameter of frontal width remains within percent of D(x1) and to be zero over the remainder of the member (Fig. 2.1 b) A default value for =10% is prescribed in Canadian Highway Bridge Design Code [4.5] and in the National Building Code of Canada [3]. The value of =10% is applicable for peak response amplitudes of greater than 2% of D.

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

For a band limited random forcing model with a Gaussian load spectrum, the induced load is described by Davenport et al. [17], Harris and Crede [41], Vickery and Clark [42,43]. The band limited random forcing model differs from the sinusoidal model by: - Allowing for a random (rather than harmonic) vortex lift force. This employs a different forcing function than the sinusoidal model. - Allowing for the energy associated with the vortex shedding to be distributed about the dominant frequency (rather than concentrated on the dominant frequency). This employs a bandwidth term, B, which is a measure of the distribution of the energy (Table 2.1). - Allowing for the three-dimensional nature of the flow the loss of correlation of the lift forces along the length of the member. This employs a term for the correlation length, L, which is a measure of the length, in diameters, that the vortices remain in phase (Table 2.1). - Allowing for the turbulence of natural wind (turbulence leads to reduction in vortex shedding correlation length and leads to a small reduction in the strength of the shedding forces). - Allowing for the variation of wind speed with height. This employs the power law wind-velocity-profile-exponent, , to obtain an apparent taper between the wind and the member. Treating vortex shedding as a sinusoidal process is an approximation leading to conservative estimates. The variation of the wind speed with height, turbulence of the natural wind and the presence of signs and other accessories all tend to disrupt the spatial correlation of the excitation. It is generally accepted to be more accurate to treat the excitation as a band-limited random process and to assume that the forcing tends to become harmonic only when the motion of the members is sufficiently large to organize the shedding vortices [4,5,6,7]. This tends to occur when the peak amplitude of the motion is of the order of 2% to 2.5% of the diameter or the width of the cross-section, and greater. For the evaluation of the vortex shedding-induced response in a particular mode of vibration, the direction of the vortex shedding excitation at any location, x, is taken in the direction of the motion of the member at that location. This is a simplification as the direction of the vortex shedding force at large body amplitudes is more likely to be in the direction of the local time derivative or velocity of the body motion. For the purpose of evaluating the generalized force, GFi, associated with a particular mode of vibration, both assumption lead to the same RMS and peak values.

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

2.2.1 Vortex Shedding Response Parameters

In table 2.1 are presented values for vortex shedding parameters, for various shapes and Reynolds numbers. Each parameter will be defined, and will be also presented values or empirical computation relations.
Table: 2.1 Vortex Shedding Parameters Cross-section Circular Subcritical Re<3x105 Supercritical and transcritical Re>3x105 Square Multisided members, rolled structural shapes S CL B L

0.18 0.25 0.11 0.15

0.50 0.20 0.60 0.60

0.10 0.30 0.25 0.25

2.5 1.0 3.0 2.75

Where: S= the Strouhal Number for the cross-sectional geometry. CL= RMS lift (across-wind) force coefficient for the cross-sectional geometry. B= band width, a measure of the variability of the vortex shedding frequency. L= correlation length, the length (as a ratio of the diameter) over which the vortices act in phase. Strouhal Number - S can be calculated also using the methodology outlined in Canadian National Building Code [3]. For circular and near-circular cylinders, the Strouhal number, S, is a function of the Reynolds number, Re. Although the Reynolds number is a function of V, a trial-and-error approach [3] to finding the critical mean speed can be avoided by examining the product, D2, and using the appropriate version of equation (2.6) as follows:
D 2 0.5m2 s, Vcr = 6D if 0.5 m2 s D 2 0.75m2 s, Vcr = 3D + (1.5 m2 s ) D 2 2 D 0.75m s, Vcr = 5D

(2.6-1) (2.6-2) (2.6-3)

Equation (2.6-1) applies when Re<2x105 and S=1/6. Equation (2.6-2) covers an intermediate region where, for computational convenience, S is taken to increase approximately linearly as Re increases to 2.5x105. Equation (2.6-3) usually governs, in which Re>2.5x105 and S=1/5. The Lift Coefficient- CL has a very random nature which makes the evaluation of vortex shedding response very difficult [38] There have been studies attempting to define lift coefficients for chimneys and stacks in both steel and concrete [45] or a general analytical model [46]. Difficulty in defining the lift coefficient is enhanced by the fact that vortices tend tom have Correlation Lengths (L) over which vortices are assumed to act in phase [2,45].

10

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

There have been several simplified approaches that can be used for determining the lift coefficient and correlation lengths for vortex shedding response. The Canadian Highway Bridge Design Code (CAN/CSA-S6-2000) [4,5] has a relatively concise approach to consideration of vortex shedding deformations and the corresponding stresses. This standard [4.5] suggests for CL and L the values presented in Table: 2.1. B= Frequency Band Width. It is assumed that there is a range of wind speeds over which the vortex shedding frequency and the structures natural frequency lock-in with one another and result large across-wind vibrations. The difficulty with lock-in phenomenon arises from finding the range of natural frequency of the structure, for each mode shape, over which lock-in can occur [2]. It has been recommended that the tendency of vortex shedding to cause across-wind vibrations be computed using a range of frequencies around the shedding frequency and comparison to modal vibration frequencies of the structure. Mathematically, this can be described as [2,38], (2.7) iL i iU Where: i is the ith natural frequency of the structure; Li is the ith natural frequency of the structure lower bound; Ui is the ith natural frequency of the structure upper bound; These values are calculated using Vz' SVz SVz' ' L i = i i = = i 1 Dz Dz Vz
Vz' SVz SVz' = i + = + = i 1 + Dz Dz Vz
U i ' i

(2.8)

(2.9)

Where: Vz is the mean wind speed at elevation z; Vz is the wind speed turbulent component at elevation z; If the structures natural frequency for mode of vibration i lies within these lower and upper bounds, vortex shedding could excite mode of vibration i. The turbulent component of the wind speed is often assumed to be zero-mean random variable with Gaussian distribution [2,45] and its magnitude depends upon height above earths surface. As a result a roughness length, z0, is often utilized to estimate this effect. There have been many experimental attempts to categorize the surface roughness parameter based upon ground exposure. The ground surface roughness parameter describes the frictional effect of the underlying surface geometry. It is also described as the surface momentum sink for the atmospheric flow [11]. It is well defined only locally for homogeneous terrain and neutral atmospheric conditions. Table.2.2 shows the best estimations of roughness as depicted by Wieringa [47] and is a reasonable basis for determining appropriate roughness lengths for mainland areas.

11

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Table: 2.2 Roughness Lengths of homogeneous surfaces Surface Type Roughness Length (m) Sea, loose sand, and snow 0.0002 Concrete, flat desert, tidal flat 0.0002-0.0005 Flat snow field 0.0001-0.0007 Rough ice field 0.001-0.012 Fallow ground 0.001-0.004 Short grass and moss 0.008-0.03 Long grass and heather 0.02-0.06 Low mature agricultural crops 0.04-0.09 High mature crops (grain) 0.12-0.18 Continuous bush land 0.35-0.45 Mature pine forest 0.8-1.6 Tropical forest 1.7-2.3 Dense low buildings (suburbs) 0.4-0.7 Regularly-built large town 0.7-1.5

Typical roughness coefficient magnitudes for various terrain topologies are also given by Dyrbye and Hansen [2]:
0.01m for open land with little vegetation and houses; z 0 = 0.05m for agricultural areas with few houses and wind breaks; 0.3m for villages and agricultural areas with manny wind breaks; Surface roughness can be also determined empirically. Lettau [48] formulated z0 from field experiments as (2.10) HS z 0 = 0.5 A Where: H is the average height of obstacles within the area A. S is the total projected frontal area of the obstacles. A - is the surface area. The 0.5 factor corresponds to the average drag coefficient of the area based upon the primary obstacles projected frontal area. For a relatively large variety of S values, the equation (2.10) produces z0 that are within 25% of measured values from wind profiles. Based on the above values can be concluded that mostly of the transmission, antenna or highmast towers are located in the areas that may be considered to have few wind breaks and trees. Therefore, a roughness length of z0= 0.3m can be used for computations. Experimental results have allowed the empirical definition of the standard deviation for turbulent horizontal wind component to be determined. An expression for the standard deviation that has been validated through experiment is [2]:
12

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

V = AV f

(2.11)

Where A2.5 for a roughness length z0=0.3m. Vf is the friction velocity which can be computed with formula: Vz K Vf = (2.12) ln(z z 0 )

Where: K is von Karman constant and can be approximated as 0.4. Equation (2.11) combined with (2.11) becomes:

( V )z =

AVz K ln(z z 0 )

(2.13)

It will be assumed that the frequency boundaries given by equations (2.8) and (2.9) can be based upon the maximum turbulent wind speed variation from the mean speed, V. This maximum can be defined as Vz,max . It will be assumed that the peak variation in turbulent component from the mean will be (v) (above or below). Equations (2.8) and (2.9) can be rewritten as: Vz',max ( V )z L i = i 1 (2.14) = i 1 Vz Vz
Vz',max ( V )z = i 1 + = i 1 + Vz Vz
U i

(2.15)

where the standard deviation of the turbulent component is computed using equation (2.13) at the given height above the ground. The Reynold Number Re can be calculated for = 1.5x10-5 m2/s. VD Re = 10 5 1.5 (2.16)

Wind exposure coefficient - Ce can be estimated based on values presented in Table.2.3 [4].
Table: 2.3 Wind Exposure Coefficient Height H [m] Exposure Coefficient, Ce 0 to 10 1.0 Over 10 to 16 1.1 Over 16 to 25 1.2 Over 25 to 37 1.3 Over 37 to 54 1.4 Over 54 to 76 1.5 Over 76 to 105 1.6

13

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

An analytical formulation for estimating the wind exposure coefficient is presented in ANSI/ASCE7-2005 [8]. The Ce coefficient is calculated with formula:
z C e = 2.01 z g
2

(2.17)

Where: z height above the ground or 5m, whichever is greater; zg is a constant which varies with the exposure conditions. Based on information presented in [8] , zg should be taken 274.3m for exposure C. - is a constant which varies with the exposure conditions. Based on information presented in [8] , should be taken 9.5 for exposure C.

14

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

3. Analytical Methodologies

Before calculating the vortex shedding induced loads based on one of the analytical models presented above one has to determine the structures locations where the vortex shedding loads potentially lock in. Resonance in a particular mode of vibration with a natural frequency i occurs when e=i.
A) Members with constant diameter or frontal width

Will be calculated the wind speeds Vcri which induce vibration frequencies e close to structures natural frequency modes i: (3.1) D i Vcri = S

i = e =

SVcri D
0.5

(3.2)

Vcri 1.24(qC e )

(3.3)

The height above ground used to compute Ce shall correspond to the height above ground of location of coordinate x. The location at which e is computed shall be taken as the top of the member. All the parameters in above equations were previously defined.
B) Members with tapered diameter or frontal width.

In the case of a tapered member, the frequency of vortex shedding at a particular wind speed varies over the length of the member. As the wind speed increases, resonant excitation occurs first at the smaller diameter portion of the member and then shifts to portions with larger diameter. Consequently, vortex shedding effects associated with a particular mode of vibration with frequency, i, must be examined for a range of critical wind speeds. Defining Dmin and Dmax as the minimum and maximum cross-sectional diameters, respectively, this range is expressed as: (3.3) i Dmin D Vcr i max i S S
Vcr 1.24(qCe )
i

0.5

(3.4)

The wind speeds Vcri will induce vibration frequencies e(x) close to structures natural frequency modes i at location x: (3.5) SVcr i i = e (x ) = D( x ) e(x) shall be computed at sufficient locations along the member to determine at which location vortex shedding excitation can occur.
15

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

All the parameters in above equations were previously defined.


3.1 Analytical Methodology for the Sinusoidal Model

The time-varying across-wind or lift excitation resulting from the shedding of vortices shall be determined as follows.
3.1.1 Non-tapered Structures

The vortex shedding induced load per unit length acting at any location x along the structure shall be taken as: FS ( x, t ) =

V 2 C L D sin[2 e t ]

(3.1.1.1)

2 The maximum force for each vibration mode will be determined - FsiMax, multiplied with the value of the corresponding mode shape amplitude i(x) and applied on the structure at location x. The resulting generalized forces will be applied on the structure alternately from one side to the other with a frequency equal with i. Will be calculated stresses at each location x and compared with the endurance limit value for the material used as well as for each critical connection. For the first natural frequency mode the equivalent static force per unit length can be well approximated with this formula [3]: FS = C1 D 2 H D 1 C2 M qD (3.1.1.2)

where: M= average mass per unit length over the top one-third of the structure, [kg/m]; For most situations: 3 H D C1 = 3 for H D > 16; C1 = for H D < 16 4 C2 = 0.6 If: D 2 < C2 M then large amplitude motions up to one diameter may result [9]. If V is low, temperature gradients may produce very low turbulence levels, and in such cases vortex induced motions are significantly increased, particularly for very slender structures. If V is less than 10m/s and H/D>12, then: C1=6 and C2=1.2.

16

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

3.1.2 Tapered Structures

2 The maximum force for each vibration mode will be determined - FsiMax, multiplied with the value of the corresponding mode shape amplitude i(x) and applied on the structure at location x. The resulting generalized forces will be applied on the structure alternately from one side to the other with a frequency equal with i. Will be calculated stresses at each location x and compared with the endurance limit value for the material used as well as for each critical connection. For tapered structures for the first natural frequency mode the equivalent static force per unit length (3.1.1.2) can be used, if the variation in diameter over the top third is less than 10% of the average diameter of the top third. If the diameter variation exceeds 10%, then the effective static load need be applied only over that part of the structure over which the diameter is within 10% of the average for that part. For tapered structures with a diameter variation exceeding 10% over the top third: C1=3 and C2=0.6 and no increase in these coefficients required for low values for V.

FS ( x, t ) =

V 2 C L D( x) sin[2 e ( x)t ]

(3.1.2.1)

3.2 Analytical Methodology using the Band Limited Random Forcing Model

Treating vortex shedding as a sinusoidal process is an approximation leading to conservative estimates. The variation of the wind speed with height, turbulence of the natural wind and the presence of signs and other accessories all tend to disrupt the spatial correlation of the excitation. It is generally accepted to be more accurate to treat the excitation as a band-limited random process and to assume that the forcing tends to become harmonic only when the motion of the members is sufficiently large to organize the shedding vortices [4,5,6,7]. This tends to occur when the peak amplitude of the motion is of the order of 2% to 2.5% of the diameter or the width of the cross-section, and greater. The below methodology simulates the vortex shedding induced effects for amplitudes smaller than 2.5D% using a band-limited random approach. To apply this methodology one has to perform a forced vibration analysis. The maximum stresses in a member due to vortex shedding excitation shall be computed by loading the member with the peak inertia loads acting statically. The magnitude of the peak inertia load per unit length at any location x along the member for mode of vibration i shall be taken as: Fi ( x ) = (2 i ) 2 y i ( x )m( x )
(3.2.1)

17

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

where:

Fi(x)= member peak inertia load at location x for mode of vibration i, [N/m] m(x)= mass per unit length of member at location x, [kg/m] yi(x)= peak member displacement due to vortex shedding excitation at location x for member mode of vibration i, [m] yi ( x ) = ai i ( x )
where: (3.2.2)

ai(x)= modal coefficient of magnitude of the oscillatory displacement for member mode of vibration i. i(x)= amplitude of the member mode shape at location x for mode of vibration i.
The modal coefficient of magnitude for a particular mode of vibration due to vortex shedding excitation shall be computed as follows.

3.1.1 Non-tapered Structures

ai =

3.5 C L D 2 0.25C

B i (4S ) GM i
2

(3.1.1.1)

Except that if yi(x) exceeds 0.025D, ai shall be taken as


2 C L D 3 i (x )dx
~ H

ai =

i (4S ) GM i

0 2

(3.1.1.2)

where:
C= x 3 i2 ( x ) dx 1 + H 2 LD H 1+3 0
H

(H D )2

(3.1.1.3)

GMi= member generalized mass for mode of vibration i, [kg]


GM i = m( x ) i2 ( x )dx
0 H

(3.1.1.4)

I= structural damping for the ith mode, expressed as a ratio of critical damping. = wind velocity profile exponent (Table: 3.1.1.1).

18

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Table: 3.1.1.1 Area City Centers or Industrial Areas Suburban or Well-Wooded Areas Open Country with Scattered Trees

0.36 0.25 0.15

3.1.2 Tapered Structures

For a member with a tapered diameter or frontal width yi ( x ) = ai ( x1 ) i ( x ) where:

(3.1.2.1)

L C L D 4 ( x1 ) i ( x1 ) ai ( x1 ) = 2 i ( x1 ) (4S )2 GM i
Except that if yi(x)>0.025D(x), ai(x1) shall be taken as:
~

(3.1.2.2)

2 C L D ai ( x1 ) =

(x1 ) D( x ) i (x1 ) dx
x1 b 2 i

x1 +b

(3.1.2.3)

(4S ) GM i

where: x1= location along the tapered member at which vortex shedding excitation potentially occurs b= the length of the member above or below location x1 for which D(x) is within of D(x1). The quantity shall be taken as 10% unless a smaller value can be justified.

( x1 ) =

dD ( x1 ) D ( x1 ) + x dx

(3.1.2.4)

For a tapered member ai(x1) shall be computed for all locations, x1, along the member at which vortex shedding excitation can occur for mode of vibration, i, as determined by 3.B. The largest value of ai(x1) computed shall be used for determining yi(x) and the peak inertia loads. In the case of a member with a uniform taper, p, the limits of integrations specified in equation 3.1.2.3 become:
D( x1 ) p D( x1 ) x1 b = x1 p x1 + b = x1 +

(3.1.2.5)

19

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

In the case of a non-uniformly tapered member, the local variation of D around x=x1 can be approximated by a linear taper p. A good approximation for tapered members is to neglect the variation of D over the limits of integration. With this assumption the modal coefficient of magnitude for a member with a taper p for the ith mode becomes:

2 2 C L D 4 ( x1 ) ai ( x1 ) = 2 p(4S ) iGM i ave

(3.1.2.6)

where: ||ave= average of the absolute values of the mode shape over the portion of the member centered on x1 for which D(x) is within percent of D(x1). Except near the node points, which do not contribute to maximum values:

ave (x1 )

(3.1.2.7)

For a free-standing tapered member, the region of maximum excitation for the fundamental mode is at approximately of the height and moves downwards for higher modes of vibration. Although the evaluation of ai for a particular mode of vibration must be carried out over the entire member, the response is normally governed by the excitation from its main components. For example, in the case of street light poles, the response in particular modes of vibration is dominated by shedding from the pole, with the excitation on the luminaire bracket being of far lesser significance. As a good first approximation, the evaluation of various modal coefficients can thus be confined to locations along the pole. Unless experimentally determined values are available, the damping ratio values - i for members in all modes of vibration shall be taken as i = 0.0075 for steel and aluminum members; i = 0.015 for concrete members. To determine the damping ratios for higher vibration modes can be used the methodology presented in the first part of this thesis.

4. Proposed Design Procedure.

The design procedure proposed to be used as Standard Design Procedure (SDP) are based on the sinusoidal formulation of the vortex shedding phenomenon. This procedure will be used in conjunction with a commercial general purpose finite element program (ALGOR) which has the provisions to apply time-varying (sinusoidal) loads on the finite element model nodes. The structural finite element model will be built using beam finite elements.

20

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

4.1 Design Procedure for Non -Tapered and Tapered Structures

1.) Build a linear beam finite element model which will simulate geometrically and structurally the investigated structure. 2.) Perform a Frequency Finite Element Simulation to obtain the first eight natural frequency modes i, i=1 to 8. 3.) Transform the linear finite element model into a nonlinear one, with isotropic material formulation for beam elements. If using ALGOR prepare the model to perform a Mechanical Event Simulation 4.) With equation (2.2) will be determined the maximum critical wind speed at the top of structure. 5.) Considering the variation of Wind Exposure Coefficient Ce with height (Table: 2.3 or equation (2.17)) and the variation of the wind speed at the top of structure in increments of 1m/s (from 1m/s to Vcr m/s), calculate the wind speeds in the center of each beam finite element. 6.) Based on the wind speeds calculated at point 5 and the diameter of structure (or across to flats dimension) calculate Reynolds Number (with equation (2.16)) in the center of each finite element. Find the Reynolds sub-critical (Re < 3x105) and trans-critical (Re > 3x106) intervals. 7.) Based on equation (2.3) calculate the frequency in the center of each finite element for wind speeds from 1m/s to Vcr m/s. Find the natural frequencies, i, i=1 to 8, locations for sub-critical and trans-critical Reynolds ranges. 8.) For each natural frequency mode locations calculate the maximum sinusoidal vortex shedding induced force (using equation (2.5)) and the moment generated by this force at the base of the structure. For each natural frequency mode consider the location with the highest over-bending moment computed at the base of structure. Consider only the locations which fall in the sub-critical or trans-critical Reynolds range. 9.) Using equations (2.13, 2.14, 2.15) calculate the lower and upper natural frequency bounds for each natural frequency mode location found at point 8. 10.) For each natural frequency mode at each considered location define the frequency domain and calculate the sinusoidal vortex shedding induced force in the center of each finite element which fall in that frequency domain. 11.) For each natural frequency mode and considered frequency domain apply the sinusoidal load (force vs. time) in the center of each finite element considered and run the model. 12.) Compare the stress in the critical areas with the endurance limit. 13.) For the areas with stress distribution higher than endurance limit build a more detailed finite element model and follow the design procedures presented in the first part of this thesis or in reference [40]. 14.) Knowing the location of the structure the frequencies and wind speeds which generate stresses over endurance limit, find the annual number of hours for the wind blowing in that range and calculate the annual number of load cycles. The structures life can be estimated using available Cycle Life Stress Range design curves [49].

21

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

5. Mechanical Event Simulation of Vortex Shedding Structural Response

The analytical experiments will apply the above outlined simulation procedure on two types of free standing structures: a non-tapered antenna tower and a tapered highmast pole.
5.1. Non-Tapered Antenna Tower

The antenna tower (Fig.5.1.1) is 40.8m tall and has four sections flanged together. The diameter of each section is 0.6m and the thickness is: 12.7mm for bottom section, 9.52mm for second section, 7.94mm for third section and 6.35mm for top section. At the top of tower will be clamped-on a frame which will hold the telecommunication antennas. Above and below each flanged connection there is an access opening with a removable cover plate and the base plate is bolted using 38mm diameter anchor bolts. Lets follow step by step procedure 4.1. 1) A beam finite element model was built. Were used 30 linear beam finite element with the proper cross-sectional properties. The model was fully constrained at the base. Were also simulated the masses of all attachment at the proper elevation. 2) The natural frequency modes were calculated:
Mode # 1 2 3 4 5 6 7 8 Frequency [Hz] 1.040 4.340 11.260 20.810 34.170 49.120 67.880 87.100

3) A nonlinear finite element model was built. 4) The maximum critical wind speed at the top of pole was calculated for: q=450 Pa (N/m2) Ce=1.4 (At top of pole) and Vcr=31.12m/s Fig.5.1.1 40.8m Antenna Tower

22

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

5) The wind speed at the top of pole was increase in 1m/s increments until reached a value close to Vcr (32m/s in our case). Considering the variation of coefficient Ce with height was calculated the wind speed for center of each finite element (Table 5.1.1).

Table: 5.1.1 Wind speed variation with height.

23

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

6) Reynolds number was calculated at the center of each finite element for wind speed increments of 1m/s from 1m/s to 32m/s. The gray area is the domain where Reynolds Number is within sub-critical or trans-critical range. (Table 5.1.2). Table: 5.1.2 Reynold Number versus wind speed and height.

It can be observed that with the increase of wind speed the areas where potentially the vortex shedding phenomenon occur, moves towards the bottom of the pole.

24

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

7) Were calculated the frequencies in the center of each finite element for wind speed increments of 1m/s from 1m/s to 32m/s.With red color is marked up the domain where Reynolds Number is within sub-critical or trans-critical range. Were marked up with red the cells where the pole frequency matches the natural frequencies. The gray area is the domain where the pole frequency is within upper and lower frequency boundaries. (Table 5.1.3). Table: 5.1.3 Reynold Number versus wind speed and height.

It can be observed that only the first natural frequency (1.04 Hz) can be found in the domain of interest. The vortex shedding phenomenon can potentially appear at the top of pole for 4m/s wind speed, at 19.72m above ground level for 5m/s wind speed and at 3.4m above ground level for 6m/s wind speed.

25

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

8) Were calculated the vortex shedding induced forces and base overturning moments for potential locations (Table 5.1.4). Table: 5.1.4 Vortex shedding induced forces and base overturning moments for potential locations.
1=1.04
Elevation [m] 40.120 21.080 3.400 D1 [m] 0.600 0.600 0.600 Hz V1 [m/s] 4.0 4.1 4.2 F1 [N] 10.1 10.8 11.0 M1 [Nm] 405.4 228.6 37.4

The top location produces the highest bending moment at the base of pole and was marked up with grey color but the largest frequency domain is for middle location (elevation 21.08m) and the overall bending moment will be higher. Therefore, this location is considered to produce the highest stress at the base of pole. 9) Were calculated the upper and lower boundary limits for the first natural frequency mode (Table 5.1.5). Table: 5.1.5 Upper and lower boundary limits for first natural frequency mode.
Mode # 1 Frequency [Hz] 1.040 Vicr [m/s] 4.14 A K

z0
0.05

z
40.12

A*Vcr*K

Upper Limit [Hz] 1.20

Lower Limit [Hz] 0.88

ln(z/z0)
2.5 0.4 -0.62

The frequency domain for the first natural frequency mode was marked up with grey color in table 5.1.3

26

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

10) For the frequency domain are calculated the vortex shedding induced forces in the center of the finite elements (Table 5.1.6). Table: 5.1.6 Vortex shedding induced forces for first natural frequency mode.

The vortex shedding induced forces are sinusoidal time varying and the values in table: 5.1.6 are the maximum values. 11) The time-varying vortex shedding induced forces will be applied in the center of corresponding finite elements.

27

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

12) The time-varying vortex shedding induced bending stress at the bottom of pole (Fig: 5.1.2) is in the range: -2 MPa to 2 MPa which is less than 31MPa which is the endurance limit for the critical welded connections at the bottom of pole, therefore the vortex shedding phenomenon will not produce fatigue damage at the bottom of this investigated antenna structure. Thus steps 13 and 14 are not required for this structure.
Time Varying Stress Induced by Vortex Shedding for First Natural Frquency Mode 2500000 2000000 1500000 1000000 Stress (Pa) 500000 0 -500000 -1000000 -1500000 -2000000 -2500000 Time (s) 0 10 20 30 40 50

Fig. 5.1.2 Time-varying stress induced by vortex shedding phenomenon.


Time Varying Top Pole Displacement Induced by Vortex Shedding for First Natural Frquency Mode 0.025 0.02 0.015 Displacement (m) 0.01 0.005 0 -0.005 -0.01 -0.015 -0.02 -0.025 Time (s) 0 10 20 30 40 50

Fig. 5.1.3 Time-varying displacement induced by vortex shedding phenomenon.

28

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

5.2. Tapered Highmast Tower

In fig 5.2.1. is shown a 38m highmast pole which will be investigated for dynamic response to vortex shedding induced loads. The structure has four sections which overlap each other. The thickness of bottom section is 7.94mm, second and third sections 6.35mm and top section 4.76mm. The structure is 12 sided with 850mm across to flats dimension for bottom and 200mm across to flats for top. The structure is bolted to foundation using 38mm diameter anchor bolts. At the bottom there is a large winch servicing cutout with a 16mm thick reinforcing ring. Inside the pole, at the bottom, there is a winch mounted on a angle profile. At the top of pole will be docked a frame which will hold 12 floodlights with their ballasts. Inside the pole are ran the steel cables required to lower or rise the frame with lamps. Lets follow step by step procedure 4.2. 1) A beam finite element model was built. Were used 30 linear beam finite element with the proper cross-sectional properties. The model was fully constrained at the base. Were also simulated the masses of all attachment at the proper elevation. The slip joint connections cross sectional properties were calculated considering the change in thickness because of the adjacent sections overlapping.

Fig. 5.2.1. 38m Highmast Pole

29

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

2) The natural frequency modes were calculated:


Mode # 1 2 3 4 5 6 7 8 Frequency [Hz] 0.750 2.570 6.280 11.420 18.660 27.250 37.590 49.440

3) A nonlinear finite element model was built. 4) The maximum critical wind speed at the top of pole was calculated for: q=450 Pa (N/m2) Ce=1.4 (At top of pole) and Vcr=31.12m/s 5) The wind speed at the top of pole was increase in 1m/s increments until reached a value close to Vcr (32m/s in our case). Considering the variation of coefficient Ce with height was calculated the wind speed for center of each finite element (Table 5.2.1).

30

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Table: 5.2.1 Wind speed variation with height.

6) Reynolds number was calculated at the center of each finite element for wind speed increments of 1m/s from 1m/s to 32m/s. The gray area is the domain where Reynolds Number is within sub-critical or trans-critical range. (Table 5.2.2).

31

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Table: 5.2.2 Reynold Number versus wind speed and height.

It can be observed that with the increase of wind speed the areas where potentially the vortex shedding phenomenon occur, moves towards the top of the pole.

32

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

7) Were calculated the frequencies in the center of each finite element for wind speed increments of 1m/s from 1m/s to 32m/s.With red color is marked up the domain where Reynolds Number is within sub-critical or trans-critical range. Were marked up with colors the cells where the pole frequency matches the natural frequencies (Table 5.2.3). Table: 5.2.3 Reynold Number versus wind speed and height.

33

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

8) Were calculated the vortex shedding induced forces and base overturning moments for potential locations (Table 5.2.4). Table: 5.2.4 Vortex shedding induced forces and base overturning moments for potential locations.

9) Were calculated the upper and lower boundary limits for the first natural frequency mode (Table 5.2.5). Table: 5.2.5 Upper and lower boundary limits for natural frequency modes.

34

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

10) For the frequency domains are calculated the vortex shedding induced forces in the center of the finite elements (Tables 5.2.6 to 5.2.9).

Table: 5.2.6 Vortex shedding induced forces 1st frequency mode. Table: 5.2.7 Vortex shedding induced forces 2nd frequency mode.

35

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

Table: 5.2.8 Vortex shedding induced forces 3rd frequency mode. Table: 5.2.9 Vortex shedding induced forces 4th frequency mode.

36

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

The vortex shedding induced forces are sinusoidal time varying and the values in tables: 5.2.6 to 5.2.9 are the maximum values. 11) The time-varying vortex shedding induced forces will be applied in the center of corresponding finite elements. 12) The time-varying vortex shedding induced bending stress range at the bottom of pole (Table 5.2.10) is less than 31MPa, which is the endurance limit for the critical welded connections at the bottom of pole, only for the first natural frequency mode therefore the vortex shedding phenomenon will potentially produce fatigue damage at the bottom of pole for second, third and forth frequency modes. Thus steps 13 and 14 are required for this structure [40].

37

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

6. Conclusions

As demonstrated, based on the above simulations, the tubular, multisided or round, tapered or nontapered free standing structures may be subject to considerable dynamic induced stress by vortex shedding phenomenon. Also could be seen that the vibrations induced by vortex shedding may occur at relatively moderated wind speeds, therefore the structures may undergo a considerable number of stress cycles. Considering these aspects the risk of structural fatigue must be taken into account at the design stage. Even if fatigue analysis is not explicitly demanded in the design code which is applied, the risk of fatigue should be kept in mind when designing slender free standing structures. For this the design engineer must have a good understanding of the vortex shedding phenomenon and follow a logic design procedure to investigate the response of the structure to vortex shedding. The design standards lack in presenting easy to use calculation methodologies for investigating vortex shedding phenomenon and its impact on structures, and usually the design procedures require the induced forces to be applied statically. Vortex shedding is a dynamic phenomenon and to get realistic structural responses to this phenomenon it must be simulated dynamically. The dynamic structural response is the worst case scenario because of the realistic account for inertial induced loads. As seen in the analytical models the wind speed is the main parameter and good wind speed estimations are critical. Based on the above outlined design procedure one can easily estimate the critical wind speed which potentially trigger the vortex shedding phenomenon as well as the upper and lower boundaries for that critical wind speed. After defining the areas where potentially the vortex shedding occurs will be calculated the time-varying induced forces which will act on the structure. The design procedure can be easily automated using MS EXCEL program in conjunction with a general purpose Fig. 6.1 High Stress Locations Pole Bottom finite element software. The interpretation of the stress range results should be made correctly considering for each critical connection the proper fatigue category. To evaluate correctly the stress in the investigated connection a detailed finite element model should be built. As seen in fig. 6.1 is very hard to predict the high stress locations for complex geometries without numerical simulations, loading the structure with the vortex shedding induced forces in all possible directions to get a good feeling about the structures response. The design procedure presented above (4.1) brings few interesting and original things:

38

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

it is an easy to follow procedure; all parameters are clearly defined; the vortex shedding induced loads are applied dynamically like in the real situation; the procedure can be fully automated.

The results obtained testing this design procedure on transmission, antenna and highmast towers are in the realistic range and comparable with experimental observations. Considering that computational fluid dynamics simulation or wind tunnel tests to investigate the vortex shedding effects for these type of structures will be very time consuming and expensive this easy to use methodology represents an efficient tool for the numerical analyst involved into the design process.

39

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

References

1. Solari, G. The role of analytical methods for evaluating the wind-induced response of structures, Journal of Wind Engineering in Industrial Aerodynamics, 90, 1453-1477, 2002. 2. Clas Dyrbye, Svend O. Hansen Wind Loads on Structures, John Wiley & Sons, 1999. 3. NBC 1995 National Building Code of Canada -1995. 4. CAN/CSA-S6-00 Canadian Highway Bridge Design Code -2005. 5. CAN/CSA-S6.1-00 Commentaries to Canadian Highway Bridge Design Code-2005. 6. Eurocode 1 Basis of design and actions on structures Part 2-4: Actions on structures Wind Actions, ENV 1991-2-4. 7. Standard Specifications for Structural Supports for Highway Signs, Luminaires and Traffic Signals, American Association of State Highway and Transportation Officials, 2001. 8. ASCE7-2005 Minimum Design Loads for Buildings and Other Structures, American Society of Civil Engineering, Reston, VA, 2005. 9. AS170.2, SAA Loading Code, Part 2 Wind forces, AS1170.2-89, Australian Standards, 1989. 10. AIJ-RLB Recomandations for Loads on Buildings, Arhitectural Institute of Japan, 1993 (English version 1996). 11. NTE 003/04/00 Design Standards for Electrical Transmission Lines over 100 kV (Replaced Standard PE 104). 12. PE 105 Design Methodology for Electrical Transmission Lines Steel Towers. 13. STAS 10101/20-90 Structural Loads Induced by Wind Romanian Standardization Institute 1990. 14. Yukio Tamura, Ahsan Kareem, Giovanni Solari, Kenny C.S. Kwok, John D. Holmes Report by Working Group WCE Dynamic Response, 2001. 15. Davenport A.G. The application of statistical concepts to the wind loading of structures, Proceedings of the Institution of Civil Engineers, London, UK, 1961;19 :449-472. 16. Davenport A.G. Note on the distribution of the largest value of a random function with application to gust loading. Proceedings of the Institution of Civil Engineers, London, UK, 1964;24 :187-196. 17. Davenport, A.G. New Approaches to Design Against Wind Action, Unpublished Text. Boundary Layer Wind Tunnel Laboratory, University of Western Ontario, London, Ontario 1976. 18. Vikery, B.J. On the reliability of gust loading factors. Proceedings of Technical Meeting concerning Wind Loads on Buildings and Structures. National Bureau of Standards, Washington, DC: 1970;93-104. 19. Simiu E. Equivalent Static Wind Loads for Tall Buildings Design. Journal of Structural Div., ASCE 1976; 102:719-737. 20. ESDU 76001 The response of flexible structures to atmospheric turbulence. London, UK, Engineering Science Data Unit, 1976. 21. ECCS Recommendations for the calculation of wind effects on buildings and structures. Brussels: European Convention for Constructional Steelworks, 1978. 22. Solari G. Equivalent Wind Spectrum Technique: theory and applications. Journal of Structural Engineering, ASCE 1988; 114: 1303-1323. 23. Kaspersky M. Extreme wind load distributions for linear and nonlinear design. Engineering Structures, 1992;14:27-34.

40

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

24. Holmes J.D. Along-wind response of lattice towers: part 1 derivation of expressions for gust response factors. Engineering Structures, 1994;16: 287-292. 25. Holmes J.D. Along-wind response of lattice towers: part 2 Effective load distributions. Engineering Structures, 1996;18: 489-494. 26. Davenport, A.G. How can we simplify and generalize wind loads. Lournal of Wind Engineering in Industrial Aerodynamics, 1995; 54-55: 657-669. 27. Tamura, Y., Kawai, H., Uematsu, Y., Marukawa, H., Fuji, K., Taniike, Y., - Wind loads and wind induced response estimations in the Recommendations for loads on buildings, AIJ 1993. Engineering Structures 1996; 18: 399-411. 28. Solari, G. Towards a global model for calculating 3D equivalent static wind forces on structures. Proceedings on 7th US National Wind Engineering Conference, Los Angeles: 1993, 719-728. 29. Piccardo, G., Solari, G. A refined model for calculating 3D equivalent static wind forces on structures. Journal of Wind Engineering in Industrial Aerodynamics, 1996; 65: 21-30. 30. Solari, G. Mathematical Model to Predict 3D Wind Loading on Buildings. Journal of Engineering Mechanics, ASCE, 1985; 111: 254-276. 31. Piccardo, G., Solari, G. Closed Form Prediction of 3D Wind-excited response of slender structures. Journal of Wind Engineering in Industrial Aerodynamics, 1998; 74-76: 697-708. 32. Piccardo, G., Solari, G. 3D wind-excited response of slender structures: Closed form solution. Journal of Structural Engineering, ASCE 2000; 126: 936-943. 33. Piccardo, G., Solari, G. Generalized equivalent spectrum technique. Wind and Structures Journal, 1998; 1: 161-174. 34. P.A.B. de Sampaio, A.L.G.A. Coutinho Simulating Vortex Shedding at High Reynolds Numbers, Proceedings of the 10th International Offshore and Polar Conference, Seattle, USA, 2000. 35. Yin Zhou, Ahsan Kareem, Ming Gu Mode Shape Corrections for Wind Load Effects, Journal of Engineering Mechanics, January, 2002, 15-23. 36. Maria Pia Repetto, Solari, G. Directional Wind-Induced Fatigue on Slender Vertical Structures, Journal of Structural Engineering, ASCE, July, 2004, 10032-1040. 37. Dexter, R.J., Ricker, M.J. Fatigue-Resistant Design of Cantilevered Signal, Sign, and Light Supports NCHRP Report 469, University of Minnesota, USA, Published by: National Academy Press, 2002. 38. Christopher M. Foley, Scott J. Ginal, John L. Peronto, Raymond A. Fournelle Structural Analysis of Structures and Luminaire Supports, Marquette University, Wisconsin, USA, 2004. 39. Giosan, I., Brockman, T. - Streamlining Metal Poles, Mechanical Engineering Magazine, April, 2000, 68-70. 40. Giosan, I. Building Better Free Standing Structures, Electrical Line Magazine, February, 2006, 63-70. 41. Harris, Cyril, M and Crede, Charles E. Shock and Vibration Handbook, Section 29, Vibration of Structures Induced by Wind, Part II, 2nd Edition, McGraw-Hill Book Co. 1976. 42. Vikery, B.J., Load Fluctuations in Turbulent Flow, ASCE, J. Eng. Mech. Div., Vol. 94, February 1968. 43. Vikery, B.J. and Clark, A. W., Lift of Across Wind Response of Tapered Stacks, ASCE, J. Eng. Mech. Div., Vol. 98, January 1972.

41

Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation

44. Wooton, L.R., The Oscilation of Large Circular Stacks in Wind, Proceedings I.C.E., August 1969.1969. 45. Simiu, E., Scalon, R.H. Wind Effects on Structures: Fundamentals and Applications to Design 3rd Edition, John Willey & Sons, New York, 1996. 46. ESDU-96030 Response of Structures to Vortex Shedding. Structures of Circular or Polygonal Cross Section. Engineering Science Data Unit, London, UK, 1998. 47. Wieringa, J. Representative roughness parameters for homogeneous terrain. BoundaryLayer Meorology, 63, 323-363, 1993. 48. Lettau, H. Note of aerodynamic roughness-parameter estimation on the basis of roughnesselement description. Journal of Applied Meteorology, 8, 828-832, 1969. 49. Handbook of Steel Construction Canadian Institute of Steel Construction, 3rd Edition, 1982.

42

You might also like