vibrations of a taut string with linear damper

Upload: snailbook

Post on 14-Apr-2018

221 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    1/45

    46

    CHAPTER 3

    Vibrations of a Taut String with Linear Damper

    To suppress problematic stay-cable vibrations like those discussed in the previous

    chapter, viscous dampers are often attached to stay cables near the anchorages to

    supplement the low levels of inherent damping. The potential for widespread application

    of dampers for stay-cable vibration suppression necessitates a thorough understanding of

    the dynamics of the cable-damper system to enable both an effective and economical

    damper design. A simple model of the stay-cable with attached damper is a taut string

    with a linear viscous dashpot. This simplified model neglects the bending stiffness of the

    cable, the sag of the cable due to its static deflection under self-weight, and any

    nonlinearity of the damper, whether unintentional or by design.

    The effect of sag is not considered in the present study, as it has already been

    studied in detail by several investigators, as will be discussed in the following section,

    and it has been observed that for most stay cables, the sag is small enough that its

    influence is not significant. The effect of bending stiffness has been shown to be

    significant for many stay cables (Tabatabai and Mehrabi 2000), and a detailed study of its

    effects is presented in the following chapter. The influence of damper nonlinearity is also

    investigated in subsequent chapters. The simplifications introduced by the taut-string and

    linear-damper approximations allow an analytical solution to be carried quite far,

    providing insights into the complicated and unusual features of the cable-damper system.

    These insights provide a basis for understanding the more complicated solution

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    2/45

    47

    characteristics in the case of nonzero bending stiffness, and the formulation is extended in

    subsequent chapters to develop an approximate solution for nonlinear dampers.

    3.1 Previous Investigations

    Carne (1981) and Kovacs (1982) were among the first to investigate the vibrations

    of a taut cable with an attached damper, both focusing on determination of first-mode

    damping ratios for damper locations near the end of the cable. Carne developed an

    approximate analytical solution to the free-vibration problem, obtaining a transcendental

    equation for the complex eigenvalues and an accurate approximation for the first-mode

    damping ratio as a function of the damper coefficient and location, while Kovacs

    developed approximations for the maximum attainable damping ratio (in agreement with

    Carne) and the corresponding optimal damper coefficient (about 60% in excess of

    Carnes accurate result) from consideration of the frequency response function for a cable

    with attached damper under harmonic excitation.

    Subsequent investigators formulated the free-vibration problem using Galerkins

    method, with the sinusoidal mode shapes of an undamped cable as basis functions, and

    several hundred terms were required for adequate convergence in the solution, creating a

    computational burden (Yoneda and Maeda 1989; Pacheco et al. 1993). Several

    investigators have worked to develop modal damping estimation curves of general

    applicability (Yoneda and Maeda 1989; Pacheco et al. 1993), and Pacheco et al. (1993)

    introduced nondimensional parameters to develop a universal estimation curve of

    normalized modal damping ratio versus normalized damper coefficient, which is useful

    and applicable in many practical design situations. Krenk (2000) developed an exact

    analytical solution of the free-vibration problem for a taut cable and obtained an

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    3/45

    48

    asymptotic approximation for the damping ratios in the first few modes for damper

    locations near the end of the cable; Krenk also developed an efficient iterative method for

    accurate determination of modal damping ratios outside the range of applicability of the

    asymptotic approximation.

    Several investigators have considered the influence of sag on the attainable

    damping ratios. Pacheco et al. (1993) extended their formulation using Galerkins

    method to investigate the influence of small sag on the modal damping ratios, and Xu et

    al. (1997) developed an efficient transfer matrix approach using complex eigenfunctions

    to consider the effect of cable sag and inclination. These studies revealed that moderate

    amounts of sag can significantly reduce the first-mode damping ratio, while the damping

    ratios in the higher modes are virtually unaffected (Pacheco et al. 1993; Xu et al. 1999).

    However, using a database of stay cable properties from real bridges, Tabatabai and

    Mehrabi (2000) found that for most actual stays, the influence of sag is insignificant,

    even in the first mode.

    Previous investigations have focused on vibrations in the first few modes for

    damper locations near the end of the cable, and while practical constraints usually limit

    the damper attachment location, it is important to understand both the range of

    applicability of previous observations and the behavior that may be expected outside of

    this range. Damper performance in the higher modes is of particular interest, as the full-

    scale measurements discussed in the previous chapter revealed that vibrations of

    moderate amplitude can occur over a wide range of cable modes, with vortex-induced

    vibrations observed in quite high modes. In this chapter an analytical formulation of the

    free-vibration problem is used to investigate the dynamics of the cable-damper system in

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    4/45

    49

    higher modes and without restriction on the damper location. The basic problem

    formulation in this chapter follows quite closely the approach recently used by Krenk

    (2000), although it was developed independently as an extension of the transfer matrix

    technique developed by Iwan (Sergev and Iwan, 1981; Iwan and Jones 1984) to calculate

    the natural frequencies and mode shapes of a taut cable with attached springs and masses.

    A similar approach was used by Rayleigh (1877) to consider the vibrations of a taut

    string with an attached mass. Using this formulation, the important role of damper-

    induced frequency shifts in characterizing the system is observed and emphasized.

    Consideration of the nature of these frequency shifts affords additional insight into the

    dynamics of the system, and when the shifts are large, complicated new regimes of

    behavior are observed.

    3.2 Problem Formulation

    The problem under consideration is depicted in Figure 3.1. The damper is

    attached to the cable at an intermediate point, dividing the cable into two segments,

    where 2> 1. Assuming that the tension in the cable is large compared to its weight,

    bending stiffness and internal damping are negligible, and deflections are small, the

    following partial differential equation is satisfied over each segment of cable:

    2 2

    2 2

    ( , ) ( , )k k k k

    k

    y x t y x tm T

    t x

    =

    (3.1)

    whereyk(xk ,t) is the transverse deflection and xk is the coordinate along the cable chord

    axis in the kth segment; m is the mass per unit length, and Tis the tension in the cable.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    5/45

    50

    L

    1

    1x

    2

    c

    TT m2x

    Figure 3.1: Taut Cable with Viscous Damper

    This equation is valid everywhere except at the damper attachment point; at this location

    continuity of displacement and equilibrium of forces must be satisfied. As in Pacheco et

    al. (1993), a nondimensional time to1= is introduced, where mTLo )(1 = . To

    solve (3.1) subject to the boundary, continuity, and equilibrium conditions, distinct

    solutions over the two cable segments are assumed of the form:

    exYxy kkkk )(),( = (3.2)

    where is a dimensionless eigenvalue that is complex in general1. Substituting (3.2) into

    (3.1) yields the following ordinary differential equation:

    )()(

    2

    2

    2

    kk

    k

    kkxY

    Ldx

    xYd

    =

    (3.3)

    Because

    is complex, the solutions to (3.3), which are the mode shapes of the system, are

    also complex. Continuity of displacement at the damper and the boundary conditions of

    zero displacement at the cable ends can be enforced by expressing the solution as

    1The assumed solution may be alternatively expressed as

    jkkkk exYxy )(),( = , where 1=j

    and is complex; this alternative formulation leads to trigonometric rather than hyperbolic expressions for

    the eigenfunctions. Although the two forms are equivalent, when the damping is small, is mostly real,

    and this alternative form leads to more direct expressions for the eigenfunctions. In cases of large damping,

    to be discussed herein, is mostly real, and the form of solution in (3.2) is more direct.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    6/45

    51

    sinh( )

    ( )sinh( )

    kk k

    k

    x LY x

    L

    =

    (3.4)

    where

    is the amplitude at the damper. The equilibrium equation at the damper can be

    written as

    11

    1122

    1

    1

    1

    2

    2

    ===

    =

    x

    xxt

    yc

    x

    y

    x

    yT (3.5)

    where c is the damper coefficient. Differentiating the assumed solution [(3.2) and (3.4)]

    and substituting into (3.5) yields:

    0)coth()coth(21

    =++Tm

    cLL (3.6)

    Eq. (3.6) is equivalent to the frequency equation presented by Krenk (2000). Its roots

    are the eigenvalues of the system, each corresponding to a distinct mode of vibration.

    For specific values of Tmc/ and 1/L, (3.6) can be directly solved numerically to obtain

    the damping ratios in as many modes as desired to an arbitrary degree of accuracy. Each

    eigenvalue can be written explicitly in terms of real and imaginary parts as follows

    ( )21

    1 iio

    ii j

    += (3.7)

    in which i is the damping ratio and i is the modulus of the dimensional eigenvalue,

    which has been referred to as the pseudoundamped natural frequency (Pacheco et al.

    1993). For convenience in subsequent manipulations, new symbols are introduced to

    denote the real and imaginary parts of the ith

    eigenvalue:

    )Re( ii = ; )Im( ii = (3.8a,b)

    It is noted that i is the nondimensional frequency of damped oscillation. The damping

    ratio i can then be computed from i and i:

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    7/45

    52

    1

    2 2

    21i ii

    i i

    = = +

    (3.9)

    Further insight into the solution characteristics can be obtained by expanding (3.6)

    explicitly into real and imaginary parts. Taking the imaginary part of (3.6) and

    expanding using the notation of (3.8a,b) yields the following equation, after some

    simplification:

    )2sin()2cosh()2sin()2cosh()2sin( 1221 =+ LLLL (3.10)

    This equation is independent of Tmc/ , and its solution branches give permissible

    values of

    for a given 1/L, thus revealing the attainable values of modal damping with

    their corresponding oscillation frequencies. Because it corresponds to enforcing equality

    of phase in (3.5), (3.10) will be referred to herein as the phase equation. While solution

    branches to (3.10) are symmetric about 0 = , only negative values of are considered,

    as positive values of are associated with negative values of Tmc/ , which are not of

    practical interest. Taking the real part of (3.6) yields the following equation, after some

    simplification:

    ( )

    ( ) ( )

    ( )

    ( ) ( )0

    2cos2cosh

    2sinh

    2cos2cosh

    2sinh

    22

    2

    11

    1 =+

    + Tm

    c

    LL

    L

    LL

    L

    (3.11)

    For a specific value of with real and imaginary parts and that satisfy (3.10), the

    corresponding value of Tmc/ can be readily computed from (3.11). The expression

    (3.4) for the complex mode shapes can be expanded explicitly in terms of the real and

    imaginary parts using the notation of (3.8a,b), yielding the following expression

    ( ) ( ) ( ) ( )[ ]LxLxjLxLxAxY kkkkkkk sincoshcossinh)( += (3.12)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    8/45

    53

    in which the complex coefficients are given by )sinh( LA kk = . This expression

    will be useful in considering the character of the mode shapes in the following sections,

    in which three special types of solutions will be considered: a non-oscillatory decaying

    solution (

    =0), non-decaying oscillatory solutions (

    =0), and oscillatory solutions

    approaching critical damping ( and 1).

    3.3 Special Limiting Solutions

    3.3.1 The Case of Non-Oscillatory Decay

    Solutions to (3.6) can exist for which is purely real, corresponding to non-

    oscillatory, exponentially decaying solutions in time; in this case

    =0, the phase equation

    (3.10) is trivially satisfied, and (3.11) reduces to:

    ( )

    ( )

    ( )

    ( ) Tm

    c

    L

    L

    L

    L=

    +

    12cosh

    2sinh

    12cosh

    2sinh

    2

    2

    1

    1

    (3.13)

    in which it is emphasized that positive values of Tmc/ are of practical interest, leading

    to negative values of

    . Eq. (3.13) reveals that purely real solutions exist only when

    Tmc/ exceeds a critical value; when , the two terms on the left-hand side of

    (3.13) each approach unity, so that the critical value of c associated with this limit is

    given by:

    Tmccrit 2= (3.14)

    For c2; for c>ccritthe magnitude of decreases monotonically with increasing c. In

    the limit as

    0, it can be shown that Tmc/ . Figure 3.2 shows a plot of vs.

    Tmc/ from (3.13) with varying damper location 1/L .

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    9/45

    54

    0.01

    0.1

    1

    10

    100

    1 10 100

    0.010.05

    0.1

    0.2

    0.3

    0.5

    cTm

    2

    1/ :L

    Figure 3.2: Nondimensional Decay Rate of Non-Oscillatory Solution vs. Nondimensional

    Damper Coefficient with Varying Damper Location

    This case of non-oscillatory decay may be compared to the case of supercritical

    damping for a single-degree-of-freedom (SDOF) oscillator in free vibration. If an

    exponential form of solution, x Ae= , is assumed for the SDOF oscillator, where x is

    the displacement of the oscillator, /k m t= is a nondimensional time, kis the spring

    constant, and m is the mass of the oscillator, then it is observed that there are two

    solutions for the nondimensional decay rate, which can be expressed as

    2

    1,2 12 2

    c c

    km km

    =

    (3.15)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    10/45

    55

    If /(2 ) 1c km , then both values of the decay rate are purely real, and the

    solution is non-oscillatory, a sum of decaying exponentials. The critical value of the

    damper coefficient for the SDOF oscillator is then given by kmccrit 2= , which has a

    remarkably similar form to (3.14) for the taut string. Using the notation of (3.8a,b),

    = because is purely real, and Figure 3.3 shows a plot of the magnitude of the two

    solutions for the decay rate from (3.15),1

    and2

    , as functions of the nondimensional

    damper coefficient /c km . The magnitude of 2 increases with /c km (plotted in

    gray), but the magnitude of1

    decreases with /c km (plotted in black). Because the

    overall decay of the solution is governed by the more slowly decaying exponential, the

    net rate of decay decreases with increasing /c km , following the black curve in Figure

    3.3, which has features similar to those observed for the taut string in Figure 3.2.

    With

    =0 in this case, the expression for the mode shapes (3.12) reduces to a

    hyperbolic sinusoid with a purely real argument:

    ( )LxAxY kkkk sinh)( = (3.16)

    Figure 3.4 illustrates the mode shape corresponding to a purely real eigenvalue,

    plotted for several different values of Tmc/ , with an arrow indicating the trend with

    increasing c. As c and 0 according to (3.13), the linear term in the series

    expansion of )sinh( Lxk becomes dominant, and the mode shape tends to the static

    deflected shape of the cable under a concentrated load.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    11/45

    56

    0.01

    0.1

    1

    10

    100

    1 10 1002

    ckm

    1

    2

    1

    2 2

    c c

    km km

    = +

    2

    2

    12 2

    c c

    km km

    =

    Figure 3.3: Nondimensional Decay Rate of Non-Oscillatory Solutions for SDOFOscillator vs. Nondimensional Damper Coefficient

    increasing /c Tm

    Figure 3.4: Mode Shape Associated with a Purely Real Eigenvalue

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    12/45

    57

    3.3.2 Limiting Cases of Non-Decaying Oscillation

    Solutions to (3.6) can exist for which

    is purely imaginary (

    = 0), corresponding

    to non-decaying oscillation or zero damping. It can be shown that when = 0, the phase

    equation (3.10) reduces to

    0)sin()sin()sin( 21 = LL (3.17)

    which indicates that there are three sets of frequencies associated with = 0,

    corresponding to the zeros of each of the sinusoids in (3.17). One of these sets of

    frequencies is associated with the limit ofc0 and the other two sets of frequencies are

    associated with the limit of c (i.e., a perfectly rigid damper.) Many investigators

    have made reference to these limiting cases in giving a physical explanation for the

    existence of an optimum damper coefficient for a specific mode: in both of these limits,

    physical intuition suggests that the damping ratios must be zero. These limiting cases can

    be observed more clearly by rearranging the eigenvalue equation (3.6) into the following

    form:

    0)sinh()sinh()sinh( 21 =+ LLTm

    c (3.18)

    In the first limiting case, when c0, (3.18) reduces to 0)sinh( = , which can only be

    satisfied if

    is an integer multiple of 1=j . Consequently, must be purely

    imaginary and

    must take on integer values:

    ioi = (3.19)

    These frequencies simply correspond to vibrations in the undamped natural modes of the

    cable.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    13/45

    58

    In the second limiting case, when c

    , (3.18) becomes

    0)sinh()sinh( 21 =LL , which yields two sets of roots, both of which are purely

    imaginary. One set of roots, corresponding to the zeros of )sinh( 2 L , yields the

    following frequencies:

    )( 2Lici = (3.20)

    These frequencies correspond to vibrations of the longer cable segment of length 2, with

    zero displacement at the damper (the subscript c indicates a clamped frequency). The

    second set of roots, corresponding to the zeros of )sinh( 1 L , yields the following

    frequencies:

    )( 1)1(

    Lici = (3.21)

    These frequencies correspond to vibrations of the shorter cable segment of length 1 (the

    superscript indicates vibrations of the cable segment with index k=1), also with zero

    displacement at the damper. As the three sets of frequencies (3.19)-(3.21) represent all

    the solutions to (3.17), there no other frequencies associated with =0. Figure 3.5 shows

    a plot of these three sets of frequencies against1/L ; this plot will prove useful in

    explaining and categorizing the complicated solution characteristics that will be observed

    under large frequency shifts, and its features will be discussed in more detail

    subsequently.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    14/45

    59

    0

    1

    2

    3

    4

    5

    6

    78

    9

    10

    0 0.1 0.2 0.3 0.4 0.5

    L1

    )1(ci

    ci

    oi

    Figure 3.5: Undamped Frequencies and Clamped Frequencies vs. Damper Location

    When =0, the expression for the mode shapes (3.12) reduces to:

    ( )LxjAxY kkkk sin)( = (3.22)

    which indicates that when =0, the mode shapes over each cable segment are sinusoids

    with purely real arguments. Figure 3.6 depicts mode shapes associated with these three

    special cases of non-decaying oscillation; the mode shape associated with o1 is depicted

    in Figure 3.6(a), and the mode shapes associated with c1 and c1(1)

    are depicted in Figure

    3.6(b) and Figure 3.6(c), respectively.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    15/45

    60

    a)

    b)

    0=c

    c)

    c

    c

    Figure 3.6: Limiting Cases of Non-Decaying Oscillation

    3.3.3 Oscillatory Solutions Approaching Critical Damping

    Complex-valued solutions to (3.6) can exist for which as the

    nondimensional frequency

    tends to a constant value. When

    with bounded,it

    is evident from (3.9) that 1, so that while the real part of the eigenvalue increases

    without limit, the damping ratio itself approaches a finite value (

    =1), corresponding to

    critical damping. The frequencies associated with these limiting cases of critical

    damping can be determined from the phase equation (3.14) by taking the limit as ,

    which yields the equation 0)2sin( 1 =L . The frequencies that satisfy this equation

    can be divided into two distinct categories:

    ( )( )1, /2/1 Liicrit = (3.23)

    ( )(1) , 1/crit i i L = (3.24)

    It is noted that the frequencies crit,i(1)

    coincide with the clamped frequencies of the

    shorter cable segment in (3.21). It can be readily shown using (3.11) that as

    ,

    Tmc/ 2, or cccrit, regardless of the value of

    . As

    , the expression for the

    mode shapes (3.12) tends to

    ( ) ( )LxjLxAxY kkkkk expsinh)( (3.25)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    16/45

    61

    Eq. (3.25) indicates that as

    , the magnitude of the mode shape over each cable

    segment takes the same hyperbolic form as (3.16) for the case of non-oscillatory decay.

    3.4 Solution Characteristics for Small Frequency Shifts

    By focusing their attention on vibrations in the first few modes for small values of

    1/L, previous investigators restricted their attention to cases in which the damper-

    induced frequency shifts are small. Consideration of the characteristics of solution

    branches to the phase equation (3.10) in these cases provides additional insight into the

    dynamics of the cable-damper system.

    3.4.1 Asymptotic Approximate Relations

    Figure 3.7 shows the first five branches of solutions to (3.10) for 1/L=0.05

    plotted in the - plane; a distinct solution branch is associated with each mode. The

    frequencies associated with the special cases of non-decaying oscillation discussed above

    are plotted along the horizontal axis; the undamped frequencies, oi, are indicated by

    circles and the clamped frequencies, ci, are indicated by crosses. Each solution branch

    originates at oi with =0, terminates at ci with =0, and takes on nonzero values of for

    intermediate frequencies.

    0 1 2 3 4 50

    0.01

    0.02

    0.03

    0.04

    05.01 =L

    Figure 3.7: Phase Equation Solution Branches (

    1/L = 0.05, First 5 Modes)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    17/45

    62

    The difference between these two limiting frequencies is given by

    ( )21/ioicii == (3.26)

    The similar form of the solution branches for each mode suggests that it may be useful to

    define a new parameter as a measure of the damper-induced frequency shift

    10;

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    18/45

    63

    generated by numerical solution of (3.10). A distinct curve is plotted for each mode, and

    for this damper location, the numerically generated curves corresponding to each of the

    five modes are indeed very nearly identical, agreeing quite well with (3.30). An

    approximate relation between the clamping ratio and Tmc/ can be established by

    substituting the expressions for i (3.28) and i (3.29) into (3.11) and using similar

    asymptotic approximations, yielding the following relation:

    ci

    ci

    1

    12

    (3.31)

    where

    is a nondimensional parameter grouping defined as

    Li

    mL

    c

    o

    1

    1

    (3.32)

    It is noted that )/(/ 1omLcTmc = , and the alternative normalization )/( 1omLc is

    used in (3.32) to facilitate comparison with previous investigations;

    is equivalent to the

    nondimensional parameter grouping plotted on the abscissa in the universal curve of

    Pacheco et al. (1993). Figure 3.8(b) shows a plot of (3.31) along with numerically

    computed curves of

    versus ci in the first five modes for 1/L=0.02, generated from

    (3.11) after numerical solution of (3.10). Again, the numerically computed curves

    corresponding to each of the five modes are nearly identical, agreeing well with the

    approximation of (3.31). The clamping ratio increases monotonically with

    , from ci =

    0 when

    = 0 to approach

    ci = 1 as

    . Substituting the optimal clamping ratio of

    ci=1/2 into (3.31) reveals that the optimal value of modal damping is achieved when =

    2 0.101, as determined by Krenk (2000). Eq. (3.31) can be inverted and solved for

    ci:

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    19/45

    64

    1)(

    )(22

    22

    +

    ci (3.33)

    Substituting this expression into (3.30) yields the following approximate relation:

    1)()/( 22

    2

    1 +

    L

    i

    (3.34)

    This expression was recently derived by Krenk (2000); it is an analytical expression for

    the universal curve of Pacheco et al. (1993) and is a generalization of the first-mode

    approximation obtained by Carne (1981). Figure 3.8(c) shows a plot of (3.34) along with

    curves of i/( 1/L) versus for the first five modes for 1/L=0.02 generated numerically

    from (3.10) and (3.11). It is evident that again the curves collapse very nearly onto a

    single curve in good agreement with the approximation of (3.34).

    It is important to note that the optimal damping ratio can only be achieved in one

    mode for a linear damper. Noting that

    is proportional to mode number, it is evident

    from Figure 3.8(b) that ci increases monotonically with mode number, so that if the

    damper is designed optimally for a particular mode, it will be effectively more rigid in the

    higher modes and more compliant in the lower modes, resulting in suboptimal damping

    ratios in other modes. As discussed in Chapter 5, recent investigations indicate that

    nonlinear dampers may potentially overcome this limitation in performance, allowing the

    optimal damping ratio to be achieved in more than one mode; however, in the case of a

    nonlinear damper, the damping performance is amplitude-dependent, and optimal

    performance can only be achieved in the neighborhood of a particular oscillation

    amplitude.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    20/45

    65

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    L

    i

    1

    ci

    5-1modes

    02.01 =L

    asymptotic

    a)

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    ci

    b)

    asymptotic

    5-1modes

    02.01 =L

    Figure 3.8: (a) Normalized Damping Ratio vs. Clamping Ratio; (b) Normalized Damper

    Coefficient vs. Clamping Ratio; and (c) Normalized Damping Ratio vs. Normalized

    Damper Coefficient.

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    L

    i

    1

    asymptotic

    c)

    5-1modes

    02.01 =L

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    21/45

    66

    3.4.2 Complex Mode Shapes

    Asymptotic approximations can also be obtained for the complex mode shapes

    when the damper-induced frequency shifts are small, affording additional insight into the

    solution characteristics in this case. The complex mode shapes over the longer cable

    segment can be expressed as follows, by setting2A jA= in (3.12):

    ( ) ( ) ( ) ( )2 2 2 2 2 2( ) cosh sin sinh cosY x A x L x L j x L x L = (3.35)

    When is small, ( )cosh 1kx L and ( )sinh k kx L x L . Using these

    approximations, expressing using (3.28), and using the approximation for in (3.29),

    (3.35) can be approximated as:

    2 1 2 22 2( ) sin ( ) (1 ) cos ( )

    ci i ci ci ci i

    x x xY x A i j i i

    L L L L

    + + +

    (3.36)

    In obtaining an approximation for the mode shapes over the shorter cable segment, it is

    noted that when1/L is small, ( )1cos 1x L and ( )1 1sin x L x L . Using

    these approximations, in addition to the approximations used in obtaining (3.36), the

    mode shape over the shorter cable segment can be expressed as

    1 11 1 1( ) (1 ) ( )ci ci ci i

    xY x A i j i

    L L

    + +

    (3.37)

    Eq. (3.37) indicates that the mode shapes are approximately linear on the shorter cable

    segment. The complex constant A1 is determined from the requirement of continuous

    displacement:1 1 2 2( ) ( )Y Y= . Because the mode shapes are known to be approximately

    linear over the shorter segment from (3.37), continuity of displacement can be enforced

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    22/45

    67

    directly by expressing the mode shapes as1 1 1 1 2 2( ) ( / ) ( )Y x x Y = , yielding the following

    expression:

    1 2 1 2 2

    1 11

    ( ) sin ( ) (1 ) cos ( )ci i ci ci ci ix

    Y x A i j i iL L L L

    + + +

    (3.38)

    It will prove useful to shift the phase of the complex mode shapes so that the imaginary

    part has zero amplitude at the damper attachment point. The mode shapes after shifting

    the phase can be expressed as:

    ( ) ( )zj

    k k k k Y x e Y x= (3.39)

    where z is the phase of the complex amplitude at the damper, which is given by

    [ ]

    [ ]2 2

    2 2

    Im ( )tan

    Re ( )z

    Y

    Y =

    (3.40)

    Substituting (3.36) with 2 2x = into (3.40) yields the following expression:

    [ ]1 2

    2

    ( / )( / ) (1 )tan

    tan ( )( / )

    ci ci

    z

    ci i

    i L L

    i L

    +

    (3.41)

    With the substitution2 1/ 1 /L L= , the argument of the tangent function in the

    denominator on the right-hand side of (3.41) can be expanded as follows:

    1( )(1 / ) [ ( 1) ]

    ci i ci ii L i + + (3.42)

    in which the approximation1 1 2

    ( / ) ( / )ii L i has been introduced, and the term

    1( / )i L has been neglected because i and 1( / )L are both assumed to be 1 . It

    is then noted that tan[ ( ( 1) )] tan[ ( 1) ]ci i ci i

    i + = , and because 1i

    , the

    tangent function can be approximated by its argument. Eq. (3.41) then reduces to

    1 2( / )( / ) (1 )

    tan( 1)

    ci ci

    z

    ci i

    i L L

    (3.43)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    23/45

    68

    With the approximations2

    ( / ) 1L and1

    ( / )i

    i L , (3.43) reduces to

    tan1

    ciz

    ci

    (3.44)

    The following approximations can be obtained from (3.44):

    sin z ci (3.45)

    cos 1z ci (3.46)

    Using the identity cos sinzjz z

    e j = with the approximations in (3.45) and (3.46),

    the phase-shifted mode shapes (3.39) can be expressed as

    ( ) ( 1 ) ( )k k ci ci k k Y x j Y x + (3.47)

    Eq. (3.47) can be expanded into real and imaginary parts; the real part of the phase-

    shifted mode shape is given by

    Re ( ) 1 Re[ ( )] Im[ ( )]k k ci k k ci k k

    Y x Y x Y x (3.48)

    and the imaginary part is given by

    Im ( ) Re[ ( )] 1 Im[ ( )]k k ci k k ci k k

    Y x Y x Y x + (3.49)

    Substituting (3.36) into (3.49) yields the following expression:

    22 2

    1 2 2

    Re ( ) 1 sin ( )

    1 cos ( )

    ci ci i

    ci ci ci i

    xY x A i

    L

    x xA i i

    L L L

    +

    +

    (3.50)

    The sine in (3.50) can be expanded as

    2 2 2 2 2sin ( ) sin cos cos sinci i ci i ci ix x x x x

    i i iL L L L L

    + = + (3.51)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    24/45

    69

    Because 1i

    , ( )2cos / 1ci ix L and ( )2 2sin / / ci i ci ix L x L , and

    (3.51) can be approximated by

    2 2 2 2

    sin ( ) sin cosci i ci ix x x x

    i i iL L L L

    + + (3.52)

    The cosine in (3.50) can be expanded as

    2 2 2 2 2cos ( ) cos cos sin sinci i ci i ci ix x x x x

    i i iL L L L L

    + = (3.53)

    And using the same approximations as used in (3.52), (3.53) can be approximated by

    2 2 2 2

    cos ( ) cos sinci i ci ix x x x

    i i iL L L L

    + (3.54)

    Substituting (3.52) and (3.54) into (3.50) and introducing the approximation

    1( / ) ii L allows some cancellation of terms, and an additional term can be neglected

    because it has a factor of1

    ( / )i L , where i and 1( / )L are both assumed to be

    1 . The real part of the phase-shifted mode shape can then be expressed as

    ( )2 2 2Re ( ) 1 sin / ciY x A ix L (3.55)

    Eq. (3.55) is an expression for the undamped mode shapes of the cable, with a scaling

    factor of 1 ci . It was previously shown in (3.37) that the mode shapes are

    approximately linear over the shorter cable segment, and because1/ 1L , the real part

    of the mode shape can be approximated as

    ( )Re ( ) 1 sin / ciY x A ix L (3.56)

    in which x is a coordinate originating at the end of the cable nearer the damper and

    spanning the entire length of the cable. Eq. (3.56) is approximately linear over the

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    25/45

    70

    shorter segment, and is equivalent to (3.55) over the longer segment, but scaled by a

    factor of 1( 1)i .

    A similar asymptotic approximation can be developed for the imaginary part of

    the phase-shifted mode shapes. Substituting (3.36) into (3.49) yields the following

    expression:

    22 2

    1 2 2

    Im ( ) sin ( )

    (1 ) cos ( )

    ci ci i

    ci ci ci i

    xY x A i

    L

    x xA i i

    L L L

    +

    + +

    (3.57)

    The sine in (3.57) can be alternatively expressed as

    [ ]2 2sin ( ) sin ( (1 ) )ci i i ci ix x

    i iL L

    + = + (3.58)

    Eq. (3.58) can then be expanded as

    2 2 2

    2 2

    sin ( ) sin ( ) cos (1 )

    cos ( ) sin (1 )

    ci i i ci i

    i ci i

    x x xi i

    L L L

    x x

    i L L

    + = +

    +

    (3.59)

    Because 1i

    , the following approximations can be used: [ ]2cos (1 ) / 1ci ix L

    and [ ]2 2sin (1 ) / (1 ) / ci i ci ix L x L . Eq. (3.59) can be approximated by

    2 2 2 2sin ( ) sin ( ) (1 ) cos ( )ci i i ci i ix x x x

    i i iL L L L

    + + + (3.60)

    Substituting (3.60) into (3.57) and introducing the approximation 1( / ) ii L allows

    cancellation of terms, and the expression for the imaginary part of the phase-shifted mode

    shapes reduces to

    [ ]2 2 2Im ( ) sin ( ) / ci iY x A i x L + (3.61)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    26/45

    71

    Noting that1 2

    ( / )i

    i = , it can be shown that (3.61) is equivalent to

    2 2 2 2Im ( ) sin( / )ciY x A ix (3.62)

    Eq. (3.62) is an expression for the clamped mode shapes of the cable with zero

    amplitude at the damper location, scaled by a factor of ci . Because (3.62) indicates a

    zero amplitude at the damper location and (3.37) indicates that the mode shapes are

    approximately linear over the shorter cable segment, the imaginary part of the phase-

    shifted mode shape has approximately zero amplitude over the shorter cable segment, and

    can be expressed as

    [ ]1 1 2Im ( ) H( )sin ( ) / ciY x A x i x (3.63)

    in which H( ) is the Heaviside function andx is a coordinate along the entire cable length

    as in (3.56). Eq. (3.63) gives a zero amplitude over the shorter segment and is equivalent

    to (3.62) over the longer segment, but scaled by a factor of 1( 1)i , which is the same

    factor by which the real part in (3.56) was scaled, so that the relative phasing is

    unchanged. Using (3.56) and (3.63) the asymptotic approximation for the complex mode

    shapes can be written as

    ( ) [ ]{ }1 1 2( ) 1 sin / H( )sin ( ) / ci ciY x A ix L j x i x + (3.64)

    The two sinusoids in (3.64) take on their maximum value of unity at approximately the

    same location, so that the amplitude at this location of peak displacementpeak

    x can be

    approximated as

    ( ) 1peak ci ci

    Y x A j + (3.65)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    27/45

    72

    Eq. (3.65) indicates that the magnitude of the peak displacement is asymptotically

    equivalent to the amplitude coefficientA:

    max[ ( )]Y x A (3.66)

    Using (3.64) the amplitude at the damper location1

    ( )Y = can be approximated as

    ( )11 sin / ciA i L (3.67)

    Eq. (3.67) indicates that 0 as 1ci

    , as would be expected.

    To understand the nature of the complex mode shapes it is helpful to consider the

    evolution of the displaced profile with time; the cable displacement can be expressed as a

    function of time using ( , ) Re[ ( ) ]y x Y x e = . Substituting (3.64) into this expression and

    expanding e as (cos sin )e j + then yields the following expression:

    ( ) [ ]{ }1 1 2( , ) 1 sin / cos H( )sin ( ) / sinci ciy x Ae ix L x i x (3.68)

    where = .

    Figure 3.9 shows plots of the first complex mode for a damper location of

    1/ 0.05L = and for three different values of the clamping ratio: 0.05

    ci = , 0.5

    ci = ,

    and 0.95ci

    = ; each column corresponds to a particular value of the clamping ratio. The

    first plot in each column shows the real and imaginary parts of the complex mode shape,

    from (3.56) and (3.63), with A = 1; as the clamping ratio increases, the amplitude of the

    real part (which looks like the undamped mode shape) decreases and the amplitude of the

    imaginary part (which looks like the clamped mode shape) increases. This decrease in

    amplitude of the real part indicates that the amplitude at the damper location is tending to

    zero as the damper becomes more rigid.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    28/45

    73

    Figure 3.9: Complex Mode Shapes for Three Different Clamping Ratios (mode 1,1/ 0.05L = ):

    ci =

    (second column), and 0.95ci

    = (third column). First Row Shows Real (black) and Imaginary (gray)

    Shapes; Subsequent Rows Show Displaced Profiles at Six Successive Instants in the First Half-

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    0 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    /5 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    2 /5 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    / 2 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    3 /4 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    =

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0 .2 0 .4 0 .6 0 .8 1

    Re[ ( )]

    Im [ ( )]

    Y x

    Y x

    0.5ci

    =

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0 .2 0 .

    R e[ ( )]

    Im [ ( )]

    Y x

    Y x

    -1

    0

    1

    0 0.1 0.2 0.3 0.4

    -1

    0

    1

    0 0.1 0.2 0.3 0.4

    -1

    0

    1

    0 0.1 0.2 0.3 0.4

    -1

    0

    1

    0 0.1 0.2 0.3 0.4

    -1

    0

    1

    0 0.1 0.2 0.3 0.4

    -1

    0

    1

    0 0.1 0.2 0.3 0.4

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0 .8 1

    Re[ ( )]

    Im [ ( )]

    Y x

    Y x

    0.95ci

    =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    0 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    /5 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    2 /5 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    /2 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    3 /4 =

    -1

    0

    1

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    =

    /x L /x L

    0.05ci

    = 0.5ci

    =

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    29/45

    74

    Subsequent plots in each column of Figure 3.9 show the displaced profile at six

    successive instants in the first half-cycle of oscillation, from (3.68) with A = 1, and

    without including the decay due to the exponential factor e . Krenk (2000) presented

    similar plots showing the evolution of the displaced profile for an optimal damper in

    mode 1. When 0= , corresponding to the first instant at which the displaced profiles

    are plotted, each profile is equivalent to the real part of the complex mode shape. At this

    instant, the slope is continuous across the damper location and the velocity at this point is

    zero, both of which imply that the damper force is zero at this instant. When / 2 = ,

    corresponding to the fourth instant at which the displaced profiles are plotted, each

    profile is equivalent to the imaginary part of the complex mode shape. At this instant, the

    amplitude at the damper location is zero, while the velocity at the damper location takes

    on its maximum value as do the slope discontinuity at the damper and the force in the

    damper. For the case of the optimal clamping ratio, 0.5ci

    = , the peak displacement

    occurs when 3 / 4 = , at which instant the contributions from the real and imaginary

    parts have the same amplitude. When = , the profile is the same as when 0= , but

    with opposite sign.

    3.5 Solution Characteristics under Large Frequency Shifts

    When the frequency difference i in a given mode becomes large, as is the case in

    higher modes and for larger values of 1/L, the approximate relations (3.30), (3.31), and

    (3.34) become less accurate. Figure 3.10 shows the same plot as in Figure 3.8(c), but for

    a damper located further from the end of the cable, at 1/L=0.05, and a distinct curve has

    been plotted for each of the first eight modes, with an arrow indicating the trend with

    increasing mode number. It is evident that the curves no longer collapse together. The

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    30/45

    75

    solution curves for the first few modes agree quite well with the curves in Figure 3.8(c);

    however, as the mode number increases, yielding larger values ofi, departures become

    significant, and the optimal damping value i,opt/( 1/L) and the corresponding abscissa

    value opt both increase noticeably with mode number. Numerical investigation in the

    first five modes over a range of 1/L revealed that i,opt/( 1/L) increases monotonically

    with i in a given mode, to take on a value of 1.2 - 1.3 as i0.5, and opt also

    increases monotonically with i, taking on a value of about 0.18 as i0.5. When

    i reaches 0.5 in a given mode, new regimes of behavior are observed which are

    discussed in the following sections.

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    L

    i

    1

    increasing mode number

    Figure 3.10: Normalized Damping Ratio versus Normalized Damper Coefficient; ( 1/L =

    0.05, First 8 Modes)

    3.5.1 Special Cases of Vertical Solution Branches

    In investigating and categorizing the complicated characteristics of solution

    branches to the phase equation (3.10) for large frequency shifts, it useful to begin by

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    31/45

    76

    considering some special cases of vertical solution branches, which serve as boundaries

    between different regimes of behavior. These branches extend from

    =0 to

    =

    [from

    =0 to =1 according to (3.13)] at a constant frequency . Inspection of (3.10) reveals

    that if the following conditions are satisfied:

    0)2sin( 1 =L (3.69)

    2sin(2 ) 0L = (3.70)

    sin(2 ) 0 = (3.71)

    then (3.10) is trivially satisfied for all values of , yielding a vertical solution branch.

    The values of that satisfy (3.69) are the two sets of critically damped frequencies

    previously introduced in (3.23) and (3.24). The condition (3.71) is satisfied ifcan be

    expressed as i or as i+1/2, where i is an integer. It is noted that the additional requirement

    (3.70) follows directly from (3.69) and (3.71), and consequently, (3.69)-(3.71) are all

    satisfied in four distinct cases, corresponding to four types of vertical solution branches.

    To aid in interpreting these special cases, Figure 3.11 presents a plot of the different sets

    of nondimensional frequencies, oi (3.19), ci (3.20), ci(1) (3.21) which is equal to

    crit,i(1) (3.24), and

    crit,i (3.23), against 1/L. The combinations of 1/L and

    corresponding to the four distinct types of vertical solution branches are indicated with

    different symbols. These four types are discussed in the following paragraphs, and the

    contexts in which they arise are discussed in subsequent sections.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    32/45

    77

    0

    1

    2

    3

    4

    5

    6

    7

    8

    9

    10

    0 0.1 0.2 0.3 0.4 0.5

    Frequency Curves

    Vertical Branches

    Type 1

    Type 2

    Type 3

    Type 4

    L1

    )1(

    ,

    )1(

    ,

    icritci

    icrit

    ci

    oi

    =

    Figure 3.11: Undamped Frequencies, Clamped Frequencies, and Critically Damped

    Frequencies versus Damper Location with Vertical Solution Branches Indicated

    Along a vertical solution branch, (3.11) can be expressed as

    ( )

    ( )

    ( )

    ( )0

    2cosh

    2sinh

    2cosh

    2sinh

    2

    2

    1

    1 =++

    ++ Tm

    c

    bL

    L

    aL

    L

    (3.72)

    where ( )La 12cos = and ( )Lb 22cos = , and because is constant along

    each branch, a and b are also constant.

    Type 1: crit,n = i. In these cases, indicated with hollow circles in Figure 3.11, the

    nth

    critically damped frequency crit,n (3.23), where n is an integer, coincides with an

    undamped frequency, oi=i. Equating these expressions and solving for 1/L yields

    1/L=(n-1/2)/i, where n

    (i+1)/2, which corresponds to a damper located at the nth

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    33/45

    78

    antinode of mode i. In this case, a=1 and b=1 in (3.72), from which it can be shown that

    for this case, Tmc/ 2 monotonically from zero as from zero.

    Type 2: crit,n = i+1/2. In these cases, indicated with solid squares in Figure 3.11,

    the nth critically damped frequency,

    crit,n (3.23), equals i+1/2. Equating these

    expressions and solving for 1/L yields 1/L=(n-1/2)/(i+1/2), where n

    (i+1)/2. It is also

    noted that in these cases crit,n coincides with a clamped frequency cp (3.20), wherep is

    also an integer. Using these expressions, the intersection frequency can be expressed in

    terms of n and p as (i+1/2)=n+p-1/2, and the corresponding damper location can be

    expressed as 1/L = (n-1/2)/(n+p-1/2), where n

    p. In this case, a=1 and b=1 in (3.72),

    and using the resulting expression it can be shown that Tmc/ takes on a minimum

    value of

    11 2

    1min

    sinh( 2 )(1 )

    cosh( 2 ) 1

    o

    o

    Lc

    LTm

    = +

    +

    (3.73)

    which is less than 2, at a value o which satisfies the equation

    1)2cosh()()2cosh()( 1221 = LLLL . As decreases from o to

    approach zero, Tmc/ increases monotonically to approach infinity, and as increases

    from o to approach infinity, Tmc/ increases monotonically to approach 2. It will be

    observed subsequently that this type of vertical solution branch actually corresponds to

    two solution branches that intersect at the intermediate value = oand diverge; along one

    branch as Tmc/ 2, and along the other branch 0 as Tmc/ .

    Type 3: crit,n(1) = i. In these cases, indicated with solid circles in Figure 3.11, the

    nth

    critically damped frequency associated with the shorter cable segment, crit,n(1)

    (3.24),

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    34/45

    79

    coincides with an undamped frequency, oi=i. Equating these expressions and solving for

    1/L yields 1/L=n/i, where n

    i/2, which corresponds to a damper located at the nth node

    of mode i. In this case, a=1 and b=1 in (3.72), and the resulting expression is the same

    as (3.13) for the case of a purely real eigenvalue, for which it was observed that Tmc/

    increases monotonically from 2 to approach infinity as 0 from .

    Type 4: crit,n(1) = i+1/2. In these cases, indicated with hollow squares in Figure

    3.11, the nth critically damped frequency associated with the smaller cable segment,

    crit,n(1)

    (3.24), equals i+1/2. Equating these expressions and solving for 1/L yields

    1/L=n/(i+1/2), where n

    i/2. In this case, a=1 and b=1 in (3.72), and it can be shown

    that Tmc/ increases monotonically from 2 to approach infinity as 0 from .

    3.5.2 Behavior in a Single Mode

    To further explore the dynamics of the cable-damper system, is it helpful to

    consider the evolution of a single solution branch of the phase equation (3.10) as the

    damper is moved from the end of the cable past the first antinode to the first node of the

    corresponding undamped mode; over this range, three distinct regimes of behavior are

    observed for a given mode. The evolution of the solution branch associated with mode 2

    will be considered, and Figure 3.12 shows a surface plot of

    versus

    and 1/L for mode

    2; the surface was generated by incrementing 1/L from 0 to 0.5 and repeatedly solving

    (3.10). Figure 3.13 is a companion to Figure 3.12 with a separate plot for each of the

    three regimes; each curve in Figure 3.13 is a slice through the surface of Figure 3.12 at a

    specific value of 1/L.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    35/45

    80

    Figure 3.12: Attainable Damping Ratios in Mode 2 vs. Frequency and Damper Location

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    36/45

    81

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    2 2.1 2.2 2.3 2.4 2.5

    05.01 =L

    15.01 =L

    19.01 =L

    199.01 =L

    a)

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    1.5 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3 2.4 2.5

    25.01 =L

    3.01 =L

    33.01 =L

    22.01 =L

    205.01 =L

    b)

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    0.4

    1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2

    34.01 =L

    4.01 =L

    45.01 =L

    49.01 =L

    c)

    Figure 3.13: Phase Equation Solution Branches for Mode 2 in Three Regimes:(a) Regime 1; (b) Regime 2; (c) Regime 3

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    37/45

    82

    Regime 1: Characteristics of regime 1 are exhibited in a given mode i when

    0

    1/L

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    38/45

    83

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3

    2.01 =L

    Figure 3.14: Phase Equation Solution Branches for an Intermediate Case

    Regime 2: Characteristics of regime 2 are exhibited in mode i when

    1/(2i+1)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    39/45

    84

    from the familiar sinusoid to a shape resembling a hyperbolic sinusoid, as indicated by

    (23). Interestingly, when c>ccrit, then no solution exists along a solution branch in regime

    2; this indicates that ifc>ccritvibrations are completely suppressed for modes in regime 2.

    However, other solutions emerge when c>ccrit; it was previously observed that solutions

    with purely real eigenvalues exist only when c>ccrit, and it will be observed subsequently

    that solutions associated with significant vibration of the shorter cable segment also

    emerge when c>ccrit.

    0

    0.5

    1

    0 0.25 0.5 0.75 1

    increasing c

    )(xY

    Lx

    Figure 3.15: Evolution with Damper Coefficient of a Mode Shape in Regime 2(

    1/L = 0.25, mode 2)

    When 1/L=1/(2i1), corresponding to the upper limit of regime 2, an intersection

    of adjacent solution branches occurs, similar to that depicted in Figure 3.14. In this case,

    the (i1)th clamped frequency and the first critically damped frequency

    crit,1 are both

    equal to i1/2, and a vertical solution branch of type 2 occurs at =i1/2.

    Regime 3: Characteristics of regime 3 are exhibited in mode i when 1/(2i

    1)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    40/45

    85

    approaches oi, and the solution branch for the i

    th mode shrinks to a single point at

    =i

    with

    =0. This indicates, as expected, that no damping can be added to a given mode

    when the damper is located at a node of that mode. Interestingly, when the damper is

    located at the node of mode i, a vertical solution branch of type 3 occurs at =i. Vertical

    solution branches of type 3 and 4 are associated with emergent vibrations of the shorter

    cable segment, and the characteristics of these solution branches are considered in the

    following section.

    3.5.3 Modes with Significant Vibration in the Shorter Cable Segment

    Figure 3.16 depicts several solution branches associated with the first clamped

    mode of the shorter cable segment for several different values of 1/L, ranging from 1/3 to

    1/2. Each curve originates at the clamped frequency of the shorter cable segment c1

    (1)

    (3.21) with

    =0 (indicated with squares along the

    -axis in Figure 3.16), and tends to the

    critically damped frequency crit,1(1), (3.24) as 1 ( ) (indicated with diamonds

    along the top of Figure 3.16). Because ci(1)

    = crit,i(1)

    , in each case the solution branches

    originate and terminate at the same frequency. When 1/L=1/3, the damper is located at

    the node of mode 3, and a vertical solution branch of type 3 occurs at

    =3, plotted at the

    right-hand edge of Figure 3.16. A circle and a cross are also plotted at this frequency

    with =0 to indicate that the third undamped frequency, o3, and the second clamped

    frequency, c2, also coincide with crit,1

    (1) and c1

    (1) for this damper location, as can be

    seen in Figure 3.11. Similarly, when 1/L=1/2, the damper is located at the node of mode

    2, and a vertical solution branch of type 3 occurs at

    =3. When 1/L=0.4,

    crit,1(1)=

    c1(1)=2.5, and a vertical solution branch of type 4 occurs at

    =2.5, plotted at the

    center of Figure 3.16. Vertical solution branches of type 3 and 4 have similar

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    41/45

    86

    characteristics, as discussed previously, and for each of these solution branches, the

    origin of the solution branch at

    = 0 is associated with c , and c decreases

    monotonically to ccrit as 1 ( ). Consequently, ifcccrit. Fig. 13

    shows a plot of the magnitude of the mode shape associated with the first clamped mode

    of the shorter cable segment for 1/L=0.4 and for several different values of Tmc/ . For

    slightly supercritical values of c, the damping is very large, and the magnitude of the

    mode shape resembles a hyperbolic sinusoid. As c and 0, the mode shape takes

    on the expected form of a half sinusoid to the left of the damper.

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3

    34.01 =

    L

    37.01 =L

    4.01 =L

    43.01 =L

    47.01 =L

    3

    11 =L

    2

    11 =L

    Figure 3.16: Phase Equation Solution Branches Associated with the First Clamped Modeof the Shorter Cable Segment

    0

    0.5

    1

    0 0.2 0.4 0.6 0.8 1

    increasingc

    Lx

    )(xY

    Figure 3.17: Magnitude of the Mode Shape Associated with the First Clamped Mode of

    the Shorter Cable Segment ( 1/L = 0.4)

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    42/45

    87

    3.5.4 Global Modal Behavior

    The previous discussion of the three regimes of behavior considered damper

    locations only up to the location of the first node, 1/L

    1/i. When 1/L is increased past

    the first node, the branches of solutions to the phase equation (3.10) initially exhibit the

    same characteristics as those observed in regime 1. As the damper is moved still further

    along the cable to approach the second antinode, a vertical solution branch of type 2 is

    encountered at a frequency of i+1/2, and for damper locations past this point,

    characteristics of regime 2 are exhibited. As the damper is moved past the second

    antinode to approach the second node, a vertical solution branch of type 2 is encountered

    at a frequency ofi1/2, and for damper locations past this point, characteristics of regime

    3 are exhibited. The cycle continues past subsequent nodes and antinodes until 1/L=1/2,

    at which point the subsequent behavior is dictated by symmetry. This periodic cycling

    through the three regimes of behavior is conveyed graphically in Fig. 14. The curves of

    limiting frequencies versus 1/L depicted in Figure 3.11 are used as the basis for this

    graphical depiction, and the regions corresponding to each of the three regimes have been

    indicated using different shadings. The shaded areas indicate ranges of and 1/L for

    which solutions to (3.10) exist, and the type of shading indicates the characteristics of

    solution branches in that region. Situations in which solution branches corresponding to

    two adjacent modes intersect at a vertical solution branch of type 2 (as depicted in Figure

    3.14) are indicated by a solid vertical line connecting the undamped frequencies of the

    two adjacent modes.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    43/45

    88

    Figure 3.18: Regime Diagram of Frequency vs. Damper Location

    In the interesting case of 1/L=1/2, the damper is located at an antinode of all the

    odd-numbered modes and at a node of all the even-numbered modes. Vertical solution

    branches of type 1 then occur at the undamped frequencies of each odd-numbered mode,

    and 1 ( ) in these modes as cccrit, while =0 in the even-numbered modes.

    When c>ccrit, vibrations in the odd-numbered modes are completely suppressed, and new

    modes of vibration emerge, including a non-oscillatory decaying mode and a set of

    modes with the same frequencies as the even-numbered modes, associated with vertical

    solution branches of type 3. As c , 0 in this set of emergent modes, and these

    modes take on an appearance similar to that of the undamped even-numbered mode

    shapes, except that they are symmetric about 1/L=1/2, with a discontinuity in slope at the

    damper.

    3.6 Summary

    Free vibrations of a taut cable with attached linear viscous damper have been

    investigated in detail, a problem that is of considerable practical interest in the context of

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    44/45

    89

    stay-cable vibration suppression. An analytical formulation of the complex eigenvalue

    problem has been used to derive an equation for the eigenvalues that is independent of

    the damper coefficient. This phase equation (3.10) reveals the attainable modal

    damping ratios i and corresponding oscillation frequencies for a given damper location

    1/L, affording an improved understanding of the solution characteristics and revealing

    the important role of damper-induced frequency shifts in characterizing the response of

    the system. The evolution of solution branches of (3.10) with varying 1/L reveals three

    distinct regimes of behavior.

    Within the first regime, when the damper-induced frequency shifts are small, an

    asymptotic approximate solution to (3.10) has been developed, relating the

    damping ratio i in each mode to the clamping ratio ci, a normalized measure

    of damper-induced frequency shift. Asymptotic approximations have also been

    obtained relating ci and i to the damper coefficient, and it is observed that these

    approximations grow less accurate when the damper-induced frequency shifts

    become large.

    Characteristics of the second regime are exhibited when the damper is located

    sufficiently near the antinode of a specific mode. For modes in this regime the

    damping ratio increases monotonically with the damper coefficient, approaching

    critical damping as the damper coefficient approaches a critical value

    Tmccrit 2= . When c>ccrit, vibrations are completely suppressed for modes in

    this regime; however, other modes of vibration emerge, including a non-

    oscillatory decaying mode of vibration and a set of modes with significant

    vibration in the shorter cable segment.

  • 7/30/2019 Vibrations of a Taut String With Linear Damper

    45/45

    Characteristics of the solution branches in regime 3 are similar to those in regime

    1 except that the oscillation frequencies are reduced, rather than increased, by the

    damper, and the attainable damping ratios approach zero as the damper

    approaches the node.

    A regime diagram has been presented to indicate the type of behavior that may be

    expected in each of the first 10 modes for any given damper location. The complexities

    of the solution characteristics in different regimes are potentially significant in practical

    applications, and the approach and analysis presented herein facilitate an improved

    understanding of the dynamics of the cable-damper system.