viv and galloping interaction for a 3:2 rectangular...

8
VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini 1 , A. M. Marra 1 , T. Massai 1 , and G. Bartoli 1 1 CRIACIV / Department of Civil and Environmental Engineering, University of Florence, Via S. Marta 3, Florence, Italy. [email protected]fi.it Abstract The interaction between vortex-induced vibration and galloping for an infinitely long 3:2 rect- angular cylinder was investigated through wind tunnel tests. This section was found to be very unstable and large vibration amplitudes were observed also at high Scruton number due to this interference effect. In order such a phenomenon not to occur it was necessary to increase signif- icantly the mass-damping parameter. In particular, no interaction was observed for a ratio of the quasi-steady galloping to vortex-induced vibration critical wind speeds as high as 7.5, while it was still present for a ratio of 5.3. The practical consequence of that seems to be the necessity to install dampers in several prismatic slender structures with a rectangular cross section similar to the one considered herein. 1 Introduction Galloping is a dynamic instability of slender prismatic structures caused by the self-excitation due to wind. Mathematically speaking it is a Hopf bifurcation and then, once the critical wind speed has been exceeded, it manifests itself as a limit-cycle harmonic oscillation whose amplitude steadily grows by increasing the flow velocity. The critical wind speed is proportional to a mass-damping parameter called Scruton number. By contrast, vortex-induced vibration (VIV) is caused by the nonlinear resonance of the force due to the alternate shedding of vortices with one mode of vibration of the structure. It is a phenomenon which starts at a critical wind speed which depends on the Strouhal number and disappears beyond a certain flow velocity. The amplitude of vibration and the extension of the so-called lock-in range depend on the Scruton number. In slender prismatic structures with bluff cross section and sufficient afterbody both phenomena are possible and therefore interaction can occur. Eurocode 1 (EN 1991-1-4, 2010) states that, if the ratio of the galloping to the VIV critical wind speed (the former calculated with the classical quasi-steady theory) is either lower than 0.7 or larger than 1.5 the two phenomena can be considered separately. Rectangular cylinders with a significant afterbody (say B/D > 0.6, being B the width and D the depth of the section) but with a width-to-depth ratio lower than the critical value for shear-layer unsteady reattachment (about 2.8 in smooth flow) are known to be very unstable with respect to gallop- ing (Parkinson & Brooks, 1961; Parkinson & Sullivan, 1979; Parkinson & Wawzonek, 1981; Novak & Davenport, 1970; Bearman et al., 1987). It was shown that prismatic towers with rectangular cross section could be prone to galloping (Novak & Davenport, 1970; Parkinson & Sullivan, 1979; Parkin- son & Wawzonek, 1981) and this may be of particular concern nowadays due to the reduction of modal frequencies and damping, as well as the increased lightness and slenderness of tall buildings. A possible interaction between VIV and galloping makes the problem even harder since it allows the onset of limit-cycle oscillations at relatively low wind speed, that is at the critical velocity for VIV, but with unrestricted amplitudes growing with the flow speed, typical of galloping, even for 1

Upload: tranthu

Post on 10-Jun-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

VIV and galloping interaction for a 3:2 rectangular cylinder

C. Mannini1, A. M. Marra1, T. Massai1, and G. Bartoli1

1CRIACIV / Department of Civil and Environmental Engineering, University of Florence,Via S. Marta 3, Florence, Italy. [email protected]

Abstract

The interaction between vortex-induced vibration and galloping for an infinitely long 3:2 rect-angular cylinder was investigated through wind tunnel tests. This section was found to be veryunstable and large vibration amplitudes were observed also at high Scruton number due to thisinterference effect. In order such a phenomenon not to occur it was necessary to increase signif-icantly the mass-damping parameter. In particular, no interaction was observed for a ratio of thequasi-steady galloping to vortex-induced vibration critical wind speeds as high as 7.5, while it wasstill present for a ratio of 5.3. The practical consequence of that seems to be the necessity to installdampers in several prismatic slender structures with a rectangular cross section similar to the oneconsidered herein.

1 Introduction

Galloping is a dynamic instability of slender prismatic structures caused by the self-excitation due towind. Mathematically speaking it is a Hopf bifurcation and then, once the critical wind speed has beenexceeded, it manifests itself as a limit-cycle harmonic oscillation whose amplitude steadily grows byincreasing the flow velocity. The critical wind speed is proportional to a mass-damping parametercalled Scruton number.

By contrast, vortex-induced vibration (VIV) is caused by the nonlinear resonance of the force dueto the alternate shedding of vortices with one mode of vibration of the structure. It is a phenomenonwhich starts at a critical wind speed which depends on the Strouhal number and disappears beyonda certain flow velocity. The amplitude of vibration and the extension of the so-called lock-in rangedepend on the Scruton number.

In slender prismatic structures with bluff cross section and sufficient afterbody both phenomena arepossible and therefore interaction can occur. Eurocode 1 (EN 1991-1-4, 2010) states that, if the ratioof the galloping to the VIV critical wind speed (the former calculated with the classical quasi-steadytheory) is either lower than 0.7 or larger than 1.5 the two phenomena can be considered separately.

Rectangular cylinders with a significant afterbody (say B/D > 0.6, being B the width and Dthe depth of the section) but with a width-to-depth ratio lower than the critical value for shear-layerunsteady reattachment (about 2.8 in smooth flow) are known to be very unstable with respect to gallop-ing (Parkinson & Brooks, 1961; Parkinson & Sullivan, 1979; Parkinson & Wawzonek, 1981; Novak& Davenport, 1970; Bearman et al., 1987). It was shown that prismatic towers with rectangular crosssection could be prone to galloping (Novak & Davenport, 1970; Parkinson & Sullivan, 1979; Parkin-son & Wawzonek, 1981) and this may be of particular concern nowadays due to the reduction of modalfrequencies and damping, as well as the increased lightness and slenderness of tall buildings.

A possible interaction between VIV and galloping makes the problem even harder since it allowsthe onset of limit-cycle oscillations at relatively low wind speed, that is at the critical velocity forVIV, but with unrestricted amplitudes growing with the flow speed, typical of galloping, even for

1

Page 2: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 2

large values of the Scruton number. This phenomenon seems to be already recognizable in the exper-iments of Novak & Davenport (1970). However, the first clear observation and discussion of such aninteraction can be found in Parkinson & Sullivan (1979) for cantilevered square prisms in turbulentflow. Parkinson & Wawzonek (1981) further investigated the phenomenon with experimental tests ona two-dimensional square cylinder in smooth flow.

Another crucial issue for rectangular cylinders with B/D in a range around 1.0 ÷ 2.0, highlightedby a careful literature survey, is the large variability of values for the quasi-steady galloping stabilityparameter ag = −dCL/dα(0) − CD(0), being CL and CD respectively the lift and drag coefficientsand α the angle of attack, obtained by different wind tunnel laboratories. It seems that this parame-ter is strongly dependent on flow characteristics and test boundary conditions (oncoming turbulence,blockage, aspect ratio, edge sharpness of the model, etc.). Nevertheless, the values of ag providedby Eurocode 1 (EN 1991-1-4, 2010) represent a lower bound with respect to the literature resultsand therefore do not appear conservative at all. Also the value of the Strouhal number reported byEurocode 1 seems to be quite low.

In this work the aeroelastic behaviour of a 3:2 sharp-edged rectangular cylinder (B/D = 1.5)was experimentally investigated through static and aeroelastic tests in order to shed some light on thecomplicated VIV-galloping interaction phenomena. In particular, two different set-ups were used forthe dynamic tests. The study was initiated during the experimental campaign performed at the Re-search Centre on Building Aerodynamics and Wind Engineering (CRIACIV) to assess the aeroelasticstability of a real structure.

2 Wind tunnel test description

Experiments were conducted in the open-circuit boundary layer wind tunnel of the CRIACIV, locatedin Prato, Italy. The test section is 2.42 m wide and 1.6 m high. A wooden sectional model, 986 mmlong, 116 mm wide (B) and 77 mm deep (D), was used to perform both static and dynamic tests(Figure 1). To enforce bidimensional flow conditions, rectangular plates in plywood were providedat the model ends; their dimensions (450 mm × 150 mm × 4 mm) were defined according to ESDUprescriptions (Cowdrey, 1963; Obasaju, 1979). The mass of the model, end-plates and supportingcarbon-fiber tube was 1.730 kg. The blockage ratio given by the model alone, calculated as D/Hwt,being Hwt the height of the wind-tunnel test section, was 4.8%. All the tests were carried out in anominal smooth flow with a turbulence intensity around 1%, but slightly varying according to the flowspeed and the actual position of the set-up in the wind tunnel.

For the static tests the model was connected to six load cells (three on each side) through a systemof connecting rods with spherical hinges, which allowed the measurement of drag, lift and moment.The aerodynamic force coefficients were determined for angles of attack ranging from about -10o

to +10o. The tests were repeated several times to verify the reliability of the measurements. Themodel was also equipped with 28 pressure taps, located at the central section of the cylinder andalong a longitudinal array, and registrations at a sampling frequency of 500 Hz were performed withpiezoelectric pressure transducers and the system PSI DTC Initium.

Two distinct aeroelastic set-ups were used for the dynamic tests (Figure 1). In the first set-up(Setup I), reported in Figure 1(a), the model was elastically suspended through eight vertical coilsprings (four on each side), each one with a nominal stiffness of 5340 N/m. The horizontal translationof the model was restrained by means of four pretensioned steel cables, while the streamwise distancebetween the pairs of springs was large enough to guarantee a very high stiffness in the rotational degreeof freedom. The plunging mode in the basic configuration resulted to have a frequency of 14.73 Hz, aratio-to-critical damping ζ = 1.6 · 10−3 and an equivalent mass m = 4.892 kg (participant mass of theoscillating system). Rolling motion was also possible, with a frequency 20% higher if no additional

Page 3: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 3

(a) (b)

Figure 1: Set-up I (a) and Set-up II (b).

masses were provided, but it has never been significantly excited during the tests.In Figure 1(b) the second set-up (Set-up II) is shown. The model was connected through its

longitudinal axis-tube to two shear-type frames. Only a vertical displacement was allowed by thetwo frames due to the very large flexural stiffness of the two vertical elements at which the modelwas connected. The aerodynamic damping due to the exposition to the flow of the plate-springs wasverified to be very small. Two configurations were studied by changing the length of the plate-springs.The mass m of the vibrating system in the basic configuration was respectively 4.49 kg and 4.72 kgfor a plate length of 490 mm and 624 mm. As an example, in the second case the plunging frequencywas 11.23 Hz and the damping ratio ζ = 0.94 · 10−3. In addition to the low damping and easiness ofassembling, there were several advantages in using Set-up II, such as the much smaller obstruction ofthe test section and the direct restriction of unwanted degrees of freedom (sway, pitch, roll, providedthat the model is stiff enough, and longitudinal motion).

The displacements of the model in case of Set-up I and II were recorded respectively with threeand two non-contact laser transducers.

3 Wind tunnel test results

3.1 Static tests

Static tests were mainly performed to estimate the drag and the lift coefficients needed to predictthe critical velocity for the quasi-steady galloping. Figure 2 reports the drag and lift coefficients(normalized with respect to the model depth D) against the angle of attack, evaluated for a wind speedU = 17.6 m/s, corresponding to a Reynolds number Re = UD/ν = 90, 350, being ν the kinematicviscosity. The symmetry of the curves highlights the quality of the model. For α = 0◦ the dragcoefficient is 1.765, which is in good agreement with the value of 1.8 reported in Norberg (1993). Asshown in the figure, the slope of the lift coefficient at zero angle of attack is -7.46. The correspondinggalloping stability parameter ag results 5.5, which is quite different from the values of 1.91, given byNovak & Tanaka (1974), and 1.7, provided by the Eurocode 1 (EN 1991-1-4, 2010). Conversely, thisis quite close to the value of 5.44 obtained through numerical simulations by Robertson et al. (2003)at low Reynolds number (Re = 250). These significant differences could be justified by considering

Page 4: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 4

−10 −5 0 5 101.3

1.4

1.5

1.6

1.7

1.8

α [deg]

CD

−10 −5 0 5 10−1

−0.5

0

0.5

1

α [deg]

CL

dCL

dα= −7.46

Figure 2: Mean values of aerodynamic drag and lift coefficients at various angles of attack.

0 0.1 0.2 0.3 0.4 0.510

−4

10−2

100

102

104

U = 7.51 m/s

nD/U

SL

L

nD/U = 0.106 nD/U = 0.318

n0D/U

0 0.1 0.2 0.3 0.4 0.510

−4

10−2

100

102

104

U = 18.81 m/s

nD/U

SL

L

nD/U = 0.106

nD/U = 0.212

nD/U = 0.318

n0D/U

Figure 3: Power spectral density of lift coefficient for two wind speeds. n0∼= 40 Hz is the vertical

bending frequency of the model.

the results available in the literature for the square cylinder. In fact, a large variability of the valuesof the quasi-steady galloping stability parameter (ag = 1.13 ÷ 5.4) is observed by comparing theresults obtained in different laboratories (Novak, 1969; Bearman et al., 1987; Norberg, 1993). Inparticular, as observed by Bearman et al. (1987), it seems that this parameter is strongly dependent onflow characteristics and test boundary conditions (oncoming turbulence, blockage, aspect ratio, edgesharpness of the model, etc.).

The power spectral density of the lift coefficient (Figure 3) as well as the one of the pressurecoefficient at the midpoint of the lower side of the cylinder (Figure 4), both evaluated at two differentwind speeds, provide a Strouhal number of 0.106, which is in good agreement with the value of 0.105reported in Norberg (1993). By contrast, a quite different value can be found in Eurocode 1 (EN1991-1-4, 2010), where St = 0.09 is suggested. In the graphs of Figures 3-4, in addition to the peakscorresponding to the vortex-shedding frequency and the natural bending frequency of the model, thefirst and second subharmonics of the vortex-shedding frequency are evident.

Page 5: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 5

0 0.1 0.2 0.3 0.4 0.510

−6

10−4

10−2

100

102

104

nD/U = 0.106

nD/U = 0.212

U = 9.88 m/s

nD/U

Spp

0 0.1 0.2 0.3 0.4 0.510

−6

10−4

10−2

100

102

104

nD/U = 0.106

nD/U = 0.212

U = 18.81 m/s

nD/U

Spp

Figure 4: Power spectral density of pressure coefficient at the midpoint of the lower side (parallel tothe flow) of the cylinder for two wind speeds.

0 0.5 1 1.5 2 2.5 3 3.50

0.02

0.04

0.06

0.08

0.1

0.12

U/Ur

y′/D

0.6 0.7 0.8 0.9 1 1.1 1.2 1.30

0.01

0.02

0.03

0.04

0.05

0.06

0.167

0.149

1

1

U/Ur

y′/D

Sc = 14.9 − Ur = 10.7 m/s

Sc = 22.7 − Ur = 7.1 m/s

Sc = 48.5 − Ur = 10.8 m/s

Sc = 84.1 − Ur = 6.5 m/s

Sc = 138.4 − Ur = 6.6 m/s

Sc = 195.4 − Ur = 6.6 m/s

Sc = 232.9 − Ur = 6.6 m/s

Figure 5: Standard deviation of cross-flow model displacements for different values of the Scrutonnumber obtained with Set-up I. The right frame is a close-up view of the left one.

3.2 Dynamic tests

The dynamic tests were performed by recording the cross-flow displacements of the model at variouswind speeds. The flow velocity was both increased and decreased to show the possible presence ofhysteretic phenomena. During the tests with Set-up I the mass of the system was varied by addingweights at the model ends, whereas the damping was modified by wrapping the springs with tape. Theresults are reported in Figure 5, where it can be noted that for Scruton numbers Sc = 4πmζ/ρD2

up to at least 138 unrestricted galloping-type oscillations start around a wind speed Ur = n0D/St,like for vortex-induced vibration, instead of the predicted quasi-steady galloping critical wind speedUg = 2n0DSc/ag. Interestingly, this seems to be true also when Ug is slightly lower than Ur (atlow Scruton number). This “quenching effect” of the vortex system on the galloping instability wasobserved also by Bearman et al. (1987). By contrast, for very high values of the Scruton number(Sc = 195 and 233) the oscillation amplitudes suddenly drop once a narrow lock-in region has beenovercome, thus following a classical VIV behaviour. Nevertheless, it is surprising the non-negligibleamplitude of oscillation reached for such high Scruton numbers, as shown in Figure 6(b). In the

Page 6: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 6

0 0.5 1 1.5 2 2.50

0.02

0.04

0.06

0.08

0.1

0.12

U/Ur

y′/D

Sc = 138.4

(a)

0 0.5 1 1.5 2 2.5 3 3.5 40

0.005

0.01

0.015

0.02

U/Ur

y′/D

Sc = 232.9

(b)

Figure 6: Standard deviation of cross-flow model displacements for high values of the Scruton numberobtained with Set-up I.

same figure it can be remarked that an increase of the oscillations was observed also around a flowvelocity corresponding to the third subharmonic of the system natural frequency (U/Ur between 2.5and 3), similarly to what observed by Bearman et al. (1987). It is also worth noting in Figures 5 and6(a) that after the instability onset the amplitude of oscillation linearly grows with the wind speed, aspreviously observed by other Authors (Parkinson & Sullivan, 1979; Parkinson & Wawzonek, 1981;Bearman et al., 1987). Nevertheless, a kink in the curves is apparent slightly after U/Ur = 1.5, foroscillation amplitudes lower for higher Scruton numbers. A similar feature was highlighted also forthe square cylinder by Bearman et al. (1987). A linear trend seems to be recovered after a transitionrange but with a reduced slope and more regular harmonic vibrations.

The value of the Scruton number is related to the ratio of the critical wind speed of quasi-steadygalloping Ug to that of vortex induced vibration Ur:

Ug

Ur=

2St

agSc = 0.0385 · Sc (1)

A Scruton number of 138.4 corresponds to a ratio of critical velocities of 5.3, whereas Sc = 195.4corresponds to Ug/Ur = 7.5. This means that a ratio of critical wind speeds between 5.3 and 7.5 isneeded in order to prevent the interaction between the two aeroelastic phenomena and be able to applythe classical theories for the calculation of the instability onset. A similar result was obtained also byParkinson & Wawzonek (1981), who found that a ratio as high as 8.4 was necessary not to observethe VIV-galloping instability for a 3:2 rectangular cylinder. By contrast, this is very different from theprescriptions of Eurocode 1 (EN 1991-1-4, 2010), which generally excludes the interaction betweenVIV and galloping outside the range 0.7 ≤ Ug/Ur ≤ 1.5.

The previously discussed dynamic results were also checked mounting the same model on Set-up II, where also lower values of the Scruton were reached. In this case, just the mass was varied andScruton numbers in the range 8÷26 were obtained. As shown in Figure 7 for a value of the mass-damping parameter around 8 the synchronization with the 1/3-superharmonic of the system naturalfrequency was registered and evident hysteretic effects were observed (Figure 7(b)). This superhar-monic resonance already disappeared for a Scruton number of about 20, in agreement with previousresults obtained with Set-up I. A VIV-galloping combined instability was observed for wind speedshigher than Ur. For higher frequencies of the oscillating system and therefore for higher critical wind

Page 7: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 7

0 0.25 0.5 0.75 1 1.25 1.5 1.750

0.02

0.04

0.06

0.08

0.1

0.12

1

0.235

U/Ur

y′/D

Sc = 8.2 − Ur = 8.2 m/s

Sc = 19.9 − Ur = 4.9 m/s

Sc = 26.1 − Ur = 7.8 m/s

(a)

0.25 0.3 0.35 0.4 0.45 0.50

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

Sc = 7.7 −Ur = 8.2 m/s

U/Ur

y′/D

U increasingU decreasingU increasing from zero

(b)

Figure 7: Standard deviation of cross-flow model displacements obtained with Set-up II.

15 16 17 18 19 20−0.02

−0.015

−0.01

−0.005

0

0.005

0.01

0.015

0.02U/Ur = 1.04

t [s]

y/D

15 16 17 18 19 20−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

0.2U/Ur = 1.48

t [s]

y/D

Figure 8: Examples of cross-flow vibration time histories for Sc = 8.2 and Set-up II.

speeds the instability onsets slightly earlier in terms of reduced velocities but the slope of the curvesdoes not seem to change. The model undergoes almost perfect sinusoidal oscillations around the peakof the superharmonic resonance curve (Figure 7(b)), while significant modulations of the vibrationsappear in the galloping range, as shown in Figure 8, their magnitude progressively decreasing with thewind speed.

The change in the slope of the galloping curves and the slight delay in the instability onset (Fig-ures 5 and 7(a)) may be due to a slight difference in the turbulence intensity (higher at low wind speedand in the position of Set-up II). However, further investigation is required to confirm this conjecture.

4 Concluding remarks

The 3:2 rectangular section was found to be extremely susceptible to galloping and in particular to theVIV-galloping interference phenomenon. It was observed that a ratio in the range 5.3 ÷ 7.5 betweenthe theoretical critical velocities for the two instabilities was necessary to avoid such an interaction;this is much higher than 1.5, as suggested by Eurocode 1. This condition corresponds to very large

Page 8: VIV and galloping interaction for a 3:2 rectangular cylinderiawe.org/Proceedings/EACWE2013/C.Mannini.pdf · VIV and galloping interaction for a 3:2 rectangular cylinder C. Mannini1,

6th European and African Wind Engineering Conference 8

values of the Scruton number and from the practical engineering point of view this can easily requirethe installation of dampers in order to guarantee the safety of a wide range of slender structures.

Moreover, the static test results highlighted that the value of the galloping stability parameterprovided by the European provisions may be not conservative. In fact, ag = 5.5 was measured insteadof 1.7 and the literature review showed that, despite the large sensitivity of this parameter to theexperimental test conditions, the Eurocode value seems to be very low.

Finally, the value 0.106 obtained for the Strouhal number is in agreement with most of the resultsavailable in the literature but it is significantly higher than the Eurocode recommendation, which onceagain does not appear as conservative.

Acknowledgements

The Authors would like to thank Dr. Gunter Schewe, from the German Aerospace Centre (DLR),Gottingen, Germany, and Eng. Guido Giorgetti, from Unitech Textile Machinery S.p.A., Prato, Italy,for the help in developing Set-up II.

References

Bearman, P. W., Gartshore, I. S., Maull, D. J., & Parkinson, G. V. 1987. Experiments on fluid-inducedvibration of a square-section cylinder. Journal of Fluids and Structures, 1(1), 19–34.

Cowdrey, C. F. 1963. A note on the use of end plates to prevent three-dimensional flow at the ends ofbluff cylinders. Aeronautical Research Council, Current Paper No. 683, HMSO, London.

EN 1991-1-4. 2010. Eurocode 1 - Actions on structures - Part 1-4: General actions - Wind actions.

Norberg, C. 1993. Flow around rectangular cylinders: Pressure forces and wake frequencies. Journalof Wind Engineering and Industrial Aerodynamics, 49(13), 187 – 196.

Novak, M. 1969. Aeroelastic galloping of prismatic bodies. Journal of Engineering Mechanics Divi-sion, 95(1), 115–142.

Novak, M., & Davenport, A. G. 1970. Aeroelastic instability of prisms in turbulent flow. Journal ofEngineering Mechanics Division, 96(1), 17–39.

Novak, M., & Tanaka, H. 1974. Effect of turbulence on galloping instability. Journal of EngineeringMechanics Division, 100(1), 27–47.

Obasaju, E. D. 1979. On the effects of end plates on the mean forces on square sectioned cylinders.Journal of Industrial Aerodynamics, 5(1-2), 189–190.

Parkinson, G. V., & Brooks, N. P. H. 1961. On the aeroelastic instability of bluff cylinders. Journal ofApplied Mechanics, 28(2), 252–258.

Parkinson, G. V., & Sullivan, P. P. 1979. Galloping response of towers. Journal of Industrial Aerody-namics, 4(3-4), 253–260.

Parkinson, G. V., & Wawzonek, M. A. 1981. Some considerations of combined effects of gallopingand vortex resonance. Journal of Wind Engineering and Industrial Aerodynamics, 8(1-2), 135–143.

Robertson, I., Li, L., Sherwin, S. J., & Bearman, P. W. 2003. A numerical study of rotational andtransverse galloping rectangular bodies. Journal of Fluids and Structures, 17(5), 681–699.