Investigation of Free In-Plane Vibration of Curved Bridges: Navid Heidarzadeh

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Investigation of Free In-plane Vibration of Curved bridges

Navid Heidarzadeh
Former graduate student, Sharif University of Technology, Tehran, Iran.

Shervin Maleki
Associated Professor, Sharif University of Technology, Tehran, Iran.

ABSTRACT:

In this paper the free in-plane vibration of curved bridges is investigated. The exact
closed form solution for free vibration of arches, proposed by Tufekci and Arpachi
(1996), which takes into account the bending effects, shear effects, axial tension,
translational and rotational inertia is adopted for two possible cases of boundary
conditions of curved bridges. The results obtained by this method are compared to those
obtained by the finite element method (FEM) using shell elements. The effect of
geometry, including span length and opening angle, on the natural frequencies and mode
shapes is parametrically studied. Results show that the difference between the natural
frequencies obtained by closed form solution and those obtained by FEM is less than 3
percent. Results also reveal that the first mode is more sensitive to boundary conditions
than other modes. Increasing the central angle subtended by the arch in the case of
constant chord length, decreases the natural frequency of the first mode, but increases the
frequency of the second mode.

Keywords: Curved bridges, Free in-plane vibration, Closed form solution, Finite element
method.

1 INTRODUCTION

As a result of complicated geometry and limited rights of way, horizontally curved bridges are
becoming the norm of highway interchanges and urban expressways. This type of superstructure has
gained popularity since the early 1960s because it addresses the needs of transportation engineering
(Chen and Duan 1999).
The out-of-plane vibration of curved bridges due to traffic movement is of strong engineering interest
and has received considerable attention in recent years. A good literature survey about the out-of-plane
vibration of curved bridges has been presented by K. Senneh et al, 2001. In contrast, the in-plane
vibration of curved bridges has not adequately been studied yet. The in-plane vibration of curved
bridges provides a good understanding of the seismic behavior and responses of such bridges. The
natural frequencies and mode shapes can directly be used for calculating the seismic responses using
an acceleration response spectrum.
Curved bridges can have asymmetrical responses similar to those of skewed bridges (Chen and Duan
1999). Coupling effects in the horizontal directions due to curved alignment of the bridge, makes the
in-plane behavior of curved bridges very complicated. It seems that the irregularities in the dynamic
behavior of curved bridges can be highly attributed to curve alignment of the deck rather than the
substructure.
For in-plane vibration analysis, classical beam theory is usually used to approximate the actual deck
behavior (Abdel-Salam et al. 1988, Singh 1996 and Sextos et al. 2004). It is well-known that for
beams having large cross-sectional dimensions in comparison to their length, and for beams in which
high frequency modes of vibration are critical, the elementary Euler-Bernouli beam, which is derived
based on the assumption that the deflection of beams are due to bending only, may yield unsatisfactory

Paper Number 004


results. In contrast the Timoshenko beam theory which takes into account the rotary inertia, shear
effect and axial extension, gives a better approximation to the actual beam behavior (Tufekci and
Arpachi 1998).
In recent years many finite element formulations have been developed for curved beam elements with
lumped masses (Wolf 1971, Palaninathan et al. 1985, Lin 1998, Patel et al. 1999, Raveendranath et al.
2000, Przemyslaw et al. 2001 and Wu et al. 2003). However, a lumped mass idealization, although
applicable, is not the natural approach for certain types of structures such as curved bridges (Chopra
2001). Tufekci and Arpachi (1998), proposed an exact solution to the governing equation of free
vibration of circular curved beams which accounts for bending effect, shear effect, axial extension,
translational and rotational inertia and the distributed mass. A similar procedure does exist for beams
with variable cross section (Tong et al. 1998).
In this paper, to identify the effects of deck curvature on characteristics of free in-plane vibration of
curved bridges, the simple case of a single span curved bridge is considered. In order to achieve exact
results, the mass has been considered distributed over the length of the curve. The exact closed form
solution for free vibration of arches, proposed by Tufekci and Arpachi (1998), which takes into
account the bending effect, shear effect, axial extension, translational and rotational inertia is adapted
and solved using two possible cases of boundary conditions for curved bridges.

2 BRIDGE GEOMETRY

Figure 1 shows typical cross sections of curved bridges. In this study, as the cross-sectional
dimensions of the deck are usually considerable in comparison to their length, the Timoshenko beam
theory is used to approximate the actual deck behavior.
At the bridge supports, the elastomeric bearings are assumed to be present. As the shear stiffness of
elastomeric bearings is insignificant, concrete shear keys are used to prevent large displacements and
falling of the superstructure due to earthquake induced motions.
In the rest state with no horizontal forces being applied to the superstructure, shear keys are not in
contact with the elastomeric bearings and there is a small gap between the two. These gaps prevents
the deck and shear keys from engaging as a result of unexpected loads during construction or service
loads. Two possible positions of shear keys are shown in Figure 2. In the case of Figure 2a, the shear
keys only prevent transverse displacements. Longitudinal displacement and rotation around the
vertical axis are unrestrained. This is a common detail at expansion joints, which are normally
provided for thermal effects. The boundary condition, corresponding to these bearings, is considered
as a roller in the analytical model.
In the case of Figure 2b, both the longitudinal and transverse displacements are restrained, but the
rotational behavior of structure is complicated due to the gap. If the gap width and the distance
between the centerline of the two nearby bearings are expressed in terms of x and y, respectively (Fig.
2b), the maximum rotation will be equal to 2x/y. For practical range of x, between 1 to 2 cm, and y,
between 4 to 8 m, the maximum rotation is in the range of 0.0025 radians to 0.01 radians. Also the
LRFD Bridge Design Specification (AASHTO 1998) recommends providing a safety margin against
applicable unfactored loads and uncertainties. It suggests the allowable rotation to be taken as 0.005
radians, unless an approved quality control plan justifies a smaller value.

b
tf

h
tw
(a) (b) (c)
Figure 1: Typical cross section of curved bridges, (a) Steel I girder, (b) Steel box girder, (c) Concrete box girder

2
dx

(a) (b)
Figure 2: Two possible positioning of shear keys (a) transverse direction only (b) both transverse and
longitudinal directions

Therefore, if the rotation of the superstructure at supports is more than the mentioned range, the
boundary condition at supports should be chosen as clamped, otherwise, a hinged condition may be
assumed. As the rotation due to free vibration is a variable quantity over time, the boundary conditions
must be frequently changed in the analytical model. Considering the difficulties in modeling time
varying boundary conditions, which needs the rigid body motion and contact mechanics to be taken
into account; two cases of boundary condition (roller-hinged and roller-clamped) are considered in this
paper. It is anticipated that the actual behavior of curved bridges is somewhat between these two cases.
It seems that this assumption is satisfactory for most of the engineering practice.

3 METHODOLOGY

As mentioned in the introduction, an exact closed form solution presented by Tufekci and Arpachi
(1998) is used to calculate circular natural frequencies and mode shapes. The theory behind this
method is explained here briefly.
A schematic of general circular arch with an arbitrary geometry, distributed mass and boundary
conditions subjected to externally distributed forces is shown in Figure 3a. In this figure, is the
radius of curvature, is the central angle subtended by the arch, P is the distributed tangential force, Q
is the distributed transverse force and T is the distributed moment. Note that, the tangential and radial
axes of the local coordinate system, which is rotated degrees from the centerline of the curve, are
shown as and , respectively. A free body diagram of a differential length of the curve and the acting
forces is shown in Figure 3b, where N is the internal axial force, V is the internal shear and M is the
internal moment. The dynamic equilibrium of this differential element can be explained by three
equilibrium equations as indicated below:

(a) (b)
Figure 3: (a) A schematic of general circular arch, (b) Free body diagram of a differential length of the
curve

3
- Dynamic equilibrium of internal forces, external forces and inertial forces in the tangential direction:

dφ dφ 2
d (1)
dN ⋅ cos − 2⋅ V⋅ sin + ρ ⋅ dφ⋅ P ρ ⋅ µ ⋅ dφ⋅ w
2 2 2
dt

- Dynamic equilibrium of internal forces, external forces and inertial forces in the transverse direction:

dφ dφ dφ 2
d (2)
2⋅ N⋅ sin + dN ⋅ sin + dV ⋅ cos + Q⋅ ρ ⋅ dφ ρ ⋅ µ ⋅ dφ⋅ u
2 2 2 2
dt

- Dynamic equilibrium of internal moments, external moments and inertial forces about the vertical
axis:

I 2
d (3)
dM + ρ ⋅ dφ⋅ V + ρ ⋅ dφ⋅ T ρ ⋅ µ⋅ ⋅ dφ⋅ Ω
A 2
dt

In these equations, = mass per length of the curve; u = tangential displacement; w = transverse
displacement; = angle of rotation about the vertical axis; I = moment of inertia with respect to
vertical axis and A = cross-sectional. Neglecting the second order terms in Equations 1-3 and
considering the external forces as inertial forces for free vibration gives:
d 2 2 I 2
N V − ρ ⋅ µ⋅ω ⋅ w ; d d
V −N − ρ ⋅ µ ⋅ ω ⋅ u ; M −ρ ⋅ V − ρ ⋅ µ ⋅ ⋅ω ⋅Ω (4)
dφ dφ dφ A

The in-plane behavior of elastic curved beams, with axial and shear deformations taken into account,
can be represented by three equations below:
d ρ ρ ⋅κ ρ
w u+ ⋅N ; d ; d (5)
E⋅ A u −w + Ω + ⋅V Ω ⋅M
dφ dφ G⋅ A dφ E⋅ I

Where E = modulus of elasticity; G = shear modulus and = shear factor. Equations 4 and 5 can be
written in matrix form as:

y( φ) C⋅ y( φ)
d
(6)

Where:
ρ
0 1 0 0 0
E⋅ A
w ρ ⋅κ
−1 0 ρ 0 0
u G⋅ A

Ω ρ
y ; 0 0 0 0 0 (7)
C E⋅ I
M
N 2 I
0 0 −ρ ⋅ µ ⋅ ω ⋅ 0 0 −ρ
A
V
2
−ρ ⋅ µ ⋅ ω 0 0 0 0 1
2
0 −ρ ⋅ µ ⋅ ω 0 0 −1 0
An exact solution exists for the particular case of a constant coefficient matrix in the form of:
φ⋅ C
y( φ) e ⋅ y( 0) (8)

4
In Equation 8, e .C is a matrix exponential term and y(0) is the initial value vector at =0. e .C can be
expressed exactly using Caylay-Hamilton theory. Nowadays powerful mathematical software are
available which have exponential matrix functions, such as MATLAB (Mathworks Inc. 2002). This
software uses the Potzer method for calculating the exponential matrix term with double precision.
The y(0) can be obtained according to the boundary conditions. Equation 8 at both ends of the curve
suggests that:
−θ θ
−θ 2 ; θ 2 (9)
y e ⋅ y( 0) y e ⋅ y( 0)
2 2

For three possible boundary conditions at each support of a curved bridge, i.e., roller, hinged and
clamped, three components of the solution vector, y, is equal to zero (roller: u=0; M=0; N=0, hinged:
w=0; u=0; M=0, clamped: w=0; u=0; =0).
Equation 9 actually contains 12 linear equations in which the 6 components of the initial value vector,
the 3 components of the solution vector at =- /2 and the 3 components of the solution vector at = /2
are redundant. Choosing the 6 equations corresponding to zero value components of solution vectors
at =- /2 and = /2, gives 6 equations in terms of 6 components of the initial value vector:
H⋅ y( 0) O (10)

Where H is the 6×6 coefficient matrix and O is the zero vector. For non-trivial solution, setting the
determinant of H to zero will give the circular natural frequencies. The initial value vector, y(0), for
each mode can also be calculated from Equation 10. Mode shapes are specified by substituting the
normalized initial values into Equation 8.

4 VALIDATION OF SIMPLIFIED MODEL

The results of the closed form solution for a typical box girder curved bridge of 40 m span with an
opening angle of 60 degrees (case A8 in Table 1) were verified against the finite element model using
the program SAP2000 (Computers and Structures, Inc. 2005). Shell elements were used to model the
deck. Figure 1c shows the cross-section of the deck. The mathematical program, MATLAB
(Mathworks Inc. 2002) was used for mathematical calculations of the closed form solution.
Figure 4 shows the first five circular natural frequencies and mode shapes obtained by the two
methods. The difference between the circular natural frequencies is less than 3 percent. Mode shapes
are also in good agreement with the FEM model.

5 PARAMETRIC STUDY

In this section, the effect of geometrical parameters in the free vibration of curved bridges is
investigated. The curved box girder bridge with spans of 20, 40 and 60 m are considered with opening
angles of 20, 40 and 60 degrees. Boundary conditions were hinged-roller (A) and clamped-roller (B).
Figure 1c shows the cross section of the deck. Geometrical properties and section properties are given
in Table 1. The circular natural frequencies, considering the hinged-roller and clamped-roller
boundary conditions are given in Tables 2 and 3, respectively. The mode shapes of A4-A6 and B4-B6
cases are shown in Figure 5 and Figure 6, respectively.
Generally, results indicate that single span curved bridges with common dimensions and boundary
conditions can be categorized as high frequency structures. The lowest circular natural frequency
through the 18 cases studied, is 12.6 rad/s.

5
Closed Form Solution Finite Element Model

Mode 1, = 39.3 rad/s Mode 1, =38.2 rad/s

Mode 2, = 105.5 rad/s Mode 2, = 105.6 rad/s

Mode 3, = 135.0 rad/s Mode 3, = 0.0470 rad/s

Mode 4, = 207.8 rad/s Mode 4, = 220.4 rad/s

Mode 5, = 296.7 rad/s Mode 5, = 305.0 rad/s


(a) (b)

Figure 4: The first five circular natural frequencies and mode shapes obtained by (a) the proposed theory and
(b) the FEM model, heavier line is the undeformed shape of the arch centreline

Table 1: Geometrical and section properties


Case (degree) L(m) R(m) h(m) b tf tw A(m2) Av(m2) I(m4) M(ton)
A1, B1 20 20 57.6 1.00 8.2 0.2 0.2 4.28 3.2 24.3 13.8 1.5
A2, B2 20 40 115.2 1.50 8.2 0.2 0.2 5.18 3.2 30.3 16.0 1.5
A3, B3 20 60 172.8 2.00 8.2 0.2 0.2 6.08 3.2 36.3 18.1 1.5
A4, B4 40 20 29.2 1.00 8.2 0.2 0.2 4.28 3.2 24.3 13.8 1.5
A5, B5 40 40 58.5 1.50 8.2 0.2 0.2 5.18 3.2 30.3 16.0 1.5
A6, B6 40 60 87.7 2.00 8.2 0.2 0.2 6.08 3.2 36.3 18.1 1.5
A7, B7 60 20 20.0 1.00 8.2 0.2 0.2 4.28 3.2 24.3 13.8 1.5
A8, B8 60 40 40.0 1.50 8.2 0.2 0.2 5.18 3.2 30.3 16.0 1.5
A9, B9 60 60 60.0 2.00 8.2 0.2 0.2 6.08 3.2 36.3 18.1 1.5

6
Table 2: Circular natural frequencies considering the hinged-roller boundary condition
A1 A2 A3 A4 A5 A6 A7 A8 A9
Mode 1: 116.0 36.1 17.3 101.7 31.9 15.3 83.8 26.3 12.6
Mode 2: 219.6 110.0 61.3 227.5 107.1 58.0 232.6 100.6 53.4
Mode 3: 325.3 121.5 76.4 316.7 124.3 80.0 305.8 129.4 84.1
Mode 4: 526.2 215.4 118.2 519.7 209.6 115.1 497.3 200.4 110.2
Mode 5: 533.9 313.1 178.6 527.2 305.3 174.2 528.7 293.1 167.0
Mode 6: 644.9 330.6 223.8 637.5 328.7 221.8 625.5 324.7 217.9
Mode 7: 698.0 410.9 240.2 698.2 402.6 235.3 694.8 389.0 227.8
Mode 8: 740.4 481.7 301.3 725.9 482.0 295.3 704.6 481.3 285.3
Mode 9: 943.0 507.2 362.1 924.8 497.4 355.0 895.0 482.4 343.5
Mode 10: 1028.9 535.8 372.6 1015.0 533.7 368.2 992.9 527.6 360.6

Table 3: Circular natural frequencies considering the clamped-roller boundary condition


B1 B2 B3 B4 B5 B6 B7 B8 B9
Mode 1: 140.5 49.8 25.1 126.4 45.3 22.9 109.3 39.3 19.9
Mode 2: 219.9 110.6 68.7 227.8 109.7 64.7 232.6 105.5 59.9
Mode 3: 334.4 132.0 77.4 326.8 132.7 81.2 316.8 135.0 84.8
Mode 4: 533.5 221.8 124.7 520.5 216.4 121.8 499.7 207.8 117.3
Mode 5: 578.1 316.3 182.8 577.6 308.6 178.5 577.1 296.7 171.5
Mode 6: 645.8 330.7 223.8 640.7 328.8 221.9 631.0 324.8 218.1
Mode 7: 739.2 412.2 242.8 724.4 404.1 238.0 700.7 390.7 230.5
Mode 8: 853.7 495.5 302.9 847.6 494.0 296.9 837.0 481.5 287.1
Mode 9: 944.9 508.0 362.9 927.4 499.7 355.9 899.3 496.3 344.5
Mode 10: 1074.5 549.6 372.6 1061.0 542.9 368.2 1038.1 531.5 360.6
For the clamped-roller boundary condition, the natural frequency of the first mode is 40 percent
greater than the hinged-roller boundary condition. The difference is less than 4 percent for other
modes.
The frequencies of the first three modes decrease up to 1/3 and 1/2, when the span length increases
from 20 m to 40 m and from 40 m to 60 m, respectively. The dynamic characteristics show
complicated trend when the opening angle changes. For example, the frequency of the first mode
decreases up to 16 percent when the opening angle increases from 20 degrees to 40 degrees and from
40 degrees to 60 degrees. In contrast the frequency of second mode increases up to 5 percent.
Mode shapes are approximately similar for each mode in all the cases of A1-A9, and B1-B9. This
indicates that mode shapes are not sensitive to geometrical parameters. Mode shapes in the case of
hinged-roller boundary condition are not very different from those with clamped-roller boundary
condition. In all cases, the first mode shape activates the tangential displacement at the roller end of
the bridge. The second mode shape has an inflection point near the centerline of the curved bridge.
The third mode also has an inflection point near the centerline. Modes 4 and 5 have three inflection
points at various locations along the span.

7
-10 -8 -6 -4 -2 0 2 4 6 8 10 12 -20 -15 -10 -5 0 5 10 15 20 25

x(m) x(m)

(a) (b)

-30 -25 -20 -15 -10 -5 0 5 10 15 20 25 30 35

x(m)

(c)
Base Mode 1 Mode 2 Mode 3 Mode 4 Mode 5

Figure 5: (a) case A4, (b) case A5, (c) case A6

-10 -8 -6 -4 -2 0 2 4 6 8 10 12 -20 -15 -10 -5 0 5 10 15 20 25

x(m) x(m)

(a) (b)

-30 -25 -20 -15 -10 -5 0 5 10 15 20 25 30 35


x(m)

(c)
Base Mode 1 Mode 2 Mode 3 Mode 4 Mode 5

Figure 6: (a) case B4, (b) case B5, (c) case B6

6 CONCLUSION

This paper investigated the effect of deck curvature on characteristics of free in-plane vibration of
single-span curved bridges. The exact closed form solution for free vibration of arches, proposed by
Tufekci and Arpachi (1996) is adopted for two possible cases of boundary conditions of curved
bridges. The Timoshenko theory which takes into account the bending effect, shear effect, axial
extension, translational and rotary inertia was used to approximate the superstructure’s behavior. The
natural frequency and mode shapes were obtained in exact closed form and compared with the finite
element method. Further, the effect of geometrical parameters on the free vibration of curved bridges
was investigated in a parametric study. The significant results are as follows:

8
• The difference between the circular natural frequencies obtained by closed form solution with
those obtained by FEM is less than 3 percent even in the range of 300 rad/s and higher mode
shapes.
• In the case of clamped-roller boundary condition, the natural frequency of the first mode is 40
percent greater than the hinged-roller boundary condition. The difference is less than 4 percent
for other modes.
• Increasing the central angle of the arc and keeping a constant chord length, decreases the natural
frequency of the first mode, but increases the frequency of the second mode.
• Mode shapes are not very sensitive to geometrical parameters and boundary conditions. Mode
shapes in the case of hinged-roller boundary condition are similar to those with clamped-roller
boundary condition.

REFERENCES:

Abdel-Salam M. N. and Heins C. P., ‘Seismic response of curved steel box girder bridges’. Journal of structural
engineering, ASCE. Vol. 114, No. 12 (1988)
Chen W. F. and Duan L., ‘Bridge Engineering Handbook’, CRC press, (1999)
Chopra A. K., ‘Dynamics of structures, theory and application to earthquake engineering’, Prentice Hall, 2001
Lin S. M., ‘Exact solution for extensible circular curved Timoshenko beams with nonhemogeneous elastic
boundary condition’, Acta Mechanica, Vol. 130, pp. 67-79 (1998)
Litewka P. and Rakowski J., ‘Free vibration of shear-flexible and comperessible arches by FEM’, Internation
Journal of Numerical Methods in Engineering, Vol. 52, pp. 273-286 (2001)
LRFD Bridge Design Specification, 2nd Ed. American Association of State Highway and Transportation
Officials, AASHTO, Washington, D.C. (1998).
MATLAB, version 7.1, Mathworks Inc.
Palaninathan R., Chandrasekharan P. S., ‘Curved beam stiffness matrix formulation’, Computers and Structures,
Vol. 21, No. 4, pp. 663-669 (1985)
Patel B. P., Ganapath M. and Saravanan J., ‘Shear flexible field-consistent curved spline beam element for
vibration analysis’, Internation Journal of Numerical Methods in Engineering, Vol. 46, pp. 387-407 (1999)
Raveendranath P., Singh G., Pradhan B., ‘Free vibration of arches using a curved beam element based on
coupled polynomial displacemebt field’, Computers and Structures, Vol. 78, pp. 583-590 (2000)
SAP2000, version 9.03 (2005). “Integrated structural analysis and design software.” Computers and Structures,
Inc., Berkeley, CA.
Senneh K. M., Kennedy J. B., ‘State-of-the-art in design of curved box girder bridges’, Journal of Bridge
Engineering/may/june. pp 159-167 (2001)
Sextos A., Kappos A. J, Mergos P., ‘Effect of soil-structure interaction and spatial variability of ground motion
on irregular bridges: the case of Kristallpigi bridge’, 13th world conference on Earthquake Engineering,
Vancouver, B.C., Canada, 2004
Singh R., ‘Seismic response analysis of curved bridge’, Journal of Individual Studies by Participants in the
International Institute of Seismology and Earthquake Engineering’, Tokyo, Japan, 1996
Tong X., Mrad N., Tabarrok B., ‘In-plane vibration of circular arches with variable cross-sections’, Journal of
Sound and Vibration, Vol. 212(1), pp. 121-140 (1998)
Tufekci E. and Arpachi A., ‘Exact solution of in-plane vibration of circular arches with account taken of axial
extension, transverse shear and rotatory inertia effects’, Journal of Sound and Vibration, Vol. 209(5), pp. 845-
856 (1998)
Wolf J. A., ‘Natural frequencies of circular arches’, Journal of Structural Division, Proceedings of the American
Society of Civil Engineers, Sep. 1971, pp. 2337-2350
Wu J. S., Chiang L. K., ‘Free vibration analysis of arches using curved beam elements’, Internation Journal of
Numerical Methods in Engineering, Vol. 58, pp. 1907-1936 (2003)

9
10

You might also like