You are on page 1of 17

GAMM-Mitt. 28, No.

1, 73 89 (2005)

Wind loads on lightweight structures:


Numerical simulation and wind tunnel tests
E. Rank1 , A. Halfmann1 , D. Scholz1 , M. Gluck
2 , M. Breuer2 , F. Durst2 , U.
3
3
3
Kaiser , D. Bergmann , and S. Wagner
1

Lehrstuhl fur Bauinformatik, Technische Universitat Munchen, Arcisstr. 21, 80290


Munchen, Germany
Lehrstuhl fur Stromungsmechanik, Universitat Erlangen-Nurnberg,
Cauerstr. 4, 91058 Erlangen, Germany
Institut fur Aerodynamik und Gasdynamik, Universitat Stuttgart,
Pfaffenwaldring 21, 70550 Stuttgart, Germany

Received 27 October 2004, accepted 17 November 2004


Key words fluid-structure interaction, partitioned solution approach, finite element method,
finite volume method, arbitrary lagrangian-eulerian formulation, wind tunnel.
In this paper basic principles of a numerical computation of fluid-structure interaction will be
discussed. Starting point are standard software systems for the simulation of fluid dynamics
and structural dynamics. These program systems are adequately extended and coupled in a
partitioned solution approach. Results of the presented methods will be compared in detail
with wind tunnel experiments for a lightweight membrane construction.
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Introduction

For certain construction types the proof of safety under extreme windloads is one of the most
demanding tasks in construction engineering [1]. Difficulties arise for several reasons. Firstly,
the load itself has stochastic nature and can only be described with very limited accuracy, when
the classical methods of civil engineering having deterministic nature are used. Assumptions
on wind profile, wind direction and strength of wind have to be made being only very limitedly
in accordance with reality.
Furthermore, the interaction of wind and structure has to be considered. In the most simple case, only the direction of wind flow is influenced by a rigid construction. The pressure
distribution acts as a surface load on the structure [2]. It is already more complicated when
the construction is deformed in such a way that the flow regime is modified which means that
an interaction of fluid and structure occurs. Most difficult is the prediction of a dynamic interaction. Induced by the fluid flow, the stucture starts fluttering or galopping. This interaction
can lead to a complete failure of the structure. The most prominent example is the collapse
of Tacoma-Narrows Bridge [3] at a wind speed which was far below the speed of assumed
failure and occured because of a torsional oscillation of the bridge.

Corresponding author: e-mail: d.scholz@bv.tum.de, Phone: +49 89 289 25115, Fax: +49 89 289 25051
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


74

Wind loads on lightweight structures: Wind loads on lightweight structures

Not that spectacular but much more frequently happens failure of light textile structures.
For this type of construction wind loads are relevant. Usually this load is very difficult to
predict as critical states appear only if deformations are large, i.e. that the feedback of the
structural deformation to the fluid flow regime is decisive.
Today, there are typically two types of proofs of safety possible. In the first type, alleged
conservative load assumptions are made for the structural computation (see e.g. DIN 1055,
Eurocode 1). This approach does not consider the feedback of the deformation to the wind
flow. If more detailed load assumptions are desired, the interaction of fluid and structure
has to be investigated in wind tunnel experiments. Besides the typically very high costs of
these experiments, it is difficult to scale the model in such a way that all important physical
phenomena can be observed.
A complete numerical computation of the interaction of wind and lightweight structures is
not yet used in civil engineering practice. One major reason is that already the formulation of
a realistic model leads to very significant difficulties. Furthermore, a reliable description of
the coupling phenomena is connected with a large number of mathematical and algorithmic
difficulties. Moreover, nearly all practically relevant cases are strictly three dimensional. A
reduction to two-dimensional flow around cross sections will lead only in very special cases
to realistic results. Yet, only the computation of a (typically turbulent) flow field around
large constructions is a very considerable challenge which can only be attacked by extreme
computational power.
Despite these general difficulties, significant progress in development and validation of
suitable coupling methods have been made during the past few years (e.g. [5, 6, 7, 8]). Considering this progress and assuming further increase of computational power gives rise to the
expectation that in only a few years numerical simulation will be a true alternative to wind
tunnel experiments in civil engineering practice. Therefore, the goal of this paper is to review
the most important algorithmic principles of fluid-structure interaction and to report on a detailed comparison of wind tunnel experiment and numerical simulation. As an example a tent
construction is considered. In order to get a meaningful result, the wind tunnel model itself
was investigated by the numerical simulation and not the original real life construction.

Numerical simulation of fluid-structure interaction

2.1 The coupled equations


The interaction of wind and structure can be described by a multi field problem. The equations
of structural dynamics (CSD) are coupled to those of fluid dynamics (CFD). Interaction occurs
as the fluid flow imposes normal and shear stresses as loads to the structure which deforms
and thus changes the geometrical shape of the flow regime. If deformations are large, flow
and loads on the structure may change significantly. In general, each field problem itself is
nonlinear and dynamic. Let us first consider a fixed flow regime. Fluid mechanics for an
incompressible Newtonian fluid are described by the unsteady Navier-Stokes equations
ui
xi
(ui uj )
ui
+

t
xj

= 0,
p
2 ui
=
+
,
xi
x2j

(1)

c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim




GAMM-Mitt. 28, No. 1 (2005)

75

Fig. 1 x-y-cut through the CFD grid of a plate at the beginning of the simulation (left) and in a deformed
state (right)

where the three velocity components u i and the pressure p are unknown. describes the
density and the dynamic viscosity of the fluid. Furthermore, suitable initial and boundary
conditions have to be defined.
In contrast to this classical flow problem, fluid-structure interaction demands for a description of the flow on a deforming flow regime. This can be accomplished by adding an additional
convective term to the Navier-Stokes equations which describes the motion of the flow regime.
The corresponding Arbitrary Lagrangian-Eulerian- (ALE-) formulation is described in detail
e.g. in [9]. The motion of the structural boundary is transferred to the spatial flow grid by
the solution of an additional field equation. In Fig. 1, for example, this grid motion is shown
for a two-dimensional problem. In the present project, discretization of the fluid mechanical
problem is performed using finite volume methods (e.g. [10, 11]), where extensions for the
simulation on moving grids have been implemented.
Structural dynamics for fluid-structure interaction has to consider large deformations in
general:
1 T
(F F I) ,
2
,
Div(F S) + 0 b = 0 d
E=

(2)

S = CE.
F denotes the deformation gradient, I the unit tensor and E the Green-Lagrangian strain
tensor. The symmetric 2 nd Piola-Kirchhoff stress tensor S, the density and volume and
are used to describe the equilibrium conditions. Compressible and
inertia forces b and 0 d
elastic material properties are assumed defining a constant elasticity tensor C. Suitable initial
and boundary conditions are to be defined for the structural problem, too.
After spatial discretization with finite elements, a set of coupled ordinary differential equations is obtained:
+ D d + R(d) = F
Md

(3)

denotes the acceleration vector, d the velocity vector and d the displacement vector collectd
ing all degrees of freedom of the finite element approach. M is the mass matrix, D the (linear)
damping matrix and R(d) is the displacement vector depending nonlinearly on the interior
forces. F describes surface loads imposed on the boundary of the structure and is varying
in time by the changing fluid field. It should be noted that all computations being presented
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


76

Wind loads on lightweight structures: Wind loads on lightweight structures

in section 4 use a damping matrix D = 0. Thus, only the damping of the fluid imposed by
surface loads onto the structure is considered. The commercial finite element code ASE of
SOFiSTiK AG [12] is used as a structural dynamics software for the fluid-structure interaction
simulation.
2.2 The partitioned solution approach
In principle, the solution of the two sets of field equations (1) and (2) can be obtained by
the two following approaches. In a monolithic approach both systems are combined. Fluid
velocities and structural deformations are considered as one set of unknowns and the complete set of nonlinear dynamic equations is solved by e.g. a finite element or a finite volume
approach. As a consequence, very significant software development has to be performed as
this coupled problem cannot be solved by only a simple extension of an existing structural
or fluid mechanical software code. The second, much more frequently used method is the
so-called partitioned solution approach. The coupled solution is obtained iteratively and has
been investigated in many different variants (e.g. [13, 14, 15, 16, 17]). Formally, the two field
problems can be written in operator notation:
Lf (vfi+1 , dis , t)
i+1
Ls (di+1
s , vf , t)

= ff (t) Csf (t, ds (t)i ) ,


= fs (t) Cf s (t, vf (t)

i+1

(4)
).

(5)

Index f is used for the fluid, s for the structure. L s and Lf denote the nonlinear differential
operators for structure and fluid dynamics, respectively. v f contains all variables which are
necessary to obtain pressures and shear forces by fluid mechanics, d s denotes the unknown
structural displacements. It should again be noted that only by the solution process itself
the shape of the flow regime and the deformation of the structure are obtained. Therefore,
the differential operators L f and Ls depend nonlinearly on the variables d s and vf . On the
right hand sides, the terms f f and fs are independent of the coupling, whereas C sf (t, ds (t))
and Cf s (t, vf (t)) are coupling terms. For the fluid mechanical problem (4) C sf (t, ds (t))
is the convective term of the ALE-formulation and for the structural mechanical problem (5)
Cf s (t, vf (t)) is the surface load resulting from the fluid flow. Considering equations (4) and
(5) it is obvious that a partitioned solution approach is formally related to a block Gauss-Seidel
iteration for solution of an equation system. Extensions of this method have been described
in [18], using e.g. predictor-corrector methods. Different iteration procedures e.g. relaxation
methods with an optimal choice of the relaxation parameters are described in [19, 20]. The
algorithmic structure of this coupled procedure is displayed in Fig. 2.
In an outer loop the dynamic time dependence is discretized, the middle loop describes the
iteration of equations (4) and (5). The inner loop describes the iteration being necessary to
solve the nonlinear fluid or structural sub-problem in each individual sub-step. The advantage
of this iterative coupling is obvious. Interaction is modeled only by exchanging loads and displacements between the two sub-problems. Therefore, both sides can use nearly unmodified
standard programs for structural and fluid dynamics, respectively, which are embedded in the
middle and exterior loop of a frame algorithm. The necessary communication between the
programs can be implemented by a suitable coupling software (e.g. MpCCI [21]), which also
supports the interpolation of data between computational grids being described in the next
section.
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

77

fluid

inner
iteration

structure

CFD
fluid solution
wind loads
CSD

inner
iteration

stucture solution
FSI

displacements

FSI

converged global solution


timestep

timestep

Fig. 2 Partitioned solution approach

2.3 Discretization of interfaces


Due to the different types of methods and programs used in the partitioned solution approach
for fluid and structural discretization, the description of the interface geometry is in general
different in the two sub-problems.
As shown schematically in Fig. 3 the interface is described by the fluid flow simulation e.g.
by a block-structured grid being used by a finite volume discretization. On the structural side,
however, discretization can be performed by an automatic mesh generator [22] using, e.g.,
quadrilateral finite elements. As both meshes approximate the same interface geometry, yet
use different discretizations, interpolation of all grid based quantities are necessary for the data
transfer. In our implementation a conservative interpolation procedure was used, guaranteeing
that after data transfer the integral of loads on both computational grids are equal (see [23]).
In the following algorithm, instead of pressure and shear forces, equivalent nodal loads are
exchanged between the simulation codes.
Firstly, for each CFD interface node S j or its projection Sj the closest element of the
CSD mesh is searched. Projection is necessary due to the different discretizations of curved
surfaces. Fig. 4 shows the described correspondence of a CFD node to a CSD element. This
first step, the so-called pairing of CFD nodes and CSD elements was implemented efficiently
using an Octree-based search algorithm [7]).

As for bilinear shape functions N i in each quadrilateral element i=4
i=1 Ni = 1 holds, it is
obvious that the described method is conservative. Integrals of loads on both computational
grids are equal (CFD grid with n
f nodes and CSD mesh with n
s nodes):
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


78

Wind loads on lightweight structures: Wind loads on lightweight structures

P,

CFD surface grid

CSD grid

ds

Fig. 3 Data transfer between the computational grids

CFD grid

CSD grid
Sj
Sj

Fig. 4 Pairing of CFD nodes and CSD elements

n
s

i=1

Fs(i) =

n
f
n
s 

i=1 j=1

Ni (j , j ) Ff (j) =

n
f


Ff (j)

(6)

j=1

Fs(i) denotes nodal forces of the structural mechanical description, F f (j) those on the
surface of the fluid flow.
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

79

Fig. 5 Membrane umbrella

Models

3.1 Wind tunnel model


Starting point of the present comparison between simulation and experiment was a textile roof
for temporary and mobile use (Fig. 5). The shape was similar to an umbrella with a diameter
of 30 m, supported in the middle by a mast and tensed with radial guys. The rigidity of the
membrane that was needed for stability was achieved by a special curvature in combination
with pre-loading. The pre-loading could be adjusted by either extending the central mast or
by varying the tension of the 32 radial guys.
Model measurements were accomplished in the boundary layer wind tunnel of the IAG
[26, 27]. It is an open-circuit type with a test section of 2.5 2.0 m 2 . The flow processing
section of the tunnel is equipped with exchangeable devices for generation of certain velocity
profiles, turbulence intensities and turbulence structures. Fig. 6 shows a schematic sketch of
the tunnel, the inserted model with downstream view and the velocity and turbulence profiles,
respectively.
A model scale of 1:50 has been chosen, yielding a model diameter of d = 60 cm. The
center of the membrane was fixed by a conical plate rigidly attached to the mast. The height
of the supports arranged in circumferential direction was 8.3 cm in model scale.
Due to restrictions in manufacturing of the membrane, not all details of the prototype could
be modeled on a scale of 1:50. For the same reason the layout of the pre-cut parts also had
to be simplified to a certain degree. To achieve an optimal reproduction of the original shape
both multi-part layouts and single part layouts have been tested. Membrane layouts consisting
of several parts turned out to be less accurate because of the adding of minor errors in each
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


80

Wind loads on lightweight structures: Wind loads on lightweight structures

air intake

windows

xxxXX

contraction
test section

end plate
propeller

diffusor

air outlet

500
Cv
Tuu

z [ mm ]

400
300
200
100
0

0.2 0.4 0.6 0.8


Cv , Tu [ - ]

Fig. 6 Schematic side view of the boundary layer wind tunnel at the IAG, wind tunnel model in the test
section with view in flow direction and velocity and turbulence profile

part. For the model used in this investigation the original shape has been modeled using a
simplified layout made from one piece.
Special efforts were undertaken in order to be able to adjust the pre-loading as evenly
as possible and in arbitrary stages. To apply loads to the outer rim of the membrane thin
filament has been attached to the membrane at 32 locations. The filament has been deflected
by supports with guide rollers to the underneath of the wind tunnel, where defined weights
could be applied. Ball bearings were used in the rollers to keep friction as small as possible.
Force transducers in the supports were used to ensure equal loading. During the tests the
filaments have been fixed.
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

81

For the measurement of the forces at the anchoring wires 6 industrial load cells were used.
The signals were amplified over a carrier frequency amplifier and filtered before data acquisition. Due to the rotational symmetry of the model each of the 32 anchoring wires could be
measured by its turn.
The wind tunnel tests were accomplished without flow processing devices. Due to the distance between intake and model position of approximately 8 m a turbulent boundary layer
builds up on the wind tunnel floor. In the test section the boundary layer thickness is approximately 200 mm, covering the model completely. The turbulence intensity at the height
of the membrane rim was approximately 5%, decreasing upward to a constant value of 2%.
The deformation of the membrane was measured contactlessly with two laser triangulators at
two points. By turning of the model the deformation could be measured on two circular cuts
at a time. By changing the radial position of the transducers several circular cuts could be
recorded.

3.2 Structure mechanical model


As material for the membrane a rubber sheet was selected, whose material properties have
been determined at the IAG: thickness t = 0.2 mm, linearized modulus of elasticity E Mem =
3 MPa, transverse extension number Mem = 0.2 and density Mem = 1100 kg/m 3 . For the
outside rim of the membrane a silicone tube with an outer diameter of d Rim,out = 1.8 mm and
a modulus of elasticity of E Rim,out = 4.3 MPa was used.
For the anchoring wires a kite string (e.g. Dyneema SK 60) with d Anchor = 0.3 mm,
EAnchor = 2260 MPa, Anchor = 0.2 and Anchor =1340 kg/m 3 was used. Due to their high
stiffness (compared to the membrane) the deformations of the mast and the supports could be
neglected. The material parameters used for the structural elements are described in detail in
[7].
Before being able to simulate the fluid-structure interaction, the geometric shape of the
unloaded membrane has to be found in a numerical form finding process, where geometric
nonlinear structural behavior has to be assumed (see [28]). The form finding process starts
with a displacement of the singular center point, so that a height of the membrane of h z = 8.5
cm and a total height of the membrane roof of H z = 16.8 cm with respect to the supports of
the columns is achieved. Furthermore, the prestress in the membrane is described with a prescribed ratio of tangential/radial prestresses with increasing radial stresses towards the center
point of the membrane. The prestresses in the outside rims and anchoring wires are chosen
accordingly, which results in:
FMem,rad
FMem,tang
FRim,out
FAnchor

= 8.50 N/m
= 2.98 N/m
= 0.80 N
= 0.50 N
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


82

Wind loads on lightweight structures: Wind loads on lightweight structures

Fig. 7 Part of the block-structured finite volume grid

3.3 Fluid dynamical model


For the CFD simulation a k- model [24] is used. Based on a measured turbulence degree T u
= 0.05 the turbulent kinetic energy at the inflow boundary can be computed to
k(z) =

3
T u2 U (z)2 .
2

(7)

In accordance to experimental measurements the velocity field U (z) at the inlet of the simulation domain was described by an exponential wind profile. The dissipation rate at the inlet
is given by
(z) =

k(z)3/2
.
lc (z)

(8)

Following [25] the characteristic length scale l c is chosen to


lc (z) =

z
C03

with C0 = 0.427

and

= 0.4.

(9)

Using density and viscosity of air ( = 1.14 kg/m 3 and = 18.24 10 6 kg/(m s)) the
Reynolds number referring to the height of the membrane h z is obtained to Re = 65 875 and
Re = 87 125 for the reference velocities U ref = 12.3 m/s and 16.2 m/s, respectively.
The geometric description of the structural model is the starting point for the definition of
a CFD model. The computational domain is discretized by a block-structured finite volume
grid into 1 228 800 control cells. Fig. 7 shows a part of this discretization. The projection of
the grid to the membrane surface is plotted in Fig. 8.
As we only consider the stationary fluid-structure interaction, only the gray part of the
algorithm plotted in Fig. 2 is relevant for this example. Further investigations using a fully
dynamic interaction and a large-eddy simulation approach for the CFD computation are described in [8].
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

83

Fig. 8 CFD interface discretization (3072 quadrilaterals)

Comparison of wind tunnel and simulation results

4.1 Deformation of the structure


Fig. 9 shows the deformed wind tunnel model and Fig. 10 the corresponding deformed numerical model. Simulation and experiment are in good qualitive agreement. To investigate
the system behavior of wind tunnel and numerical model in more detail the following three
different cases are examined. The upper row in Fig. 11 shows measured displacements of the
wind tunnel model and below the corresponding results of simulation starting from an inflow
velocity Uref and a given prestress F Anchor (Fig. 11, left). In a first step the prestress is
doubled yielding a reduction of the structural deformation (Fig. 11, center). In a second step
the configuration is changed again by increasing the inflow velocity. The system reacts by an
increase of deformation (Fig. 11, right). In all three cases good correspondence of wind tunnel experiments and numerical simulation can be observed. For a more detailed comparison
the deformed structures are shown in cutlines in Fig. 12 (in flow direction and vertical to the
flow direction). Good agreement can be observed again even when small differences in the
absolute displacement values are obtained. Concerning the high complexity of the model and,
e.g., the inaccuracies in the definition of prestresses in the wind tunnel model, the agreement
of the simulation and experiment is highly satisfactory.
4.2 Windloads
Fig. 13 shows the horizontal load components in the cables for the three cases of comparison.
Forces are plotted over the circumferential angle. 0 degree corresponds to the luvward part
and 180 degrees to the leeward part. Qualitatively, again good agreement can be observed,
although there are quantitative differences in the prestresses at individual columns. This discrepancy is due to the simplifications which had to be made for the experiment as well as for
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


84

Wind loads on lightweight structures: Wind loads on lightweight structures

Fig. 9 Deformed wind tunnel model

Fig. 10 Deformed structural model during simulation

the numerical approach. Further investigations are necessary to clarify the exact source of
these discrepancies.

Conclusions and future work

The most significant result of the presented investigation is the good agreement of experiment
and simulation, being dependent on the exactness of the definition of the corresponding models. A very significant advantage of numerical simulation is the possibility of coupling it to
CAD models and automatic mesh generation and thus being able to modify shape and parameters of the structure, easily. On the other hand, modification of a wind tunnel model is in
general time consuming and expensive. Furthermore, in a physical experiment only pointwise
or integral quantities can be measured, whereas numerical simulation gives a very detailed
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

85

Fig. 11 System response

spatial solution of the quantities of interest. Yet, the very high requirements on computational effort for a numerical simulation of fluid-structure interaction is still very demanding
and today only possible in academic research projects. Furthermore, when stepping to dynamic wind-structure interaction in a highly turbulent atmospheric boundary layer even more
computational power would be necessary. Considering this situation, the wind tunnel experiment could investigate a complete configuration with different prestresses over a time span
corresponding to a storm of 36 h on the original structure. Future increase of computer power
and additional developments in computational mechanics let yet expect that in only few years
even complex practical problems of wind-structure interaction can be numerically simulated.
These computations will be a highly welcome alternative to expensive and rather inflexible
experiments in wind tunnels.
One important final experience of our investigations should be mentioned. The definition
of the right model of reality is much more difficult for a fluid-structure problem than for
classical single field problems. Even seemingly small modifications of load conditions, of
the shape of the membrane or the inflow conditions can lead to very significant differences in
the systems answer. The necessity of a very detailed model description can be seen in the
numerical simulation as well as in the wind tunnel experiment and underlines that multi-field
problems need still a lot of research in computational and experimental mechanics.
Acknowledgements The authors would like to thank the Bayerische Forschungsstiftung for funding
this work in the Competence Network for Technical, Scientific High Performance Computing in Bavaria
(KONWIHR). Furthermore the support ofOFiSTiK AG, especially of Dr. Bellmann and Dr. Katz, is
acknowledged.

c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim




86

Wind loads on lightweight structures: Wind loads on lightweight structures


0.32 N 84 Pa: numerical simulation
0.32 N 84 Pa: wind tunnel test
0.65 N 84 Pa: numerical simulation
0.65 N 84 Pa: wind tunnel test
0.65N 150 Pa: numerical simulation
0.65N 150 Pa: wind tunnel test

150

-300

z [mm]

-200

100
50
0

-100

0
x [mm]

100

200

300

100

200

300

z [mm]

150
100
50
0
-300

-200

-100

0
y [mm]

Fig. 12 Comparison of deformed systems

References
[1] K. Zilch , Ein anschauliches Lastkonzept fur Hochhauser in boigem Wind, Habilitationsschrift, Technische Hochschule Darmstadt (1983).
[2] P. Droll, R. Sieber, A. Cozzani and M. Schafer, Numerical investigation of the performance of a deep space antenna under environmental loads, Bauingenieur, January,
pp. 712 (2002).
[3] http://www.enm.bris.ac.uk/anm/tacoma/tacoma.html
[4] K. J. Bathe, H. Zhang, Finite element developments for general fluid flows with structural interactions, International Journal for Numerical Methods in Engineering , vol. 60,
pp. 213232 (2004).
[5] W. A. Wall, Fluid-Struktur-Interaktion mit stabilisierten Finiten Elementen, Dissertation, Institut fur Baustatik, Universitat Stuttgart (1999).
[6] A. Halfmann, E. Rank, M. Gluck, M. Breuer, F. Durst, J. Bellmann and C. Katz, Computational Engineering for Wind-Exposed Thin-Walled Structures, Lecture Notes in Computational Science and Engineering, vol. 21, pp. 6370, Springer, Berlin, Heidelberg,
New York (2002).
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

2.5

87

Experiment
experiment
num. Simulation
numerical
simulation

F [N ]

1.5

0.5

Fv=0.32N V=12.3m/s
Fv=0.65N V=12.3m/s
Fv=0.65n V=16.2m/s
0

45
90
135
horizontale
Richtung
der Abspannung
horizontal
direction
of anchor [][ ]

180

Fig. 13 Comparison of cable forces

[7] A. Halfmann, Ein geometrisches Modell zur numerischen Simulation der FluidStruktur-Interaktion windbelasteter, leichter Flachentragwerke, Dissertation, Lehrstuhl
fur Bauinformatik, Technische Universitat Munchen (2002).
[8] M. Gluck, Ein Beitrag zur numerischen Simulation von Fluid-Struktur-Interaktionen
Grundlagenuntersuchungen und Anwendung auf Membrantragwerke, Dissertation,
Technische Fakultat, Universitat Erlangen-Nurnberg (2002).
[9] I. Demirdzic and M. Peric, Space Conservation Law in Finite Volume Calculations of
Fluid Flow, International Journal for Numerical Methods in Fluids, vol. 8, pp. 1037
1050 (1988).
[10] C. Bartels, M. Breuer, K. Wechsler and F. Durst, CFDApplications on ParallelVector
Computers: Computations of Stirred Vessel Flows, Computers & Fluids, vol. 31, no. 1,
pp. 6997, Elsevier Science Ltd, Oxford (2001).
[11] F. Durst and M. Schafer, A Parallel BlockStructured Multigrid Method for the Prediction of Incompressible Flows, International Journal for Numerical Methods in Fluids,
vol. 22, pp. 549565 (1996).
[12] C. Katz and J. Bellmann, ASEHandbuch, SOFiSTiK AG (19881995).
[13] K. C. Park and C. A. Felippa, Partitioned Analysis of Coupled Systems, in Computational Methods for Transient Analysis (Eds. Belytschko, T. & Hughes, T.J.R.), pp. 157
219 (1983).
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


88

Wind loads on lightweight structures: Wind loads on lightweight structures

[14] C. Farhat and M. Lesoinne, On the Accuracy, Stability, and Performance of the Solution of Three-Dimensional Nonlinear Transient Aeroelastic Problems by Partitioned
Procedures, AIAA-96-1388-CP, pp. 629641 (1995).
[15] A. C. Frank, Organisationsprinzipien zur Integration von geometrischer Modellierung,
numerischer Simulation und Visualisierung, Dissertation, Lehrstuhl fur Ingenieuranwendungen in der Informatik, Numerische Programmierung, Fakultat fur Informatik, Technische Universitat Munchen (2000).
[16] M. Gluck, M. Breuer, F. Durst, A. Halfmann and E. Rank, Computation of Fluid
Structure Interaction on Lightweight Structures, International Journal of Wind Engineering and Industrial Aerodynamics, vol. 89, no. 1415, pp. 13511368 (2001).
[17] M. Gluck, M. Breuer, F. Durst, A. Halfmann and E. Rank, Computation of Wind
Induced Vibrations of Flexible Shells and Membranous Structures, Journal of Fluids
and Structures, vol. 17, pp. 739765 (2003).
[18] M. Lesoinne and C. Farhat, Improved Staggered Algorithms for the Serial and Parallel
Solutions of Three-Dimensional Nonlinear Transient Aeroelastic Problems, Proceedings
of 4th World Congress on Computational Mechanics, Buenos Aires, pp. 483485 (1998).
[19] W. A. Wall, D. P. Mok and E. Ramm, Partitioned Analysis Approach for the Transient,
Coupled Response of Viscous Fluids and Flexible Structures, Proceedings of European
Conference on Computational Mechanics, Munchen (1999).
[20] D. P. Mok, Partitionierte Losungsansatze in der Strukturdynamik und der Fluid
StrukturInteraktion, Dissertation, Institut fur Baustatik der Universitat Stuttgart (2001).
[21] R. Ahrem, M. G. Hackenberg, P. Post, R. Redler and J. Roggenbuck, MpCCI Mesh
based parallel Code Coupling Interface, Institute for Algorithms and Scientific Computing (SCAI), GMD, http://www.mpcci.org/ (2000).
[22] E. Rank, M. Rucker and M. Schweingruber, Automatische Generierung von FiniteElement-Netzen, Bauingenieur, vol. 69, pp. 373379 (1994).
[23] C. Fahrhat, M. Lesoinne and P. LeTallec, Load and Motion Transfer Algorithms of
Fluid/Structure Interaction Problems with Non-Matching Discrete Interfaces: Momentum and Energy Conservation, optimal Discretization and Application to Aeroelasticity,
Computational Methods in Applied Mechanics and Engineering, vol. 157, pp. 95114
(1998).
[24] B. E. Launder and D. B. Spalding, The numerical computation of turbulent flows, Computer methods in Applied Mechanical Engineering, vol. 3, pp. 269289 (1974).
[25] J. R. Garratt, The atmospheric boundary layer, Cambridge University Press, Cambridge,
pp. 71 ff (1992).
[26] S. Wagner, D. Bergmann and U. Kaiser, Kraftmessungen an einem Membranschirm
im Original unter Windeinfluss, Bericht, Institut fur Aerodynamik und Gasdynamik,
Universitat Stuttgart (2001).
c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


GAMM-Mitt. 28, No. 1 (2005)

89

[27] U. Kaiser, Windwirkung auf schwach vorgespannte Membranstrukturen am Beispiel


eines 30m - Membranschirmes, Dissertation, Fakultat Luft- und Raumfahrttechnik, Universitat Stuttgart (2002).
[28] J. Bellmann, Membrantragwerke und Seifenhaut - Unterschiede in der Formfindung,
Bauingenieur, March, pp. 118123 (1998).

c 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim




You might also like