Sie sind auf Seite 1von 10

IEEE Transactions on Power Apparatus and Systems, Vol. PAS-104, No.

11, November 1985


AEOLIAN VIBRATION OF SINGLE CONDUCrORS AND ITS CONTROL
0. Nigol, Senior Member R.C. Heics, Member Ontario Hydro Toronto, Canada
H.J. Houston, Member Hamilton, Canada

3245

SLACAN

of

ABSTRACT - This paper describes the development a mechanical-impedance model of a vibrating single-conductor-damper system and its experimental verification. Measurements are made by the forcevelocity method and compared with the corresponding model-generated data. The effects of each of the damper components is demonstrated, and the optimal design of aeolian vibration dampers for an ideal viscous damping medium is established. The optimum locations of dampers are also discussed. INTRODUCTION
In examining the design limitations imposed by the vibration of overhead conductors, it is found that more efficient aeolian vibration dampers would permit the use of higher conductor tensions. This would allow the use of longer spans or shorter towers. It is apparent that an increase in the every-day design tensions from the present 20 per cent rated tensile strength (RTS) of conductors to 30 or 40 per cent RTS could result in large economic benefits.
Many investigators have studied the application of analytical models to the damping of conductor vibration. Claren and Diana [5] used the method of principal modes, but most others have used methods based on impedance models. Tompkins, Merrill and Jones [1) developed electrical-mechanical impedance analogy models to study damper-conductor interactions; Rawlins [2] expressed these models in purely mechanical terms; and Slethei Huse [3] used electrical-mechanical impedance models in their analysis of vibration decay tests; mechanical impedance representations were the basis for Ervik's [6] analysis using the Laplace transform.

The purpose of this paper is to describe the development of a general method for analyzing the combined response of various single conductor-damper arrangements, present a summary of its experimental verification, and to use the models to demonstrate optimum damping concepts. The verification of the general method and the Canadian the by supported resultant models is and Development Research Association Electrical (CEA-RD) contract "Development of Accurate Analytical M4odels for Vibrating Conductor-Damper Systems". To provide the necessary data and to develop an improved method for evaluating vibration dampers, CEA-RD is also the following related under research supporting contracts: "Internal Damping of ACSR, Alumoweld and Steel Conductors"; "Methods of Evaluating Vibration Performance of Dampers, Spacers, Spacer-Dampers and of Aerodynamic Other Conductor Hardware"; "Study Excitation Forces in a Long Conductor Span"; and "Fatigue Life of Aluminum Strands in All Aluminum and ACSR Conductors". This work is expected to be completed in 1984.

VIBRATING CONDUCTOR-DAMPER MODEL


form a clear understanding of conductor To vibration problems it is essential to develop accurate conductor-damper vibrating for analytical models Such models are developed in the Appendix. systems. When the equations are applied to a single conductor span held rigidly at both ends and excited by a sinusoidal force F at the center of the span, the lateral velocity of the conductor at Q/ is given by the following equation

The model used by Tompkins et al, and Rawlins was an open-ended representation of the conductor span with energy moving in only one direction, from the shaker end of the span, to the damper and terminal. The model developed here is an extension of this work. Both ends of the span are represented, and energy flows in both directions from an excitation source located anywhere This method allows simulation of the in the span. behaviour of spans containing more than one damper, and may be also applied to the modelling of bundleconductors with spacer-dampers. A two-conductor-bundle rnodel with up to two spacer-dampers is presently being experimentally verified for motion in one plane.

V2 )=2

F
0

tanhy9/2

F 2Z

sinh&Q + jsin$z coshaR + cos RZ

(A)

and antiresonances at

The expanded form of the above equation shows that the lateral velocity of a conductor goes through a series of harmonically related resonances at cos8Z=-1

cos8k=1.

From the force requirements and the availability of these forces from the wind, it is apparent that only occur at resonant can vibration conductor frequencies. At these frequencies the lateral velocity, Crom equation (A), is

A paper recommended and approved 85 SMl 404-9 by the IEEE Transmnission and Distribution Committee of the IEEE Power Engineering Society for presentation at the IEEE/PES 1985 Summner Meeting, Vancouver, Manuscript subB.C., Canada, July 14 - 19, 1985. Tnitted January 31, 1983; made available for printing April 22, 1985.

V(Z/

,o)
=

F
o

2F 0
c

(B)

The above equation shows that the velocity is directly proportional to the excitation force and inversely proportional to the total internal damping of the conductor span. This equation will also apply to conductors excited by the wind when an equivalent force is applied to the aerodynamic integrated

0018-9510/85/1100-3245$01.00O1985 IEEE

3246
conductor instead of the concentrated force Fo. Assuming that the aerodynamic force Fa per unit length of conductor is constant throughout the span, the expression for the velocity is independent of the span length.

The impedance of the simple viscous damper shown in Figure 3 as viewed from the conductor has the following form

j0M

(D-

jK
K

(E)

V(Q

/2'

a))=

2Fa Q

2Fa
D

D + j (WM -

)
conductor

(C)

The characteristics of aerodynamic force Fa in terms of wind speed, direction, frequency and conductor determined from are diameter currently being measurements of natural wind excitation on a full-sized This data is span under a separate CEA-RD study. of conductor expected to improve the prediction response compared to estimates made from wind tunnel tests on short conductor samples. When dampers are present in the conductor span, their characteristics may be modelled by a complex mechanical impedance which is the ratio of applied force and the resultant velocity at the point of attachment to the conductor. From equation (A) it can be seen that the conductor span can also be represented a by mechanical impedance. Figure 1 shows the equivalent impedance representation for a conductor span with a single damper located at its centre. Here the point of excitation and observation are also at the center. The velocity of the conductor may be expressed
as

Figure 3: Simple model of

single-conductor damper

It can be seen that the damper behaves like a mass joM below the resonant frequency, (wM)2/D + jLM at resonance tIM = K/i, and as a damping element D at higher frequencies. From equation (D) it can be shown that the resonant nature of a conductor span is eliminated when the impedance of the damper is real or dissipative, and numerically equal to /ifiK. This agrees with the findings of Tompkins et al[1]. When the damper impedance is either very high or very low, the travelling waves will be reflected by or transmitted through the damper without dissipating energy in the damper. Thus the vibration of the conductor is not damped. Analysis shows that the range of impedance values for effective dampers is 0.5 VTim < Z < 3VT-Y. This is demonstrated by the measurements that follow. To provide the necessary damping for single conductors, the damping element must work against a An effective damper reasonably stationary platform. with minimum mass is obtained when the mechanical impedance of the mass M is numerically equal to the the lowest aeolian damping term D = kvTi' at The choice of the spring frequency to be damped. and the constant is not critical as long as D < K/w static deflection of the spring supporting he mass is not -xcessive. The X or f value is typically 30 rad sec oi 5 Hz for traonsmission line conductors, and 60 or 10 Hz for smaller distribution conductors rad sec
0

V2//)

oF =2Z0cothyk/2
Zo
coth

(D )

2 rI12
I

Zo

cothT.:I/2

V4~~~~~~~~~~~~~~~~IF~
1/2

1/2

Figure 1: Mechanical impedance representation of single conductor span with damper


This expression is the same as equation (A) except for the damper impedance Z which is added to the distributed mechanical impedance of the conductor. It should be noted that equation (D) also applies to a test span used to measure the internal damping of conductors when the impedance Zg of the force generator is included. In this case Z=Z . The impedance of the force generator is defined in (he Appendix. The physical arrangement described in the above example is not practical because a damper located at the center of the span would be at the node of all even harmonics, and as a result it would be completely practical For ineffective at these frequencies. reasons, dampers on single conductors should be located near one or both ends of the span, with the distance determined by the aeolian frequency spectrum, the conductor internal damping, and the damper characterA typical impedance representation of a istics. conductor span with two dampers is shown in Figure ?
ZO
Coth yx1

and groundwires.

ZA

ZB

ZO coth yx2

X1

(1/2- xi)

,1'
FIGURE 2

B2 - x2)

,-x2

MECHANICAL IMPEDANCE REPRESENTATION OF SINGL-E CONDUCTOR SPAN WITH TWO DAMPERS

As described earlier a damper should not be placed at the exact center of a conductor span because all even harmonics would have a node at that location. This applies to all points of symmetry along the conductor. For example, dampers located at the 1/3, 1/4, 1/5 etc points of the span would fail to provide vibration at protection every 3rd, 4th, 5th etc harmonic respectively. When this process is continued, it is found that the most practical and effective location for one damper on a single conductor span is at a distance x 1 from one of the terminals. With a single damper, x is chosen so that the frequency at which the damper would be at the standing wave node, and hence become ineffective, is high enough for the internal damping of the conductor to provide the necessary For larger conductors at normal damping. design tensions, this frequency would be in the 40 to 50 Hz range, and the corresponding distance x in the 1.2 to m range. For smaller conductors, the above 1.8 are and the frequencies higher distance x1 correspondingly smaller.
1

3247

With higher tensions and/or large conductors, it becomes necessary to employ two damnpers per span. The distances x1 and x2 of the dampers from the two terminals are chosen so as to cover the complete aeolian frequency range. To achieve this, the first damper is located at x1 to cover the low frequency end of the aeolian frequency spectrum, and the second damper at x to cover the high frequency end. For the conductor sizes used at transmission voltages, x would be in the 2.4 to 3.6 m range, and x2 in the l.0 to 2.2 m range.

elements. The damper impedance was measured at the resonant frequencies of the combined conductor-damper Since system using the,arrangement shown in Figure 1. the conductor impedance at an antinode is extremely small at resonance frequencies, the measured response is that of the damper with the conductor providing a With the low-loss suspension system for the damper. damper impedance now known, the damper was located near the ends of the span, as would be the case in practice, and the response obtained for the conductor-damper system corresponding to Figure 2. For each of the measured vibration responses, corresponding responses were produced by a computer program implementing the models shown in Figures 1 and 2. The inputs to the program included the effective tension and stiffness of the conductor which were the determined from a least-squares curve-fit of resonant frequencies; conductor mass per unit length; internal damping of the conductor, also determined from the measured data; and damper mass, spring constant, damping and location. The impedance of the shaker was also measured and is included in the model. The modeland measured responses are graphically generated compared in the figures that follow.

If, for example, the ratio of x /x = 0.5 or 0.67 is chosen, the damper at x; is effecBive when the other damper is at the first standing-wave node from the span terminal. However, with these ratios both dampers will be at the nodes of the standing waves at frequencies which are only 2 or 3 times higher than the first nodal frequency. If the first nodal frequency is 20 Hz, the two dampers will not provide the required damping at 40 and 60 Hz. To extend the frequency range for the two the first nodal damper arrangement to 5 times 100 Hz, the ratios of 0.4, 0.6, 1.4, frequency, i.e. 1.6 etc. should be used. The equations covering the placement of dampers on single conductors are

Xi

2f

= 2.4 to 3.6 m for 120 < v < 150 m

sec1
(F)

VERIFICATION OF THE CONDUCTOR-ALONE MODEL Measured Responses

and 20 < f1 < 25 Hz

X2

(0.4 or 0.6)x1

For smaller conductors, x1 and x2 are considerably smaller because the peak of the aeolian frequency spectrum to be damped is around 30 Hz rather than the 10 to 15 Hz range used in the above analysis.

A typical vibration response for a conductor without dampers is shown in Figure 4. This response was obtained with the force applied at the center of the span, and the velocity measurement made at the same Figure 5 is a plot of the measured response location. for the same conductor-tension combination obtained at 1.52 m from one end of the span.
t1

APPROACH TO THE MODEL VERIFICATION PROCESS


To check the validity of analytical models of a vibrating conductor span with and without dampers, a systematic examination of all conductor and damper parameters is required over a complete range of vibration several at vibration frequencies, and amplitudes. The parameters to be examined are conductor tension, mass, damping and stiffness; damper mass, sDring constant, damping and location; and excitation

-MODEL

...... EASLRED x----xCOPLING

- - -/(2*ZO) oANTI-RES'S

z
u

tco-,
io2

t>rce location.

- 0

The conductor-alone model was verified on a 23.5 m test span with the conductor held rigidly at its terminals, and excited by a constant-amplitude sinusoidal Measurements were made by the force-velocity force. method.[7] To cover the range of conductor mass and stiffness;of ACSR conductors, three sizes were included in the study: 585 (26/7), 1277 (42/7) and 1843 (72/7) kcmil. For each conductor size the tension Was adjusted to three different values: 15%, 25% and 40% RTS. The excitation forces were adjusted to three different levels so as to cover the range of amplitudes which occur with natural wind excitation. At each constant force level, the frequency was varied from 3 to 70 Hz. Most of the measurements were made at the center of the test span in order to simplify the analysis. However, a few measurements were performed at 1.52 m from one ern to check the analytical models and to illustrate tbcomplexity of the resultant response.
All1 the verification-tests3 of the conductor-danp 'nodels were performed on a 23.5) m test span of 5o kcmil conductor at 25% and 40% RTS with a simple viscous damper attached. The damper was constructed th along the lines of Figure 3, which allowed independent variation of its mass, spring and dampi

- e 4: Response of 1277 kcmil conductor at 25% for F =.044 N


0

RTS

calculated and drawn on both graphs to show the various A;curve has also been drawn for the meassymmetries. ured transformer coupling of the shaker's drive coil into the built-in velocity-sensing coil. This represents the "apparent" velocity that is measured with the armature blocked. The ratio of "apparent" velocity to excitation force is independent of amplitude. Due to the unknown phase relationship between the coupled signal and the real velocity signal, the magnitude of the error in the vicinity of the "coupling" curve can not

In Figures 4 and 5, the neasured data points have been superimposed uipon plots of the corresponding model-generated responses. A 1/(2ZO) value has been

3248
be defined. It ranges from zero, when the magnitude of the real signal is the same as its vector sum with the coupled signal, to very large values, when the span is in antiresonance. Therefore, true velocity measurements cannot be made in the vicinity of the coupling curve using the velocity coil. The measurements of antiresonances, shown in Figure 4 below the coupling level, were made using a linear velocity transducer (LVT) connected to the conductor at the shaker attachment point.
io i lo 10

10 I

F-. 13 N

F-1.3 N

--F_-J3

10

* U
IL.
49

101

10-2

O.'
,O,,

:;z
u

jo-i to-3 to -C
FREQUENCY
10t

(HZ)

10 2

tto
---

Figure 7: Measured response at resonance of 1843 kcmil conductor at 15% RTS

General Shape of the Conductor Response


In general shape, the measured and calculated responses shown in Figures 4 and 5, and those obtained for other conductors, are virtually identical. The resonant frequencies match to within 1% and the amplitudes of the peaks are basically the same. Minor differences are noticeable in the measured high amplitude peaks at low frequencies due to slight variations in the shaker damping. The model predicts that the shape of the conductor (velocity divided by force) response at the center of the span should be a hyperbolic tangent which, in a semi-log plot, is symmetric in amplitude about the

40

45

FREGUENCY (HZ)

Figure 5: Response of 1277 kcmil conductor at 25% RTS for F =.044 N, off-centre shaker
The response curves for other conductor-tension combinations and for other force levels are similar in shape to those shown above. Differences occur in the spacing between resonant frequencies, the 1/(2Z ) line, the height of the resonant peaks due to variaFions in conductor damping, and the skew of the peaks due to conductor-stiffness effects. These effects are discussed in the sections that follow.

The data shown in Figures 6 and 7 was obtained at resonance frequencies, and are intended to show effects of vibration amplitudes on conductor damping and resonant frequencies. Shaker mechanical losses limit 8-he loy frequency response to values below about 2 m.N .sec which is very noticeable in Figure 6. Excluding this region, the conductor loss can be seen to be proportional to frequency squared at low force levels, and approaches frequency to the exponent 1.5 at high force levels. The drop in resonant frequency as the excitation force is increased can be readily seen in the high frequency part of Figure 7.
span the

1/(2Z0) line, with the resonant and antiresonant peaks determined by conductor damping. The response is also expected to be symmetric in terms of frequency. That is, for idealized zero conductor stiffness, the width of the upward or resonant peaks, is the same as the width of the downward or antiresonant peaks at the 1/(2Z 0 ) line. For real conductors with non-zero the peaks spread further apart as frequency stiffness, increases, but the widths of adjacent peaks should remain approximately equal. These symmetries are further altered by use of an- electromechanical shaker for performing the vibration measurements.
The response obtained at the shaker is the combined response of the conductor-shaker system. The shaker loss limits the amplitude of resonant peaks at lower Above the frequencies. shaker resonance frequency, the shaker armature mass skews the resonant peaks downwards; below the shaker resonance frequency, the stiffness of the armature support skews the resonant peaks upwards. The smaller the characteristic impedance of the conductor, the larger the effect of shaker impedance. These effects are illustrated by the calculated responses shown in Figure 8.

10 1
U

OF-.03 N

oF-.11

--.F-.11N N

10 o

o..,

A
C. 0

5 10-t

4J

10

2 to 10

to1 FREQUENCY

(Hz)

to 20

t; gure 6: Measured response at resonance of 585 kcmil conductor at 15% RTS

3249
Effect of Conductor Stiffness

U)

10)

o0

10

24

at 151TofrF=.2

oo

Mesue
10 0 5

vauso
10 15
20

wr
25
30 35
40

bandorpial
45
50

55

60

65

70

FREQUENCY

(Hz)

In all the measured responses the separation between successive resonant frequencies was observed to become larger with increasing frequency. This is due to conductor stiffness, and has been observed by others including Claren and Diana.[5] Conductor stiffness has the effect of increasing the travelling wave velocity with frequency. The spacing of resonant peaks in the calculated responses was matched to the spacing in the curvefitting the measured by responses measured resonant frequencies to obtain effective values for tension and stiffness. These values were then used in All resonant frequencies, except the the model. fundamental were fitted to within 1%. The fundamental frequencies appear to be slightly affected by terminal effects, and differed by as much as 5% from the calculated vibrating string frequency.
It should be noted that conductor stiffness is also amplitude dependent. At very low amplitudes, the resonant frequencies are close to those expected with the conductor behaving as a solid aluminum rod. As the amplitude increases, the conductor becomes less stiff and the resonant frequencies are reduced. This amplitude dependence produces a skew in the resonant frequency peaks of the measured responses. Skewed distortion was very noticeable for the largest and stiffest conductor, the 1843 kemil, when tested at a This skew high force level, as shown by Figure 10. could be incorporated in the model by making the travelling wave velocity both frequency and amplitude dependent. For the 1843 kcmil conductor ah 15% RTS, theffective stiffness varied from 3300 N-im to 2700 N-n when the force excitation was varied from 0,13 N rms to For comparison, the solid 2aluminum rod 13 N rms. stiffness was calculated to be 7596 N-m , and the sym of individually-acting-strands stiffness as 77.6 N-mr Both 'these values were determined using the methods Amplitude-dependent outlined in Chapter 1 of Ref.[7]. frequency the stiffness is also responsible for difference between the measured and model-predicted The effective shown in Figure 4. antiresonances stiffness of the 1277 kcmil conductor varied2from 2680 N-im at the antiresonances to 2350 N-m at the resonances. This latter value was used in the model.
10t

Figure 8: Calculated respo'nses showing the effect of shaker impedance on 585 kcmil conductor at 15% RTS for F =.028 N
0

were obtained graphic'ally Neasured values of Z from the centres of symmetry of the measured responses, and are compared with calculated values in Figure 9. The average value of the measured Z divided by the calculated Z was 1.00, with a standar8 deviation of Considering that the accuracy of the graphical 0.05. method was about 5%, the agreement is excellent.
1.5
0
0 N w
4

1.4

1.3

1.2
tl4

LA
0 w

i .1
X
--

.
X'

--

'X'''

-----

---

-y

a:
U)

.9

.8
.7

.6
.5
50

ioo

150

200
Zo

250

300

350

400

450

500
10

Meapured

(N-seq/rm)
Z
U)

Figure 9: Comparison of measured and calculated ZO


Antiresonances were measured for the 1277 kcmil only. The results of these measurements are shown in Figure 4. Due to the extremely low level of response at antiresonances, the measurements were made at the highest force level used (13 N), and with the highest sensitivity LVT commonly available.

t-

conductor

u L
0

oo-2

10

o >A

1074

The antiresonance measurements at a force level of 13 N were plotted on the response for the same conductor and tension which was measured at the much lower force level of 0.044 N. This was done because the resonant response at low amplitudes was found to be approximately independent of the excitation level. For the extremely small amplitudes of vibration which result at antiresonances due to a-force of 13 N, the response should also be independent of excitation amplitude. Indeed, the vibration response shown in Figure 4 is very close to being symmetric with respect to amplitude, and has the expected hyperbolic shape.
antiresonance frequencies are calculated The slightly shifted from the measured values. This is due to conductor stiffness effects.

10t-6

Figure 10: Response of 1843 kcmil conductor at 15% RTS for Fo 13 N


Location of Excitation Force

Up to this point all of the discussion has centered around measurements of the vibration response with the shaker at the span centre. This location of the excitation greatly simplifies the analysis of the

3250

resultant response. However, even harmonics of the span cannot be excited at the centre and the models should also be verified for off-centre excitation. A location of 1.52 m from one end of the span was chosen for the response shown in Figure 5.
In this case the shape of the measured response is considerably more complex. The response is no longer symmetric about the 1/(2Z 0.) line with respect to amplitude or frequency, except where the shaker is at, or near, an antinode of the conductor, as occurs at 23 Hz. The amplitude of the resonant peaks is modulated by the shaker location as its attachment point changes from being an antinode to a node. The shaker is close to a node at about 55 Hz. Note that the frequency at which the shaker is at a node is more than twice the frequency at which it was at an antinode. This is due to the increasing effect of conductor stiffness with increasing frequency. In order to fit the calculated and measured responses, it was necessary to locate the shaker 1.36 m from one end of the span model instead of 1.52 m. Since while the the model assumes pivotted terminals, conductor was in fact rigidly clamped, the 0.16 m difference is the effective amount by which rigid clamping reduces the length of the freely vibrating span.

these curves, it appears that the range damping is approximately Z /2 < Z < 3ZO.
10~~~~~~
Optimum
-- 0

of

effective

Overdamped

......

Underdamped

1t0

L*UNOAMPEO
U0
C,.................................

o-3 0

10

15

20

25

30 35 40 45 50 Frequency (Hz)

55

60

65

70

75

Figure 11: Measured responses demonstrating optimum

damping

VERIFICATION OF THE CONDUCTOR-DAMPER MODEL


A simple viscous damper, as shown in Figure 3, was for the conductor-damper model verification process. This damper allowed the individual control of the damper parameters of mass, stiffness, damping and location. In the plots that follow, this damper was used to demonstrate the effect of each of these parameters. The linear behaviour of the damper components makes the response plots more easily understandable. Although there are slight nonlinearities in the damping which produce some small differences between the calculated and measured responses, these are far less than would be found in a non-viscous damper. Also, the simple damper allows a more general verification of the model than would be possible if a commercial damper such as the Stockbridge was used. The measured and calculated responses of the conductor-damper system are summarized in Figures 11 to 18. All measurements were performed on a 23.5 m test span of 585 kcmil ACSR conductor at 25% and 40% RTS.

10 i

chosen

-j
U

10

LU.
.0

"IL o
4J .,t

L 0

10 -1

102

_l

27O
10 0

10

15

20

25

30 35 40 45 50 Frequency (Hz)

55

60

65

70

75

Figure 12: Calculated responses demonstrating optimum damping

Effect of Damper Damping and Location


The measured and calculated responses, shown in Figures 11 and 12 respectively, were determined for two identical dampers located symmetrically 1.73 m from the terminals, with the shaker and observation point at the The damping value was varied by about a span centre. factor of four on either side of the theoretical optimum of Z for each damper. The spring element was 3800 Nm-1, and the mass 2.9 kg, except for the overdamped case were the mass had to be increased to 5.8 kg to provide a stable platform for the higher impedance damping element. The undamped conductor response and the 1/(2Z ) line are drawn for reference. The line is the response an optimally1/(2Z ) terminated or infinite line will exhibit, and thus represents optimum damping. It can be seen that the response measured for the damper approaches this line when the damper is near an antinode, such as 15 or 60 Hz, and approaches the undamped conductor response when it is near a node, at about 45 Hz. The over and underdamped cases can be seen to be less effective over the frequency range shown. It appears that slight overdamping is preferrable to slight underdamping. From

In this paper, "optimum" refers to conditions produce maximum energy dissipation by the damper when placed at a conductor loop antinode, and thus minimum conductor vibration amplitude, for any given excitation in the aeolian frequency range.

which

The effectiveness of the dampers can be improved over the symmtrical case by staggering the dampers as was discussed above. Figures 13 and 14 show this for a ratio of x.1/x,-0 .6 using the "optimum" damper of Figures 11 and l2.

"optimum"

10

X1 '
Z

2.29 m, x2

1.33m

100

~~~~~~~~~.........................
L ~~~~~~~~~~~~~~~~~~~~~~~~~~............. *- -*. . . .FOAPE
UNDAMPEO

10

U'

107

.............................................................................................................................

i 2Zo

lo~ ~ ~ ~l ~ . . .- NOHE
3251

'' ..

------

---

~~~~~~~~~~~~~~UNDAMPED

2ZO

10

J0

15

20

25

30

35

40

45

50

55

60

10

10

15

20

25

30

35

40

45

50

55

60

65

Frequency (Hz)

Frequency (Hz)

Figure 13: Measured responses showing effect of staggering the dampers

Figure 15: Measured responses showing effect of damper mass

101
-

, -

1.

73 m

10
1.39
m

x*

2.29

m,

X2

0.57 kg.

2.9 kg.

---

4.2 kg.

...............~~~~~~~~~~~~~~~~~~~~................... N UN~ ~ ~ ~ ~ ~ ~ LDAMPED

................. UNOAMPEO ------

U.

L 0

l. o-i
0 LL

...~~~~~~~~~~~~~~~~~~~~~~~~~..................

N,

lo.

t.

io,<2

so

Frequency (Hz)

Frequency (Hz)

Figure 14: Calculated responses showing effect of staggering the dampers


Effect of Damper Mass
The "optimum" damper is compared with dampers of reduced and increased mass in Figures 15 and 16. All dampers were located symmetrically 1.73 m from the terminals, with only the mass element varied. It can be seen that increasing the damper mass above 2.9 kg will not provide a significantly better platform for the The impedance of the 2.9 kg mass damping element. at becomes equal to that of the damping element and Z On the other Sand, 9.7 Hz, the design frequency. reducing the mass seriously affects the low frequency The design frequency corresponding to effectiveness. 0.57 kg may be calculated to be 50 Hz, which is damper approximately the frequency at which this becomes effective.

Figure 16: Calculated responses showing effect of

damper

mass

Effect of Damper Stiffness

Figures 17 and 18 show the effect of increasing the damper stiffness above ZO at the damper design frequency f0. The "optimum" damper was again uied as a The reference and its response is labelled 3800 Nm . dampers were located at the span centre so that the predominantely low-frequency effect of the stiffness element could be observed uncomplicated by damper position effects. At f =9.7 Hz, a stiffness _f 3800 has an has an impedance o? 0.35 Z0, and 16200 Nm Nm impedance of 1.5 Z . It can be seen that mid-frequency effectiveness is slightly improved as the resonant frequency of the damper is shifted upwards, at the expense of low-frequency effectiveness.

3252
CONCLUSIONS
New analytical models have been developed and verified for vibrating single conductor-damper systems. With the aid of these models, the optimum design parameters of dampers have been established for an Optimum locations of ideal viscous damping medium. dampers have also been established.
1 ~~~~~~~~~~~~~~~UNDAMPED

'I

Iz

to

fi

~ ~ ~ ~. . . . . . . . . . . . .
lo- ................

REFERENCES

IOU.
0

...............

[1] Quantitative Relationships in Conductor Vibration


Using Rigid Models, J.S. Tompkins, L.L. Merrill, B.L. Jones, AIEE Transactions October 1956.

I,o-2
............................................................

[2)
5

................................................

Conductor Vibration in Developments Recent Research, C.B. Rawlins, Alcoa Technical Paper No 13, 1958. Conductor Vibration - Theoretical and Experimental Investigations on a Laboratory Test Span, T.O. Slethei, J. Huse, IEE Proceedings, June 1965.
Decay Test to Transmission Line Conductors, A.R. Hard, R.D. Holben, IEEE Transactions, February 1967.
Line Transmission of Analysis Vibration, R. Claren, G. Diana, IEEE Transactions, December 1969.

l- e

10

15

20

25

30 35 40 45 50 Frequency (Hz)

55

60

65

72

[3)

Figure 17: Measured responses showing effect of damper stiffness

[4] Application of the Vibration


[5] Mathematical

to I
z

3800 N/l

---

16200

N/l

[6)
\
0

Vibration

*-- - ..

/ ~~~~~~UNOAMPED

Damping on Long Fjord Crossings. IEEE M. Ervik, Investigations, Theoretical Transactions, April 1981. Induced

L&
(.

.~~~~~~~~~~~~~~~~.................

[7)

Transmission Line Reference Book, Wind Conductor Motion, EPRI Publication, 1979.

NOMENCLATURE
io2
-

A,B
2ZO

- Constants
-

Dc
40

to-3

10

15

20

25

30

35

45

50

55

60

65

70

75

Internal damping p r ugit length of m conductor, kg sec


Linear damping of
a

Frequency

(HZ)

damper, kg

sec

-1

Figure 18: Calculated responses showing effect of damper stiffness


Summary

g
-

Damping of force generator, kg

sec

EI
F
F
0

Flextural stiffness of conductor, Nm

The mass and spring requirements for a damper can be summarized by the following general equation D = woM the to be matched The damping must > K/uk. characteristic mechanical impedance of the conductor. If the mechanical impedance of the damper is very high in relation to the characteristic impedance of the conductor, it will force a node at the damper location and as a result become ineffective. If the impedance is very low, it will also become ineffective. Slightly too much damping is better than too little. When more than one damper is required per span, the dampers should be placed at distance ratios such as 0.4 or 0.6 to extend the frequency range of effective damping to beyond 100 The correspondence between the measured and Hz. model-generated responses is good, with only minor differences due to non-linear behaviour of the damping element. Therefore the model shown in Figure 2 is valid the behaviour of conductor-damper simulating for systems for motions in the vertical plane.

Sinusoidal lateral force, N Sinusoidal excitation force, N

Fa

Equivalent sinusoidal aerodyndmic force per unit length of conductor, Nm


Frequency, Hz
Lowest design frequency at which damper becomes effective,Hz

f
f
0

K
K
g

Spring constant of

damper, Nm1

Spring constant of force generator, Nm1


Damper design constant Span length,
m m

m
M

Conductor mass per unit length, kg

Damper mass, kg
Armature mass of force generator, kg
Number

MN
p

(1,2,3....)

3253
s

- Dimensionless stiffness correction of velocity

The practical conductors with flexural stiffness EI the propagation velocity can be shown to be equal to
v

T
t
u

Conductor tension, N
[1

2T

(Q) 2I

(1

s)

(9)

- Time, sec - Harmonic number - Conductor velocity, m sec 1


-

VY
v x
x
0

Conductor

velocity

at x0 *

m sec

are obtained

The arbitrary constants A and B in the above force and velocity equations can be evaluated in terms of terminal conditions. When the forces F1 and F2 and velocities V1 and V2 at the two terminals x = 0 and x = Q are introduced the following general expressions

- Propagation
- Distance

velocity,

m sec

1
m

F(x,w)

F1 coshYx + F2 coshY( 9- x) +

along the conductor,

V1Zo sinhyx
V(x,W)
=

V2Zo sinhy( t -

x)

(10)

- Distance to the point of excitation, m


- Mechanical impedance of the damper, kg sec
-

V1 coshyx + F

Y0

sinhYx

ZO

Charactiristic kg sec

impedance of the conductor

V2 coshy ( Q - x) + F2Y0 sinhY(9- - x)

(11 )

- Radial

frequency, rad

sec

co0 0
ct 8 Y

- Lowest radial Radial design frequency at which damper becomes effective, Hz


- Attenuation cqnstant per unit

These equations can be used to saatisfy any particular set of terminal and excitation cconditions. For example, when a conductor is rigidly heold at both terminals V1 V2 = 0 and it is excited by a sinusoidal force F0 at x x0

Fo
V

conductor, m

length of

coshYx0 + F2 coshY( Q x) F1Y0 sinhyx0 = F2Y0 sinhY(- x0)


1
-

(12)

(13)

- Phase coqstant per unit length of conductor,

From above,

(rad) m

of conductor, m

Complex propagation constant per unit length


APPENDIX

F1

Fo

sinhy(Q - x )
sinh y9

(14)

F2 Fo
F(x,Q)
=

sinyx0

sinhyQ

(15)

VIBRATING CONDUCTOR-DAMPER MODEL The behaviour of a flexible conductor at any point in a span can be described by the following partial differential equations

After substitution

Fo
FOYO

sinhy(t-x )coshyx + sinhyx coshY(Q-x) sinhYk


sinhY(Z-x
)sinhYx
s inhY

(16)

aF 3x
DV

D cV
1

+m3t
'

(1)
(2)

V(x,w)
V(x,w)
For x
=

for x <

xo

(17)
(18)

-x T

at

FY00

sinhyx sinhy(9Q-x)

sinhyPZ

for x > x 0

For sinusoidal excitation the differential equations are obtained (D0 +' amM)V dx =(c im) dV Q 1
dx
dF

following ordinary
(3)
('4)
follow-

x0

= Q/2

F(Q'/2',) =

Fo
V
=

(19)

jco-1F T
= =

V(Q/2,W)

-2- tanhyt/2
0

F 2Z
0

sinhoc9 + coshoiQ +

cosf-Q

jsin8k

(20)

The solution of these equations takes the ing general form

F(x,w)

AeYx

Be-Yx

(5)
(6)

The above exhibits a number of equation harmonically related resonances at cos,B = -1 and antiresonances At resonant cos,3t = 1. frequencies f = (2u-1)v/29, the velocity and the displacement of the conductor become 2F
=

V(x,O)
where Y y

Y0 (AeYx - BeYX)

V (k/

a))

VO=
c

(21 )
(22)

(7)
D
205
+
v

(8)

Y (U/2 X)

= c-

V0

2F 0

3254
It should be noted that because of the symmetry, all the even harmonics have a node at the center of the span. When the conductor is excited and observed at some other point, it will exhibit resonances at all frequencies given by f = uv/2Q.

Similarly

B=
where
Z

When a damper, or a force generator, described by its impedance Z, is attached to the conductor at the center of the span, equation (20) becomes

ZR2 cothy(C/2-x ) R2 2 2 2 ZR2 Zo zR -z0


F2

Z2
o

R2

(29)

V(~2,ct

2Z

,cothYt~/2

=-- = Z + Z o cothyx R2 v 2 2

(30)

+ Z

(23)

From this expression, it is apparent that for a dissipative Z = 2ZO the velocity of conductor at resonant frequencies is reduced to Fo/2Zo and the resonant behaviour of the is eliminated. span Effectively, this is equivalent to moving two dampers, each with the theoretical optimum impedance Zoo from the terminals to the centre of the span. The resultant of these coincident dampers is an impedance of 2ZO, and the span appears "optimally-terminated" at the driving point.
The mechanical impedance of a simple viscous spring-mass supported damper as viewed from the conductor and shown in Figure 3 is given by the following

jcM (D -j
D + j ((OM
-

(24) ()
used in
the experiments

The vibration generator had the following impedance

Zg

=
=

D +

(M

K K)
-

(25)

0.47 + j(0.307to

1670)

The impedances ZA and ZB shown in Figure 2 may be determined from parts of equations (10) and (11) as follows:

left

ZAZRi ZA + ZRi
Z

F1

coshy(x-x1)

+
+

V1

coshy(x-x1)

F Y

V1Z0

sinhY(x-x 1) sinhY(x-x )
(26)

cothY(x-x1) + Z Ri o Z o cothY(x-x ) + Z z0 Ri 1
= F where Z RlV

F1

1 =

Z1 + + Z o cothYx
x=.

(27)
/2

Combining equations (26) and (27), and setting


7

A_- o(2i
=

z2 cothy(k/2-x )

Z2

ZR
%'

zR -zo

Das könnte Ihnen auch gefallen