You are on page 1of 12

EACWE 5 Florence, Italy 19th 23rd July 2009

Flying Sphere image Museo Ideale L. Da Vinci

Fluid-structure interaction analysis with the ANSYS software in bridge aeroelasticity


1st G.Szab, 2nd J.Gyrgyi
Phd student 1 mr.gergely.szabo@gmail.com Szolnok, 5000, stks u. 4., HUNGARY Associate professor 2 gyorgyi@ep.mech.me.bme.hu

Keywords: bridge aeroelasticity, 3D flutter analysis, CSD, CFD, fluid-structure interaction.

ABSTRACT Bridge flutter assessment is usually carried out by means of wind tunnel tests on section models. Aerodynamic derivatives can be obtained either through free vibration or forced vibration methods. By using the Ads, equations for flutter stability can be solved and the critical wind speed can be determined as well. Nowadays, with a strong computational background, CFD simulations appear to be powerful rivals of the wind tunnel tests. Recently, a number of 2D simulations have been made in order to calculate the Ads of an ordinary bridge cross section according to Scanlan & Tomko (1971). In the work of Zhiwen et al. (2008) the Ads of a box bridge girder were calculated with the forced vibration method. Examples for the free vibration method can also be found in literature. In case of the section modeling approach, it is difficult to capture the rather complex three-dimensional coupling of the bridge deck motion and the fluid flow around it. For bridge flutter prediction it is hard to find 3D coupled CSD-CFD simulations in literature, but for wing flutter analysis Wang et al. (2008) have remarkable results. This paper firstly presents 2D CFD simulations for a flat plate and a box bridge deck to obtain the Ads. Reasonably agreement was found with theoretical and measured values. Secondly, a three-dimensional coupled CSD-CFD simulation is introduced. At this stage of our research, preliminary results can be presented only.
Contact person: 1st Gergely Szab, Phd student, Szolnok, stks u. 4.,HUNGARY, +3630/327-9262, +3612262096 E-mail mr.gergely.szabo@gmail.com

1. TWO-DIMENSIONAL STUDIES In our two-dimensional studies two different problems were solved with the ANSYS-FLUENT 6.3.26 version. The main goal was to determine the Ads of the chosen cross sections, and validate the results afterwards. The first case is a flat plate, where analytical results are available, while the second case is a bridge deck, which was tested in wind tunnel. The forced vibration method was applied for both. 1.1 Oscillating flat plate

The problem of an oscillating plate in a uniform airflow is a well-known problem in fluid mechanics. Theodorsen (1935) showed an analytical solution for an arbitrarily moved flat plate. The results are the lift force and moment acting on the plate due to rotational or heave motions. In our study, a 0.12m wide flat plate was investigated. Rotational and heave motions were considered. The CFD mesh around the flat plate boundary can be seen in Figure 1.

Figure 1: mesh around the flat plate.

In the numerical solution, the flat plate is not entirely flat but there is a certain height of it. The maximum angular amplitude was 10, while the vertical displacement was 0.02m. The frequency of the oscillation was set to 6.00Hz. The time step was 0.00005sec. For modeling turbulence, many approaches were tried. In Figure 2 the lift force signal can be seen with the inviscid theory, the k- and the Spalart-Allmaras turbulence models and the Theodorsen theory. As it can be seen, the k- model over predicts the moment, while the others show great agreement with the theoretical solution. This case belongs to Ured=4.00 value.

Figure 2: lift force acting on the flat plate by using different turbulence models.

As the Theodorsen theory assumes inviscid flow, it is reasonable to undertake the simulations by assuming the same. For flutter prediction, results within a certain Ured range are required. The previously mentioned frequencies and motion amplitudes were constant but the inflow velocities were varied. From the force and moment signals the Ads could be extracted. The calculated (thick line) and the theoretical Ads can be seen in Figure 3. The agreement between the theory and simulation for H1*, H3*, A1*, A3* is excellent. However, in the case of the most important A2* values, the differences are noticeable for higher Ured values.

Figure 3: calculated and theoretical aerodynamic derivatives of the flat plate.

1.2

Bridge deck section test

In this section, results from a wind tunnel test have been compared with numerical solution. In the wind tunnel a 1:100 rigid section model was oscillated. A professional robot-arm, which can be seen in Figure 4, ensured the precise sinusoidal motions while force and moment signals were recorded. To check the results of the wind tunnel tests, a 2D mesh was made around the bridge shape. In Figure 5 the numerical mesh can be seen. On the left side the mesh details are shown. At the near region of the wall boundary the mesh has no distortion, so its good quality can be preserved. The motion of the bridge is handled by the distortion of the farther mesh region, which is of no great importance. The motions of the deck (rotation and heave) and the inflow velocity were varied in order to set the required reduced velocities. The test was made using end plates on just one and on both sides of the section. In our study the two ends plated case is included only.

Figure 4: sectional model attached to a professional robot arm.

Figure 5: numerical mesh around the bridge deck.

For recording the force and moment signals, a force receiver was placed between the bridge deck and the robot arm. Firstly, the forces and moments were recorded at different angles of attack without any motion of the bridge section. The force coefficients have been computed with the CFD simulation. Three different turbulence models were tried in order to analyze their effects on the results; the k-, the k- and the SpalartAllmaras models were included. In each case second order schemes were used. The number of iterations was chosen to 2000 in steady state simulations.

In Figure 6 the measured and the calculated force coefficients are shown. The inlet velocity in each case was 10m/s. The coefficients were calculated using the reference area of the bridge deck; the width was 0.2954m and the length was 0.50m.

Figure 6: measured and calculated force coefficients.

The comparison shows good agreement between results and measurement. The calculated drag coefficients are somewhat under predicted with both of the turbulence models. In small angle of attack the lift and the moment coefficients fit well. In the case of the moment coefficients differences between the three turbulence models are the biggest.

After the steady state results, the oscillating problem must be solved. In the CFD code it is possible to define the regions to be oscillated. In our CFD model the bridge boundary contour with a rigid mesh region were moved together with a prescribed sinusoidal motion. The initial state and an internal transient time state of the mesh can be seen in Figure 7. The corresponding velocity contour plot is displayed in Figure 8.

Figure 7: mesh around the bridge deck at two different time steps.

Figure 8: velocity contour plot around the bridge deck at two different time steps.

Figure 9: velocity contour plot around the bridge deck at two different time steps.

In Figure 9 the measured and the calculated moments against time on the shear center of the bridge deck is shown (Ured=6.42, rotational motion). In the unsteady calculations the k- turbulence model was used only. As in the wind tunnel test, the reduced velocities were set in the CFD simulations; the motion amplitudes, frequencies, inlet velocities were defined properly. The time step was set to 0.0002s in each case, which ensured the time step number of 500 within a bridge deck oscillation phase and the CFL<1 condition. As in the case of the flat plate, the aerodynamic derivatives were calculated. The calculated and measured Ads can be seen in Figure 10 where the thick lines represent the simulated result.

The agreement between the measurement and the simulation overall is reasonable. In the case of the A3* and H3* derivatives the coincidence is really good in lower reduced velocities. With enlargement of the reduced velocity the differences are larger. The curves of other derivatives calculated with CFD show the tendency well but there are certain differences. Such comparison can be found in the work of Larsen & Walther (1998).

Figure 10: calculated and measured aerodynamic derivatives of the bridge deck.

2. THREE-DIMENSIONAL STUDIES However, the 2D simulations are to be developed and validated, there is a possibility to exploit the 3D CSD-CFD facilities of the ANSYS software to study bridge flutter. Experiences in wind tunnel testing according to Lajos et al. (2006) and 2D CFD runs made by Gyrgyi & Szab (2007) were strongly required to start a 3D coupled simulation. The ANSYS 11.0 version was used with a FEM modeling facility for the bridge structure and the CFX module for modeling the fluid flow. These two physics can be combined using the MFX multi-field solver. As there are measurements for section models available only, a fictive wind tunnel model was considered which is being constructed. This model is a two meters long suspension bridge. The wind tunnel model and its CSD model have to provide the same dynamic behavior though it is not easy to accomplish. While the wind tunnel model consist of a core mechanical part bounded with a light material which gives the shape of the bridge, the CSD model is made up using shell elements, by means of the FEM. The material properties and the thickness value of the shell elements have to be adjusted in the case of the CSD model in order to reach similarity. In the fluid-structure interaction calculation the force-displacement transfer between the structure and the fluid flow happens on the bridge deck surface but the cable elements have a role on the structural behavior only. In Figure 11 the details of the FEM model can be seen. The FEM mesh consists of regular rectangular 4-node shell elements with six degrees of freedom per each node while the cable elements are modeled with link elements without bending capabilities. The surface of the shell must be stiffened with diaphragms at every single vertical cable joints.

Figure 11: FEM model of the aeroelastic bridge model.

The first four mode shapes of the bridge model can be seen in Figure 12. The first and the second modes are asymmetrical and symmetrical with pure vertical motion. The third one shows a horizontal motion of the bridge deck. The first torsion mode was the fifth one.

Figure 12: four mode shapes of the FEM model.

For the CFD calculations a relatively coarse mesh was made to reduce the computational efforts. The number of cells was around 200.000. The CFD mesh can be seen in Figure 13. The key momentum in the coupled simulation is that the fluid flow induces forces on the bridge deck, which deforms accordingly based on the FEM calculations. The deformation is fed back to the CFD mesh. The time step was set to 0.0008 sec and the k- turbulence model was applied. The end time was 0.48s, which needed 6 days run time on a four core 2.40 GHz computer with 4.0 GB RAM memory. The first torsion frequency of the model is 6.33Hz so within the run time, approximately three vibrating phases could be observed.

Figure 13: three-dimensional CFD mesh around the bridge deck.

The inflow velocity was varied from 5 to 10m/s. From a non-deformed initial state of the bridge, the developing of the flutter motion needs much time. Thus, the bridge deck was loaded with a torsion moment for a certain transient time and then it was released, and the free vibration was investigated afterwards. If the amplitudes do not grow, the system could be regarded as stable under these conditions. At the velocities of 5.0 and 7.5 m/s flutter does not occur. At the speed of 10.0 m/s the vibration amplitudes grew from the initial value. This state showed a typical flutter phenomenon.

In Figure 14 the flow around the deformed bridge deck is shown. As the cable elements do not take part in the solid-fluid coupling, they are not shown at all. In the total deformation the torsion of the bridge deck dominates coupled with a little vertical displacement.

Figure 14: streamlines around the deformed bridge deck.

Due to the rather difficult solution introduced above, there is a need to check roughly its reliability. For this purpose, the dynamic loading from the airflow was strongly simplified; for different angles of attack the lift forces and moments can be calculated assuming flat plate theory with the following expressions: F = 2 M =

U 2
2

L B sin( ) L B 2 sin( )

(1)

U 2
2 2

(2)

In the expressions above is the air density, U is the airflow velocity, L is the section length and B is the plate width. By using (1) and (2) expressions, a quasi-static approach can be applied. To do that, the dynamic equation of the bridge must be written:
Mx(t ) +

Kx(t ) + Kx(t ) = q (t ) 0 r

(3)

In (3), M and K are the mass matrix and the stiffness matrix respectively, q is the time dependent load vector, and 0r are the logarithmic decrement and the r-th circular frequency of the bridge. To solve (3), a three-dimensional simplified CSD model was made using beam elements instead of the time-consuming shell model in Figure 11. The dynamic properties of the beam model must coincide with that of the shell model. The mode shapes of the beam model were collected in a V vector. The 2.0 m long bridge deck was subdivided into 20 elements so there are 19 nodes. At each node, there are vertical and torsion degrees of freedom. By using the V vector and seeking the solution in the form of x=Vy, (3) yields: V T MVy (t ) +

V T KVy (t ) + V T KVx(t ) = V T q (t ) 0 r

(4)

Considering the following equations:


V T MV = E
2 2 2 V T KV = 01 ...... 0 r ...... 0 n

(5)
(6)

The (4) differential equation of motion reads:

&&r (t ) + y

& 0 r y r (t ) + 02r y r (t ) = v rT q(t ) = f r (t )

(7)

To solve (7), detailed description can be found in the work Bathe & Wilson (1976). The solution methodology is as follows; at every single time step the angle of attack relative to the airflow must be determined and for the next time step the lift force and the moment can be calculated using (1) and (2). This approach neglects the induced flow forces due to the motion of the deck. To take them into account, Starossek (1998) proposed a solution for a three-dimensional beam model including the Ads of an arbitrary bridge section. To calculate the critical wind flutter speed of the bridge model, the inflow velocity was increased step by step, and the free damped oscillation of the structure was recorded as in the case of the coupled CSD-CFD model. In Figure 15 the vertical motion of the middle point of the bridge deck is shown at the wind speed of 5.0m/s with the CSD-CFD and the simplified beam model. As it can be seen, both models show a damped free oscillation after releasing them under this condition.

Figure 15: vertical displacement of the two different models.

As the CSD-CFD coupled model, the beam model was used with increased wind velocities as well. At the speed of 10.0 m/s it also showed flutter-like condition. This means that for this bridge model the critical wind speed of flutter can be between 5.0 and 10.0 m/s. To find all the instabilities versus the flow velocities, at each flow velocities the CSD-CFD simulation should be done. The problem is that every run requires much time. In addition, the present coarse CFD mesh is not adequate for capturing the vortex shedding for instance. Naturally, considering the results from the two strongly different three-dimensional modeling techniques, perfect coincidences cannot be expected. Nevertheless, the agreement in order is very promising for the future work.

3. CONCLUSIONS To conclude, in this paper two-dimensional and three-dimensional approaches for bridge flutter analysis were presented. At first, the oscillating flat plate problem was handled with 2D CFD simulation by using the forced oscillation approach. The theoretical and the numerical Ads were compared and really good agreement was found except for one derivative. Secondly, a wind tunnel section test results were compared with that of a simulation. The main results of the CFD simulation were the force coefficients against the angle of attack and the time dependent forces and moments from which the Ads could be extracted. Acceptable agreement between measurement and simulation shows the niceties of the CFD simulations but great care must be taken in accepting its results. Many simulations must be done while investigating the effects of the mesh quality, time step size and the chosen turbulence models on the results. To develop our 2D simulations further; a 3D coupled CSD-CFD simulation was performed. In an attempt to make a validation in the near future, a fictive full aeroelastic wind tunnel model was considered. The FEM model of the bridge was made using shell elements on the border between the fluid and the solid domain. At the inlet of the CFD model, constant velocities were defined, and the free oscillating motion of the bridge deck was studied under uniform airflow. The methodology of finding the critical flutter speed was to apply an inlet velocity increasingly in different runs. When the motion amplitude started growing, the critical velocity was found. The critical wind speed was checked with a simplified beam model of the bridge combined with a simple load model for the wind. The guess value of the critical wind speed was in the same order with the two different threedimensional models. The presented 3D CSD-CFD simulation is rather crude, indeed, and cannot be regarded as a solution for 3D flutter prediction at this stage of our research. Nevertheless, in the near future, the aeroelastic wind tunnel model should be made so that the methodology can be validated and become a reliable tool in bridge aeroelasticity. ACKNOWLEDGEMENT The authors are grateful for the support of the Department of Fluid Mechanics (University of Technology and Economics, Budapest) and the CFD.hu Ltd., Hungary. REFERENCES
Scanlan R. H., Tomko J. J., (1971): Airfoil and bridge deck flutter derivatives. ASCE J. of Eng. Mech. 97, 1717-1737. Zhiwen Z., Zhaoxiang W., Zhengqing C. (2008): Computational fluid dynamic analyses of flutter characteristics for self-anchored suspension bridges, Front. Archit. Civ. Eng. China 2008, 2(3): 267-273 Wang Y., Lin Y. (2008): Combination of CFD and CSD packages for fluid-structure interaction, Journal of Hydrodynamics, 2008, 20(6):756-761 Theodorsen T., (1935): General theory of aerodynamic instability and the mechanism of flutter. TR 496, NACA. Larsen A., Walther J. H. (1998): Discrete vortex simulation of flow around five generic bridge deck sections. J. Wind Engng. And Industrial Aerodynamics 77-78, 591-602. Lajos T., Balcz M., Goricsn I., Kovcs T., Rgent P.,Sebestyn P. (2006): Prediction of wind load acting on telecommmunication masts #paper A-0206, pp.1-8, IABSE Symposium on Responding to Tomorrow's Challenges in Structural Engineering, Budapest. Gyrgyi J., Szab G. (2007): Dynamic calculation of reinforced concrete chimneys for wind effect using the different codes and analysing the soil-structure interaction, #paper 1371, pp.1-12, ECCOMAS Thematic Conference on Computational Methods in Structural Dynamics and Earthquake Engineering, Rethymno, Crete, Greece. Bathe K. J., Wilson E. L. (1976): Numerical methods in finite element analysis. Englewood Cliffs, New Jersey: Prentice-Hall, Inc., 528 pp, ISBN 0-13-627190-1 Starossek, U. (1998): Complex notation in flutter analysis, ASCE, Journal of Structural Engineering, Vol. 124, No. 8. Ansys Help, Release 11.0 Documentation for ANSYS.

You might also like