1. Introduction
Suspension bridges are long, slender and flexible structures which are very sensitive to dynamic actions induced by wind, which potentially cause a variety of instability phenomena. The importance of wind-resistant design for these structures has been highly recognized and has led to many research works and investigations on bridge aerodynamics. Some historical views of long-span bridge aerodynamics are presented in [
1,
2].
The first studies were addressed on the dynamic of suspended cables, pertinent to suspension bridges but without the inclusion of the stiffening girder, see Irvine and Caughey [
3] and Irvine [
4]. Since the 1980s, starting with Hagerdon and Schafer [
5] and Luongo et al. [
6,
7], there has been a large number of studies on the free and forced nonlinear vibrations of suspended cables (see the review of Rega, in [
8,
9]). In particular, some papers investigated wind-induced aeroelastic instabilities, as [
10,
11,
12].
After the Tacoma Narrows Bridge failure in 1940, major efforts were made to understand the dynamic of suspension bridge, including the effects of the stiffening girder. Bleich [
13] modeled the stiffening girder as a flexural beam attached to the elastic cables by inextensible hangers. Additional contributions to the continuum model were made by Pugsley [
14]. Abdel-Ghaffar [
15] extended the approach to include coupled vertical–torsional vibrations and the effects of cross-sectional distortion. Hayashikawa and Watanabe [
16] and Kim et al. [
17] enriched the linearized continuum model including the effects of shear deformation and rotary inertia of the stiffening girder, by concluding that these effects are relatively small and only affect the higher modes of vibration. More recently, in [
18] the classical continuum model for the free linear vertical vibrations of a suspension bridge is re-examined.
Remarkable attention has been addressed to the aeroelastic behavior of long-span bridges, and especially to the aeroelastic instability, as galloping (1 d.o.f) or flutter (2 d.o.f.), which may cause catastrophic effects, like in the famous Tacoma Narrows bridge. The modern era of bridge aeroelasticity was launched by Farquharson [
19]. Bleich [
20], for the first time, analyzed the bridge flutter behavior by exploiting the Theodorsen theory of thin airfoils, with the potential flow assumption. During the same period, empirical formulas were proposed, based on the bridge structural properties only, to directly obtain the critical flutter wind velocity without involving any consideration of the flow field (e.g., [
21]). Scanlan and his colleagues [
22] introduced a semiempirical linearized analysis, referred to as “unsteady theory”, based on the so-called “flutter derivatives”, which is nowadays still popular among researchers. In [
23], advances in aeroelastic analyses of suspension bridges are described, with focus on the numerical simulations carried out for a long-span suspension bridge, using finite difference and discrete vortex methods. More recently, many linear and nonlinear aeroelastic analyses for cable-supported bridges were developed, e.g., [
24,
25,
26,
27,
28]. In [
24], a nonlinear aerodynamic force model and associated time domain analysis is performed to predict the aeroelastic response of bridges under turbulent winds. In [
25], a numerical method for full-bridge aeroelasticity is presented, based on unsteady cross-sectional wind loads and on the finite-element modeling of the structure. In [
26], a fully nonlinear model of suspension bridge is proposed to describe overall three-dimensional motions. Torsional-divergence and flutter are studied by numerical methods, able to tackle the coupled nonlinear partial-differential equations of motion. In [
27], after a review of the existing linear and nonlinear analyses, two advanced nonlinear models are introduced to study the aeroelastic effects under smooth/turbulent wind conditions: (a) artificial neural network and, (b) Volterra series-based models. In [
28], the limit-cycle oscillations exhibited by long-span suspension bridges in post-flutter condition are investigated. The governing aeroelastic partial differential equations are solved by the Faedo-Galerkin method.
Most of the studies about the aerodynamic instability of the bridge were devoted to flutter phenomena, according to which the aeroelastic instability involves two coalescent modes, one torsional, the other flexural. This is the case of bridges with low torsional stiffness, for which the natural torsional and flexural frequencies are close each other, prone to coalesce under aerodynamic excitation. On the contrary, bridges with high torsional stiffness (e.g., box-girder bridges) have the fundamental torsional frequency much larger than the flexural one, so that the aerodynamic instability exclusively occurs in the flexural mode (galloping instability). This class of bridges also attracted some interest, justified by the fact that, after the Tacoma collapse, many existing bridges were torsionally stiffened and the new bridges were designed with stiffer trusses (e.g., the Verrazzano bridge in New York, Tago bridge in Lisbon or the Japanese bridge Akashi Kaikyo). Many works in literature dealt with the galloping instability separately in the beam (see, e.g., [
29,
30,
31,
32,
33]) and in the cable (see, e.g., [
34,
35,
36,
37]). With reference to the class of bridges with high torsional stiffness, the interest was addressed mainly to the galloping control (see, e.g., [
38,
39,
40]).
In this paper, the interest is focused on galloping of suspension bridges. Differently from the current research trend, according to which fine discretized models are used to analyze particular bridges, here the aeroelastic behavior is investigated via a continuous model, in order to study families of bridges, and to identify the key dimensionless parameters which govern the mechanical behavior. The main innovative aspect, therefore, relies in directly attacking the continuous problem by perturbation methods (similarly to what done by the same authors in [
41]), which leads to approximate formulae, suitable for preliminary design. A simple continuous model of suspension bridge, subjected to steady wind flow, is proposed to analyze the aeroelastic in-plain instability (galloping). Linear visco-elastic beam and cable are considered, coupled by vertical suspenders, modeled as axially rigid links uniformly distributed. Both external and internal damping are accounted, both for the beam and cable, according to the Kelvin-Voigt rheological model. A steady and uniform wind flow is applied exclusively to the beam. The aeroelastic effects of the wind are evaluated via the classical quasi-static theory (a discussion about strengths and weaknesses of the method is reported in [
42]). First, the linear free dynamic of the undamped model is studied; then, the nonlinear bifurcation problem is tackled by the Multiple Scale Method (MSM) which naturally reduces the infinite-dimensional system to the proper finite-dimensional Center Manifold. From the bifurcation equation, the limit cycles and their stability are evaluated as function of the mean wind velocity. The asymptotic results are validated via numerical time-integration of ordinary differential equations derived by in-space finite differences. A case study of suspension bridge is analyzed.
The paper is organized as follows. In
Section 2, a continuous visco-elastic model of standard single-span suspension bridge under uniformly distributed steady wind flow, is formulated. In
Section 3, the free dynamic of the undamped problem is addressed. In
Section 4 the Multiple Scale Method is directly applied to the partial differential equations, to get the bifurcation equation governing the motion close to a Hopf bifurcation. In
Section 5, a finite-dimensional model is formulated. In
Section 6, the numerical outcomes are presented. Finally, in
Section 7 the main findings of the work are summarized. An
Appendix A, containing details about damping parameter identification, closes the paper.
2. Continuous Model
A standard single-span suspension bridge is considered, consisting of a main cable, a stiffening girder, equispaced hangers (or suspenders) and two supporting towers (
Figure 1).
The main cable, modeled as uniform, elastic and perfectly flexible (no bending stiffness) is connected to the tip of towers, taken to be infinitely rigid. The girder is modeled as an Euler-Bernoulli beam, simply supported at both ends. The structure is subjected to a steady and uniform wind flow of velocity U. The dynamics of the structure are studied in the vertical plane, while the out-of-plane motions are ignored.
The main cable is modeled according the Irvine’s theory [
3], holding for sag-to-span ratios up to about 1:10. The cable assumes a parabolic profile under the initial dead load, and experiences vibrations in which the tangent component is statically condensed. Beam and cable are coupled by vertical hangers, here assumed to be massless, inextensible and continuously distributed along the span of the bridge (curtain of hangers) which apply equal and opposite distributed reactive forces
to the two substructures.
2.1. Equations of Motion
Kinematics of the beam are assumed to be linear; nonlinearities are accounted for in the description of the aerodynamic forces only. The current configuration, both of the beam and of the cable, is described by the common transverse displacement field
, with
an abscissa along the beam axis and
t the time. The equations ruling the in-plane motion, obtained eliminating the reactive forces
between the cable and the beam equations, are:
where
is the flexural stiffness of the beam,
is the axial stiffness of the cable, and
and
are the beam and cable linear mass density, respectively (and therefore the total bridge linear mass density is
). Here
is the prestress of the cable, assumed constant on
s, and
, with
g the acceleration gravity, is the prestress curvature, also assumed to be constant. The integral term accounts for the increment of tension, which depends on the time only.
In Equation (
1), the non-inertial contribution of the external loads are decomposed in the damping and aerodynamic parts, as:
where the subscripts
refer respectively the beam and cable.
The Equation (
1) is a generalization of the motion equations provided for the first time by Bleich et al. in [
13], and recently re-examined in [
18]. They include damping and aerodynamic loads, not present in the original formulation.
2.2. Damping Forces
A damping continuous model, accounting for both external damping (due to firm air-structure interaction) and internal damping (due to material dissipation), is used. This approach is different from that usually followed in literature, in which an ‘overall’ damping is attributed to the modes of the discrete system on experimental (or of literature) ground.
The contribution of the internal damping is related to two different physical mechanisms (see [
37]): (a) elongation of the longitudinal fibers, both in cable and in beam, and (b) internal friction of the prestressed fibers of the cable. By adopting the Kelvin-Voigt model for mechanism (a), the elastic modulus
must be replaced by
with
an internal viscous damping coefficient, having physical dimensions of a time. By adopting the same rule for mechanisms (b), the pretension
must be replaced by
. In symbols:
Accordingly, the time-differential damping operator is proportional to the elasto-geometric stiffness operator. Such a type of damping is usually referred as proportional, or Raleigh model, in the context of the finite-dimensional systems.
The external damping is modeled by forces proportional to velocities, as:
in which
and
are beam and cable external damping coefficients, from which
is the coefficient of total external damping.
Note that the damping model introduced here is able to distinguish the beam and cable contributions. The four damping parameters can be calibrated to reproduce modal damping of as many modes.
2.3. Aerodynamic Forces
The aerodynamic part of the load is caused by the wind, which blows orthogonally to the plane of bridge, with a steady velocity U. It triggers forces on the beam and on the cable, acting both in-plane and out-of-plane of the structure. Due to the higher out-of-plane stiffness of the beam and to the smaller dimensions of the cable (which, being of circular shape, is exclusively subjected to drag forces), only in-plane forces acting on the beam are taken into account.
The quasi-static theory (see [
29]) is used here to express such forces. The choice is corroborated by the discussion carried out in [
28], where it is stated that: (a) the use of a nonlinear quasi-steady theory is justified in the range of low reduced frequencies, in which long-span bridges are expected to fall; and, (b) the previsions of the quasi-steady model are conservative, a few percent lower than those of a fully unsteady model. In particular, it is known that (see e.g., [
43,
44]), the quasi-steady assumption is valid only if the main frequency of periodic components of fluid forces, associated with vortex-shedding, is well above the dominant structural frequency, i.e.,
, where
is the natural frequency of the crosswind vibration mode of the structure and
, with
the Struhal number and
b the deck width. The occurrence of the frequency condition will be verified in the numerical application. When the inequality is not satisfied, it is needed to resort to a non-steady theory. However, while this latter is well codified in the linear range, few studies concern nonlinear dynamics (e.g., [
44,
45,
46]).
In the framework of the quasi-steady theory, the forces depend on the structural velocity
only (not on the frequency of the oscillations), as result of the aeroelastic interaction. By expressing them in series of powers of the
ratio, by assuming the cross-section is symmetric with respect the horizontal axis, and truncating the series to the fifth order, they read:
Here, is the air mass density, are dimensionless aerodynamic coefficients (depending on the shape of the cylinder cross-section), D is a transverse characteristic length, and ( ) are dimensional coefficients, introduced to simplify the notation.
2.4. Dimensionless Equation of Motion
By introducing the following non-dimensional quantities:
and parameters:
the equations of motion are recast in non-dimensional form:
In Equation (
8), prime and dot denote differentiation with respect to non-dimensional space and time, respectively, and the tilde is dropped for convenience. It is noticed that nonlinearities are only due to aerodynamics.
The dimensionless parameter
is the well-known Irvine-Caughey parameter [
3], accounting for geometric and mechanical properties of the cable. The dimensionless parameter
(also used in [
13,
47,
48] and others), describes the beam-to-cable stiffness ratio. For long bridges, it tends to zero, so that such bridges behave as a cable with an appended mass. However, still in this case, the
parameter causes the occurrence of boundary layers close to the constraints and the point-loads, where the flexural behavior cannot be neglected. In this paper, attention will be devoted to bridges of medium length, in which the role of the beam is not negligible, as it will be discussed soon. A parametric study about the influence of both parameters on the free dynamic of the bridge is carried out in [
18].
6. Numerical Results
The behavior of the long-span suspension bridges is often well-described by the cable only. Here, however, the focus is addressed to cases in which the beam significantly influences the overall mechanics, as it happens when the dimensionless parameter
is of order
, (see, e.g., [
13,
18,
48]). Therefore, a case study is considered: (a) the geometrical and mechanical characteristics are
m,
Nm
,
N,
kg/m and
kg/m; (b) the prestress is evaluated as
, where
m is the cable sag; (c) the damping ratio is taken as
in the antisymmetric mode, and
in the symmetric mode (corresponding to
s,
Ns/m
,
s, and
Ns/m
, whose calibration is shown in
Appendix A). A rectangular boxed cross-section of dimension
m (and so with an aspect ratio 2:1) is assumed for the beam. The aerodynamic dimensionless coefficients in Equation (
5) are drawn from literature (obtained by wind tunnel tests in [
30]), namely
,
,
. The air mass density is
kg/m
.
This example is representative of a class of bridges, whose nondimensional parameters are: , , , , , , , . All numerical values assumed by the parameters are consistent with the ordering performed in the perturbation analysis.
6.1. Linear Stability Analysis
The modal properties of the undamped model were first analyzed. The first natural mode was found to be anti-symmetric (defined by Equation (
11) for
) and the corresponding natural frequency is
rad/s. The second natural mode was found to be symmetric (defined by Equation (
13) for
) and the natural frequency is
rad/s (see
Figure 3).
The critical conditions corresponding to the antisymmetric and symmetric modes are shown in
Table 1. It is observed that the bridge galloped in its first antisymmetric natural mode, in corresponding of which the critical wind velocity assumed the minimum value
m/s.
Since the critical velocities corresponding to the antisymmetric and symmetric modes were sufficiently far from each other, nonlinear interactions between modes, leading to multimodal galloping, were excluded.
It was observed that the quasi-steady theory was valid in the case study considered. Indeed, the natural frequency of the first crosswind vibration mode was
Hz, while the vortex-shedding frequency was
Hz at the critical galloping velocity
m/s, according to a Struhal number
(as indicated by [
54] for a rectangular cross-section with a 2:1 aspect ratio) and a deck width of 2 m. Note that the quasi-steady theory worked mainly due to the narrow width of the cross section.
6.2. Galloping Response
The postcritical behavior relevant to the case study is described by the bifurcation diagram in
Figure 4. There, the (dimensionless) steady amplitude
a of the anti-symmetric mode, experienced by the bridge on the limit-cycle, is plotted
the (dimensionless) wind velocity
u. It is seen that, the hard loss of stability caused amplitudes of about
of the length (i.e., of the order of 2 m), as far the critical wind velocity was passed.
Such a bad mechanical behavior is due to the unfavorable shape of the cross-section ([
23,
30]).
6.3. Validation by Finite-Dimensional Model
The asymptotic solution was validated against numerical results, relevant to the discretized version (Equations (
37) and (
38)) of the nonlinear continuous model (Equation (
8)). The space interval was divided in
subintervals, and numerical integration carried out, via commercial routines, for given initial conditions. Regardless of these latter, after a transient had been exhausted, the motion stabilized either on a limit cycle or comes back to the trivial state. In order to compare the transient motion
to that predicted by the MSM via the amplitude
, an initial deflection shape coincident with the first natural mode of the undamped bridge was chosen.
Four typical time-histories are shown in
Figure 5 for different wind velocities. It was observed that: (a) when
(
Figure 5a) the motion stabilized on the trivial solution
; (b) when
(
Figure 5b) the motion stabilized either: (I) on the trivial solution
, if the initial condition was sufficiently small, or (II) on a limit cycle of amplitude
, if the initial condition was sufficiently large; (c) when
(
Figure 5c) the motion stabilized on a limit cycle of amplitude
. All the time-histories were compared with the numerical integration of the bifurcation Equation (
27). It was seen that this latter captured at the best extent the exact (numerical) solution.
To compare the numerical to the asymptotic limit-cycles, the numerical amplitudes were extracted by the recorded time-histories for different wind velocities, and bullets superimposed to the asymptotic bifurcation diagram (see
Figure 5d), showing an excellent accordance. Note the power of the perturbation methods in providing accurate predictive formulae (as for the galloping velocity, Equation (
29) and for the amplitudes of the limit-cycles, Equation (
30)).
7. Conclusions
The nonlinear aeroelastic behavior of suspension bridges, undergoing dynamic in-plain instability (galloping), has been analyzed. A nonlinear continuous model of bridge has been formulated, made of a visco-elastic beam and a parabolic cable, connected each other by axially rigid suspenders, uniformly distributed. A damping continuous model, accounting for both external damping and internal damping (according to the Kelvin-Voigt rheological model), has been used. The structure is loaded by a uniform wind flow, acting normally to its plane. The nonlinear aeroelastic effects have been evaluated via the quasi-static theory, while structural nonlinearities have been ignored. The nonlinear equations ruling the dynamics of the aeroelastic system have been tackled by the Multiple Scale Method, directly applied to the partial differential equations, to get bifurcation equations for the Hopf bifurcation. The limit-cycle amplitude and its stability have been evaluated as functions of the mean wind velocity. The analytical solutions have been validated against numerical integration of the finite-difference discretized equations.
Numerical results have been obtained for a case study of suspension bridge, whose beam has a rectangular boxed cross-section, with an aspect ratio 2:1. It has been found that the bridge gallops in its first antisymmetric natural mode (having natural frequency
rad/s). An explicit formula has been derived for the critical wind velocity, assuming the value
m/s. The validity of the quasi-steady theory, when related to the vortex-induced vibration phenomenon, has been checked. The postcritical behavior has been investigated, and the bifurcation diagram provided. The well-known (and dangerous) phenomenon of “hard loss of stability” has been found to occur, according which, as far the critical wind velocity is passed, the motion amplitude reaches values equal to
of the bridge length (i.e., of the order of 2 m). Such a bad mechanical behavior is due to the unfavorable shape of the cross-section. All the asymptotic results, relevant to the transient as well to the steady motions, have been fully validated by the numerical analysis. This work is susceptible of farther insights, namely, (a) analysis of the influence of the flexibility of the pylons on the bridge dynamics; (b) study of the interaction between vortex-induced vibrations and galloping; (c) validation of the results by wind tunnel tests. Concerning item (a), elasticity of supports could cause (in principle, and to be investigated in the filed of the real bridges) an inversion between the now dominant antisymmetric and the first symmetric mode, as discussed in [
6]. Such an inversion could also trigger (d) nonlinear interaction between the two modes. Concerning item (b) and (c), study of literature, as described in [
55] could be exploited.