response of electrical transmission line conductors

153
RESPONSE OF ELECTRICAL TRANSMISSION LINE CONDUCTORS TO EXTREi^E WIND USING FIELD DATA by RADHAKRISHNA R. KADABA, B.E., M.E. A DISSERTATION IN CIVIL ENGINEERING Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY Approved May, 1988

Upload: anonymous-vi1lrh

Post on 02-Dec-2015

11 views

Category:

Documents


2 download

TRANSCRIPT

RESPONSE OF ELECTRICAL TRANSMISSION LINE CONDUCTORS

TO EXTREi^E WIND USING FIELD DATA

by

RADHAKRISHNA R. KADABA, B.E., M.E.

A DISSERTATION

IN

CIVIL ENGINEERING

Submitted to the Graduate Faculty of Texas Tech University in

Partial Fulfillment of the Requirements for

the Degree of

DOCTOR OF PHILOSOPHY

Approved

May, 1988

ACKNOWLEDGEMENTS

I would like to express my special thanks and sincere

gratitude to the chairman of my committee. Dr. Kishor C.

Mehta, for his guidance, everlasting inspiration, constant

encouragement and patience in teaching me throughout my

graduate study program. I am also grateful to Drs. Eric L.

Blair, H. Scott Norville, William Pennington Vann, and Y.C.

Das, the members of my advisory committee, for their helpful

review and constructive criticism of this dissertation.

The research was accomplished under the financial

support of the Bonneville Power Administration (BPA) and the

Institute for Disaster Research (IDR) at Texas Tech

University. Support of these organizations is appreciated.

Special thanks are extended to Mr. Leon Kempner, Jr. for his

assistance in providing the field data details of the BPA

project and encouragement. Also, I would like to express my

deep appreciation to the Chairman of the Civil Engineering

Department, Dr. Ernst W. Kiesling, for his encouragement and

support during the course of my graduate study.

I shall ever remain indebted to my sister Vijaya

Bhanuprakash, brother Udaya Kumar, sister-in-law Latha

Ananth, and my other family members for their kind blessings

11

and the moral support during my stay away from home.

Special thanks are given for the assistance and friendship

provided by my colleagues Suresh Jonnagadla, Marc Levitan,

Saranga Kidambi and Pankaja Kidambi.

I must acknowledge with deep appreciation and

encouragement of my brother, Sri. Anantharamu, who instilled

in me the value of education and stood by my side giving all

the moral support he could and most of all his kind prayers

and love.

Finally, I would like to express 'Thanks' to my

parents, to whom I dedicate this dissertation.

Ill

11

vi

CONTENTS

ACKNOWLEDGEMENTS

ABSTRACT

LIST OF TABLES viii

LIST OF FIGURES x

CHAPTER

I. INTRODUCTION 1

Objectives ' 5

Contents of the Dissertation 6

II. STATE OF KNOWLEDGE 7

Response 8 Mean Response 10 Fluctuating Response 12

Aerodynamic Admittance Function 16 Mechanical Admittance Function 19

Peak Factor 20 Wind Characteristics 24

Mean Wind Speed 24 Mean Wind Profile 25 Turbulence Characteristics 28

Turbulence Intensity 30 Gust Spectrum 31

Davenport Analytical Model 34 Conductor Damping Ratio 37 Conductor Fundamental Frequency 38

III. FIELD DATA 39

Description of Test Site 41 Instrumentation 43 Data Accjuisition 47 Recording Procedure 48 Description of Recordings 50

IV

IV. FIELD DATA ANALYSIS 52

Validity of Wind Data 53 Power-Law Exponent for Wind Profile 60 Turbulence Intensity 65 Kaimal's Gust Spectrum Constants 68 Validity of Conductor Response Data 74 Effective Conductor Force Coefficient 80 Response Spectrum 85

V. COMPARISON AND REFINEMENT OF THE ANALYTICAL MODEL 90

Comparison of Analytically Predicted Mean Scjuare Response With Field Measured Values 91

Field Measured Mean Square Response 92 Analytical Model Predicted Mean Scjuare

Response 94 Refinement of the Analytical Model 104

Background Response 104 Determining the JAF Coefficients 108

Resonant Response 115 Determining the Aerodynamic Damping Ratio 115

Resonant Response with Suggested Damping Ratio 123

Peak Factors 127 Probabilistic Peak Factors from Field

Data 130

VI. CONCLUSIONS 134

REFERENCES 138

ABSTRACT

Conductors are long, slender, flexible, and wind

sensitive structures. In most cases of transmission lines,

60-80% of wind loads coming to the support structures are

transferred from conductors. Thus, assessment of conductor

response due to extreme wind is an important part of the

overall prediction of wind loads on transmission structures,

Extreme winds not only contain high wind speeds but also

randomly fluctuating gusts. These gusts cause fluctuating

dynamic responses of conductors. Since the responses

fluctuate randomly, they need to be assessed in

probabilistic terms.

An analytical model for estimating dynamic response of

transmission line structures to wind loads is developed by

Davenport. The model can be verified using field data to

determine its effectiveness. Bonneville Power

Administration (BPA) has conducted experimental studies in

the field at the Moro site to collect wind and electric

transmission structure response data. During 1981-1982 BPA

collected 23 separate 12-minute duration records that

included wind speed, wind direction and conductor response

data. The conductor response records include load cell.

VI

transverse swing angle, and longitudinal swing angle

recordings.

The BPA data of wind and conductor response are

analyzed in detail to gain as much information as possible.

The analysis of wind data include determination of wind

characteristics of mean wind profile, turbulence

intensities, and gust spectra. The conductor response in

terms of peak responses, effective force coefficients, peak

factors, and response spectra are obtained. The response

spectra are further analyzed to obtain contributions of

background response and resonant response. Comparison of

the analytical model with field data reveals that the model

underestimated background response and overestimated

resonant response.

Results of these data analyses are used to improve the

analytical model to predict conductor response in extreme

wind. The significant improvement includes determination of

peak factors from upcrossing rates, refinement in the

expression for background response and determination of

conductor aerodynamic damping ratios from field data.

VI1

LIST OF TABLES

1. Typical Values for Gradient Height and Power-Law Exponent (ANSI, 1982) 28

2. Typical Values for Turbulence Intensity and

Surface Drag Coefficient (Kempner, 1982) 31

3. Field Data Records During 1981-1982 40

4. File Description of Mode 22 49

5. Mean Azimuth and RMS in Degrees of Wind 55

6. Mean Wind Speed in mps 57

7. Power-Law Exponent and Kaimal's Gust Spectrum

Constants 64

8. Turbulence Intensity 66

9. Mean and RMS Values of Conductor Response (Transverse Load Component) 78

10. Field Measured Conductor Effective Force

Coefficients 82

11. West Conductor Response Spectrum Data Analysis 95

12. East Conductor Response Spectrum Data Analysis 96

13. Central Conductor Response Spectrum Data Analysis 97

14. Fixed and Assumed Parameters Used in the

Analytical Model 99

15. Variable Parameters Used in the Analytical Model 100

16. Background Response of West Conductor 116

17. Background Response of East Conductor 117

18. Background Response of Central Conductor 118

Vlll

19. Estimated Aerodynamic Damping Ratios in

Percentages 122

20. West Conductor Resonant Response With 40% Damping 124

21. East Conductor Resonant Response With 40% Damping 125

22. Central Conductor Resonant Response With 40% Damping 126

23. Peak Factors for Conductor Response 129

IX'

LIST OF FIGURES

1. Fluctuations of Wind Speed 3

2. Fluctuations of Conductor Response 3

3. Conductor Force Coefficients Based on Wind Tunnel

and Full-Scale Tests (Davenport, 1980) 13

4. Elements of Response Spectrum Analysis 15

5. Example of a Random Variable Showing Upcrossings of a Given Threshold 22

6. Idealization of Gust Spectrum Plot Over an

Extended Range (Davenport, 1972) 26

7. Topography of Site and Orientation of Power Lines 42

8. Schematic of Tower 16/4 44

9. Elevation Along the Test Line (Vertical Scale Exaggerated) 45

10. Time History Plot of Wind Speed for Record NOl at 34.7 m on Tower 16/4 59

11. Mean Wind Speed and Direction Recorded at 34.7 m

on Tower 16/4 of 23 Records 61

12. Power-Law Plot for Record N15 63

13. Gust Spectrum Plot for Record NOl Recorded at 34.7 m on Tower 16/4 70

14. Time History Plot of West Conductor Response for Record NOl 76

15. Response Spectrum Plot for West Conductor Response for Record NOl 86

16. West Conductor Response Spectrum Plot for Record NOl 93

X

17. Analytical Model Background Response Versus Field Measured Background Response 102

18. Analytical Model Resonant Response Versus Field Measured Resonant Response 103

19. Frequency Transfer Function of West Conductor Response for Record NOl 110

20. West Conductor JAF Coefficients Contour Plot for

Record N15 112

21. Joint Acceptance Function Plot 114

22. Refined Model Background Response Versus Field Measured Values 119

23. Analytical Model Resonant Response With 40% Damping Versus Field Measured Values 128

24. Cumulative Probability Distribution of Upcrossings for Conductor Response 133

XI

CHAPTER I

INTRODUCTION

Electrical transmission line systems are engineered

structures that traverse over all types of terrain. Wind

loading is an important factor in the design of these

transmission line systems, consisting of towers, conductors,

and ground wires. Transmission line conductors are long,

flexible, and wind sensitive structures. Probably no other

structure has as much of its mass in highly flexible form,

and so continuously exposed to the forces of wind, as do

transmission line conductors. The loads due to the effect

of wind acting on the conductors, which in turn, transmit

loads to the supporting tower, are more than the loads due

to the wind acting directly on the tower itself. Wind loads

on conductors with spans of around 300 m account for 60 to

80% of the total wind load effect on the support tower

structure. Accurate and reliable prediction of wind loads

that are transferred from conductors to the towers are

desirable to produce an economical and safe design of

support tower structures.

Transmission line tower structures are usually designed

for five different types of loads (Kempner, 1985): (1)

extreme wind, (2) wind on ice, (3) National Electric Safety

Code (NESC, 1984), (4) broken conductor, and (5)

construction loading. Records show that more than 50% of

tower structure failures are due to extreme winds. Under

these conditions, any improvement in the understanding of

conductor behavior under extreme winds which leads to better

definition of loads and better accompanying design of the

tower structure is desirable.

Extreme winds not only contain high wind speeds but

also randomly fluctuating gusts (see Figure 1). The gusts

cause fluctuating wind loads on transmission line

conductors. Transmission line conductors respond to random

gust loading in a randomly fluctuating manner (see Figure

2). Response of conductors due to wind can be considered as

a combination of mean response (static) associated with the

mean wind and fluctuating response (dynamic) associated with

wind gusts. Fluctuating response is conveniently expressed

as the product of a peak factor, g, and root mean scjuare

(RMS) value of fluctuations about the mean response.

The response fluctuations about the mean response can

be represented in the frecjuency domain by a response

spectrum. A response spectrum is a plot of spectral density

values versus frecjuency. The area under the response

spectrum is ecjual to the mean square value of the

a

•a a> o Q.

en c

Mean + Sigma

Mean

Mean - Sigma

Time (minutes)

Figure 1: Fluctuations of Wind Speed

o c o a. a>

o t5 •a c o O

Time (minutes)

Figure 2: Fluctuations of Conductor Response

fluctuating response. Fluctuating response can be viewed as

background response due to low frequency wind turbulence and

resonant response near the natural frecjuencies of conductor

vibration. The fluctuating response is discussed in detail

in subsecjuent chapters. The use of a peak factor to

establish an equivalent static design load for conductors

due to wind is convenient. The peak factor is defined as

the number of standard deviations by which the peak value

exceeds the mean value.

There are a variety of traditional methods for

computing wind loads on transmission line structures that

estimate ecjuivalent static response of the structure to

these loads. These traditional methods are simple,

empirical, and usually conservative; therefore the resulting

designs are adecjuate in resisting the design wind.

An analytical model for estimating dynamic response of

transmission line structures to wind loads has been

developed by Davenport (1980). This analytical model

obtains the gust response factor in a single equation using

the frecjuency domain approach. The model considers all the

major wind characteristics and structural properties to

estimate dynamic response. Natural frecjuencies of the

towers are generally much higher than natural frecjuencies of

the conductors. Hence it is assumed that conductor response

is not influenced by the motion of the supporting tower

structure. With the above assumption, the response due to

wind on the conductor and on the tower structure can be

analyzed separately.

Wind tunnel experiments and full-scale field tests have

been conducted on transmission line structures. Because of

their slenderness and flexibility full-scale tests are

especially significant to assess dynamic response of these

structures. Full-scale tests are of great value to compare

and refine the analytical model. The most comprehensive

source of full-scale data is the experiments conducted on

John Day-Grizzly transmission line 2 located in northern

Oregon. These field data, collected by the Bonneville Power

Administration, are used in this study. The data include

simultaneous recordings of wind and transmission line

response during extreme winds. Analysis of field data of

wind and transmission line conductor response can assist in

substantiating and improving the analytical model.

Objectives

The general objective of this research is to determine

the dynamic response of conductors due to extreme wind using

field data. The Specific objectives are: (1) to assess

wind parameters from the field data, (2) to develop

probabilistic peak factors for conductor response using

upcrossing rates, (3) to determine the aerodynamic damping

and the joint acceptance function for conductors from the

field response data, and (4) to compare dynamic response of

conductors measured in the field with a refined analytical

model.

Content of the Dissertation

A brief description of the contents of this

dissertation is given here. The next chapter contains an

overview of the state of knowledge concerning wind

characteristics, and details of the analytical model to be

used to predict dynamic response of conductors. A

description of the measurements, site characteristics, and

instrumentation for the field data is given in Chapter III.

Analyses of wind and conductor response field data are

described in Chapter IV. Comparisons of field measured

conductor response data with responses calculated using the

analytical model are part of Chapter V. Determination of

probabilistic peak factors for conductor response using the

upcrossing rate principle, refinement of the background

response expression, and determination of conductor

aerodynamic damping from the field response data are also

presented in this chapter. Conclusions reached in this

study are presented in Chapter VI.

CHAPTER II

STATE OF KNOWLEDGE

Wind loads on transmission line conductors depend on

wind characteristics and on interaction phenomena of the

wind with conductors. Wind speed fluctuates randomly; it

can be considered to consist of mean and fluctuating (or

gust) components. A knowledge of both the mean wind speed

and the random fluctuations are recjuired to evaluate wind

loading. In addition, structural properties (natural

frecjuencies, damping, size, shape,..etc) play an important

role in prediction of response of the conductors in extreme

winds.

The difficulty of proper simulation of the natural wind

characteristics and scaling of transmission line structures

in wind tunnels leads researchers to depend on full-scale

experiments. There have been only a few full-scale

experiments for wind loads on transmission line structures.

Field measurement programs have been conducted in the United

States, Canada, Europe, and Japan to monitor the wind

response of transmission line systems. A review of each of

the field test programs is given in a report by GAI

consultants (1981). These field tests measured wind and

8

response data for a specific design objective such as

determination of the span factor or the gust response

factor. Detailed analyses of field data are not reported in

the open literature. Bonneville Power Administration (BPA)

has conducted field experiments on transmission line

structures since 1976. Results of analysis of BPA data are

available in several reports and papers (Kempner and

Laursen, 1977, 1979, 1981; Kempner and Thorkildson, 1982;

Ferraro, 1983; Norville, 1985). These reports and the

present study indicate that conductor response to extreme

winds is governed by wind turbulence characteristics,

aerodynamic characteristics of wind structure interaction

and structural dynamics of the conductors. The wind and

conductor response data collected by BPA during 1981-1982

are used in the present study. The literature on wind,

aerodynamics, and structural dynamics is very extensive.

Only the items that are obtainable through analysis of field

data and are pertinent to this study are discussed in this

section.

Response

The design of a transmission line tower structure is

generally based on the peak transverse load component of the

conductors when subjected to extreme winds. The transverse

load component transferred to the tower is the direct result

of the transverse response of the conductor. The the

transverse load component is considered as conductor

response in the present study. The prediction of this peak

response value, rather than a mean response value, is needed

for design purposes. Peak response is predicted by the

summation of mean and fluctuating responses. For a time

period, T, the peak response can be estimated by

ft = R + g aj (2.1)

where ft = peak response,

R = mean response,

a„ = root mean scjuare (RMS) of the fluctuating

response about the mean response, and

g = statistical peak factor.

In ecjuation 2.1, the mean response is based on the mean

wind speed. The fluctuating response is a product of peak

factor and RMS value of response. The dimensionless peak

factor, g, is probabilistic because of the random nature of

the fluctuating response. The peak factor is determined

from the probability of the upcrossing rate or,

ecjuivalently, a specified number of occurrences in a given

interval of time. The RMS of the fluctuating response

10

depends on wind characteristics such as the turbulence

intensity and structural characteristics such as the

damping, frecjuency, shape,., etc.

Mean Response

The mean response of conductors is obtained from the

mean wind pressure acting at the effective height of the

conductors. The effective height of the conductor is

considered as the average height of the conductor above the

ground level. The ecjuation for mean wind pressure is

1 -2 P = y P V C^ (2.2)

where P = mean wind pressure,

p = mass density of air, 1.226 kg m' , at 60 F,

at sea level,

V = mean wind speed at the effective conductor

height, and

C^ = conductor force coefficient.

The mean response of conductors, R, can now be expressed as:

R = P L d (2.3)

where L = effective conductor span, and

d = conductor diameter.

11

The mean response of a conductor depends on the mean

wind speed and the aerodynamic relationship in terms of

conductor force coefficient. The force coefficient converts

the stagnation pressure term (- pV ) in ecjuation 2.2 to a

transverse force on the conductor. In most cases, the force

coefficient is determined from wind tunnel tests and, in

general, it is a function of Reynolds Number, the angle of

incidence of the wind, and the shape and roughness of the

conductor. Published results of wind tunnel measurements

show wide variability in force coefficient values, because

of difficulty in simulation of Reynolds Number and varying

tests conditions (Potter, 1981).

The force coefficient of a cylindrical shape is

strongly influenced by the Reynolds Number, Nj^, which is

given as

Np = -B^ (2.4)

R \x

where V = wind speed,

d = conductor diameter, -5 -2

1 = dynamic viscosity of air, 1.79x10 N-sec m

at 60°F, at sea level, and

p = mass density of air.

12

A plot of force coefficient versus Reynolds Number is

shown in Figure 3. The region of the curve where the force

coefficient decreases sharply with Reynolds Number is called

the critical flow range. This decrease in force coefficient

is related to the transition from laminar flow to turbulent

flow. For a typical conductor diameter and design wind

speed, the Reynolds Number is usually above the critical

4 range (N„ > 5 x 10 ) . A constant value for the conductor

force coefficient is usually given in transmission line

design recommendations (ASCE, 1984). As indicated in Figure

3, full-scale measurements tend to give lower force

coefficient values than the wind tunnel experiments. These

discrepancies are not yet resolved in the published

literature. Additional data on force coefficients from

field measurements are desirable.

Fluctuating Response

Conductor response to fluctuating wind depends upon the

dynamic characteristics of the conductor as well as

turbulence in the wind. To determine the response of a

conductor subjected to fluctuating wind, frecjuency domain

methods are usually used. Frecjuency domain methods are

popular for computation because they are cost effective and

efficient. In frecjuency domain analysis, fluctuations in

13

Curve

1 2 3 4 5

Source

Wind Tunnel Tests Wind Tunnel Tests Wind Tunnel Tests Full-Scale Tests Full-Scale Tests

Conductor diameter (in)

1.125 0.770 1.695 1.108 1.602

1.2

o

o o

o o P

B o •o c o O

to

0.9

0.8

0.7

0.6

0.5

>

\

w

\ \

\

(D\

1 (1) \ \

\ \ V

^^-\ ^ —

/ /

^

^

/

1 1

i

Reynolds Number *10•^ NR

Figure 3: Conductor Force Coefficients Based on Wind Tunnel and Full-Scale Tests (Davenport, 1980)

14

the wind and conductor response are represented by a

spectrum. A spectrum is a plot of energy at each frequency

versus the frecjuency. Therefore, it represents a

distribution of energy over the entire frecjuency range. The

area under the spectrum is ecjual to the mean scjuare value of

the fluctuations. The frecjuency domain approach to compute

the peak response is briefly described below.

Several steps to obtain the mean scjuare value of

conductor response from the wind gust spectrum are shown in

Figure 4. The first step in the analysis involves the

transformation of the gust spectral density function, S (f),

into the force spectral density function, Sp(f), by

2 multiplying by the aerodynamic admittance function, x (f)-

The second step involves the determination of the response

spectral density function, S„(f), by multiplying the force

spectral density function by the mechanical admittance

2 function, H (f). The aerodynamic admittance function and

the mechanical admittance function are frecjuency response

functions. The third step is to calculate the mean square

value of the response, CT^, from the area under the response

spectrum. Once the RMS value of response, cjj, is obtained,

a peak value of the fluctuating response is determined by

Gust Spectrum

15

- 2 / 3

Force Spectrum

CO

Response Spectrum

Aerodynamic Admittance

logl

Mechanical Admittance

Figure 4: Elements of Response Spectrum Analysis

16

multiplying the RMS value by a statistical peak factor, g.

The peak response ft, is the addition of the mean response,

R, and the fluctuating response, <3 ^r^, as indicated in

ecjuation 2.1.

Field measurements of wind and conductor response

provide gust and conductor response spectra. Appropriate

analysis of field data leads to estimation of the frecjuency

response functions, as indicated in Chapter V. In addition,

tJie field data provide the peak factor in probabilistic

terms using the upcrossing rate principle. Theoretical

expressions for the aerodynamic admittance function,

mechanical admittance function and peak factor are presented

below.

Aerodynamic Admittance Function

For a given body immersed in a flow, wind fluctuations

can be used to determine the information on resultant forces

by empirical coefficients. The time-varying transverse

force on a body completely enveloped by wind is given by the

formula

F = [i. p (V + u)^ C^i A (2.5)

where F = transverse force,

p = mass density of air.

17

V = mean wind speed,

u = fluctuating component of wind speed,

C^ = force coefficient, and

A = area of exposure.

The time-varying fluctuating force is divided into two

components, mean force, F, and fluctuating force, F'. Then

ecjuation 2.5 can be expanded as

F -f F' = ^ p (V +2VU + U'') C^ A. (2.6)

2 If the term of the order u is neglected, the mean and

fluctuating forces can be separated as

1 -2 F = -i. p V C^ A (2.7)

and

F' = p V u C^ A. (2.8)

The power spectrum for fluctuating transverse force,

F', is then related to the gust spectrum as follows:

2 Sp(f) = (p V A C^) S^(f) (2.9)

or

_2

Sp(f) = ± ^ S^(f). (2.10)

V

18

Ecjuation 2.10, is valid over the range of frecjuencies

contained in the gust spectrum provided all effects remain

perfectly correlated. In practical conditions where gust

effects over the entire length of the conductor may not be

correlated, an adjustment factor or aerodynamic factor is

included in the ecjuation. This factor is called the

aerodynamic admittance function, x (f)/ and equation 2.10

becomes:

Sp(f) = ^ X^(f) S^(f)- (2.11)

The aerodynamic admittance function is a frecjuency

transfer function which transfers the gust spectral density

function to a force spectral density function. It accounts

for the correlation of gusts over the structure. The

distribution of gusts over the structure depends on the

relative size of the structure and the gusts. A large gust

totally enveloping the structure will be well correlated

over the structure, while small gusts acting over only a

portion of the structure are uncorrelated. In general,

low-frecjuency gusts are assumed to be correlated over the

2 structure; that is x (f) is assumed close to unity. The

2 value of X (f) fall below unity at frequencies in the range

19

of interest for the effects of winds on conductors. This

variation in aerodynamic admittance function as a function

of frecjuency is illustrated in Figure 4. The aerodynamic

admittance function for a structure is generally obtained

from wind tunnel tests (Blevins, 1977). In this study

coefficients of this function are obtained from the field

data (see Chapter V)-.

Mechanical Admittance Function

After the force spectral density function, Sp(f), is

obtained by means of ecjuation 2.11, the response spectral

density function, S„(f), is obtained by multiplying Sp(f) by

2 the mechanical admittance function, |H(f)| :

Sj (f) = |H(f)|2 Sp(f). (2.12)

2

The mechanical admittance function, |H(f)| , is

determined from an analysis using the stiffness, mass, and

damping characteristics of the structure. For a single

degree of freedom system, the mechanical admittance function

is the scjuare of the structural dynamic amplification

function; it is expressed as (Bendat, 1980):

H(f)|2 = J- ^ —^ (2.13)

k' f 2 2 f 2

o

20

where f = fundamental frecjuency,

C = damping ratio, and

k = spring constant.

The form of the mechanical admittance function is

illustrated in Figure 4. The resulting spectrum of the

response, shown in Figure 4, is peaked at the fundamental

frecjuency of the structure. This peak is the resonant

response of the structure at that frecjuency. One of the

major unknowns in the mechanical admittance function is the

dcunping ratio, C,. Damping can be due to structural and

material properties, and for wind response, aerodynamic

interaction. Structural damping can be assessed only from

experiments. A theoretical expression for aerodynamic

damping is presented in a subsecjuent section of this

chapter. In this study damping of conductors is determined

from field data; this is presented in Chapter V.

Peak Factor

Another important component in ecjuation 2.1 is the

statistical peak factor, g. Davenport (1977) has shown that

for a stationary random process, the statistics of the peak

response values may be represented by a Type I extreme-value

probability distribution. For this case the peak factor

21

corresponding to the peak response occurring in time period,

T, can be approximated as:

0.577 g = V2 In yT + ^•-^'' . (2.14)

V2 In yT

Where y is the cycling rate of the process; that is, the

number of times the mean response value is crossed per unit

time.

The peak factor has also been determined from the rate

of upcrossing. Melbourne (1975) used this principle of rate

of upcrossing as developed by Rice (1945) to analyze wind

tunnel aero-elastic model data.

Consider a continuous random process that can be

differentiated at least once. A sample function of the

random process is shown in Figure 5. The crossings of the

level x(t)=Ti, with a positive slope (upcrossings) are shown

in the figure. The number of crossings of the level in the

time interval T is a random variable. For a long period of

time the expected or mean number of crossings will approach

some fixed value. Based on this average value, the average

crossing rate can be determined.

Rice (1945) showed that the average crossing rate can

be computed for any stationary random process x(t), if the

joint density distribution is known for x(t) and x(t) (the

sample functions of x(t) being dx(t)/dt). The average

22

Up-Crossings

time, t

Figure 5: Example of a Random Variable Showing Upcrossings of a Given Threshold

number of upcrossings of the value x per unit time is

expressed as:

N (x) = j X p(x,x) dx (2.15)

where p(x,x) = the joint density of x and x,and

N (x) = the average number of upcrossings.

For a linear single degree of freedom system excited by

a stationary Gaussian load, the joint probability density

can be written as (Nigam, 1983)

23

1 2 .2 P^""'^) = ?nn\ e x p [ - J L - - ^ l ( 2 . 1 6 )

X X

2 where a = mean scjuare of x, and

2 a = mean scjuare of x.

There is no covariance term a in the density equation

2.16 because x and x are assumed to be uncorrelated.

Substituting ecjuation 2.16 into ecjuation 2.15 and performing

the indicated integration yields

1 "x x^ N ( ) = ^ —^ e x p { - ^ l . (2.17)

2" ^x 2al

For a narrow band random process, the spectral energy

is centered close to the fundamental frecjuency, f , of the

structure. Thus ecjuation 2.17 can be written as (Reelect,

1969):

x^ N = f e x p l - ^ l - (2.18)

2^x

The cumulative probability distribution in terms of

upcrossings can be stated as

-* 2 P(>x) = EJ2LL = exp{--^l. (2.19)

^o 2ol

24

The upcrossing rate formulation is for a narrow band

vibration process. The upcrossing rate technicjue is applied

to the data of conductor response in Chapter V to obtain

probabilistic peak factor, assuming that the conductor

vibrates at its fundamental frecjuency (as indicated by data

in Chapter IV).

Wind Characteristics

Wind fluctuates randomly both in time and space. Wind

speed over a given time interval can be considered as

consisting of a mean wind speed and a fluctuating component.

Knowledge of both the mean wind speed and the fluctuating

component assists in evaluating wind loads on transmission

line conductors. The mean wind speed, wind profile,

turbulence intensity, and gust spectra are presented as part

of this chapter.

Mean Wind Speed

Mean wind speed is defined as an average wind speed for

a specified time interval. The numerical value of the mean

wind speed can have large variations depending on the

interval used for averaging the wind speed. A shorter

averaging time leads to a higher mean wind speed value,

while a longer averaging time leads to a smaller mean wind

speed value. This is primarily due to short gusts of high

wind speed that last for short periods of time.

25

The length of the record for which the mean value and

the RMS value of wind speed are determined is somewhat

arbitrary. The record should be long enough to reflect the

effects of low frecjuency components of mechanical turbulence

generated by the terrain roughness, but short enough so that

a reasonably stationary time history, free of significant

trends is obtained. Analysis of the power spectral

densities of wind speed provides the background for an

appropriate selection of the averaging time interval for

mean wind speed. The gust spectrum reveals that wind is

made up of two distinct types of air flow: (a)

macrometeorological or climate fluctuations, and (b)

micrometeorological fluctuations or gusts. These

fluctuations are separated by a stationary stable interval

(spectral gap) between 10 minutes and 1 hour, as indicated

in Figure 6. Based on this spectral gap, mean values

averaged over 10 minutes to 1 hour are optimum for stability

(Davenport, 1972). In this study, the wind speeds are

averaged over record length of 12 minutes.

Mean Wind Profile

An important characteristic of wind is the variation of

wind speed with height. The surface friction effects of the

ground retard the movement of air close to the ground

26 xt

rum

2S.

Ene

rgy

1 I 1

1. 1 t

i

1

1

u. Ivctevh

Ik iun

11 yt \u»-HK.I i;yck

I* ;!

•1 ii 1 t 1 * ( •

ii ! ;

n 1 1

to-«

lOUUO

1 t 1

ArniM.!

Afui V M Dcr llovcn

- - • Spck:uUli»« JIICI .\ ti. Oa«cnp<Ml

*-fi*r

M«;ruiii*irar<tlu|Hal ruifc r ' . . . t l l . ( m.^ ' llMlVtlMIHI

Scnutliufn.!

MkiunwicafolofivJ ruifc IfKMtl

Figure 6: Idealization of Gust Spectrum Plot Over an Extended Range (Davenport, 1972)

surface. This retardation causes a reduction in wind speed

near the ground. At some height above the ground, the

movement of air is independent of the ground obstructions.

This unobstructed wind speed is termed the 'gradient wind

speed,' and the corresponding height at which the air

movement is not retarded is termed the 'gradient height.'

Mean wind profiles near ground level are currently

represented by either power-law or logarithmic-law profiles

The logarithmic law profile is based on the assumption of

physical phenomena and is valid particularly up to 30 m

27

above ground (Simiu, 1984). The power-law profile is

empirical and is assumed to be valid up to gradient height

(approximately 500 m). The power-law profile is primarily

used in structural analysis and design, because of its

simplicity. It is essential to determine the wind profile

at a particular site, so that the mean wind speed at

effective height of structure can be determined. Power-law

is used in this study, and is briefly described below.

The power-law profile was developed by Davenport

(1960). He modifiecj the exponential profile developed by

Brunt (1952) to obtain a mean wind speed profile. In

horizontally homogeneous terrain, it is assumed that the

power-law is valid with a constant exponent (a) up to the

gradient height, Z . Both gradient height and power-law y

exponent are functions of the terrain roughness. The mean

wind profile is expressed as

Ii£l = (-?-)" (2.20)

where V(z) = mean wind speed at height, z,

V = mean wind speed at gradient height, Z , and

a = power-law exponent.

The power-law is used in both the American National

Standard ANSI A 58.1 (1982) and in the National Building

28

Code of Canada (NRCC, 1980). Typical values of the gradient

height, Z , and the power-law exponent, a, for different

terrains, as specified by ANSI (1982), are summarized in

Table 1.

TABLE 1

Typical Values for Gradient Height and Power-Law Exponent (ANSI, 1982)

Terrain Category

Coastal Areas

Open Farmland

Forest/Suburban

City Centers

Gradient He ight(ft)

Z g 700

900

1200

1500

Power Law Exponent

a

0.10

0.14

0.22

0.33

Turbulence Characteristics

The fluctuating part of wind is termed as the

turbulence. The turbulence present in the wind flow is due

to the ground roughness characteristics of the terrain over

which it is passing or due to thermally-induced convection

or both. The turbulence due to ground roughness is known as

mechanical turbulence and that due to heat convection is

29

known as convective turbulence. Depending on the relative

importance of convective to mechanical turbulence, the

stability conditions of the atmosphere are classified as

stable, neutral, and unstable (Simiu, 1985). The extreme

winds in which structural engineers are interested are

categorized as neutral stability conditions. In a neutrally

stable condition, the temperature related buoyancy forces

and resulting vertical air motions are minimum. For

engineering purposes, it is generally assumed that neutral

atmospheric stability conditions can be assumed for wind

speeds higher than 20 mph. Details of atmospheric neutral

conditions are discussed in detail by Kancharla (1987).

Analysis of turbulence includes determination of the

turbulence intensity and the gust spectrum. Of these two,

the turbulence intensity expression is simpler. It

indicates relative amplitudes of the fluctuations compared

to the mean wind speed. A complete representation of

fluctuating components of wind is the gust spectrum, which

gives the distribution of the mean scjuare over the frecjuency

domain. The gust spectrum is useful in assessing dynamic

response of structures. Both representations of turbulence

characteristics are discussed below.

30

Turbulence Intensity

Turbulence intensity is a measure of the gustyness of

the wind. It is expressed as

T^ = -^ (2.21) V

where T = turbulence intensity,

a^ = RMS of wind speed fluctuations, and

V = mean wind speed.

In statistical terminology this number is often called

the coefficient of variation. Turbulence intensities are

higher for records which have lower mean wind speeds than

for records that have high wind speeds for the same terrain.

The turbulence intensity is strongly related to the terrain

roughness; a greater turbulence is caused by a rougher

terrain (refer to Table 2). A decrease in turbulence

intensity with height is also expected; at greater heights

both mean and RMS values of wind speed increase but the RMS

value increases less because the shearing action of the

ground surface is less (Jan, 1982).

Parametric study of the Davenport analytical model

(Twu, 1983; GAI, 1981) shows that the turbulence intensity

is the most influential parameter in predicting the response

of a transmission line structure to extreme wind.

31

TABLE 2

Typical Values for Turbulence Intensity and Surface Drag Coefficient (Kempner, 1982)

Terrain Category

Coastal Areas

Open Farmland

Forest/Suburban

City Centers

Surface Drag Coefficient

0.

0.

0.

0.

K

001

005

015

050

Turbulence Intensity

T u

0.

0.

0.

0.

L

07

12

22

39

Therefore, emphasis is given to the computation of

turbulence intensity in the wind data analysis.

Gust Spectrum

A randomly fluctuating phenomenon such as wind speed

can be conceived of as the superposition of a large number

of harmonic fluctuations with frecjuencies ranging from zero

to infinity. The spectral representation of turbulence is

related to this concept, and it provides information on the

contributions of fluctuating components (energy) with

various frecjuencies. The energy of any random process, like

wind, is usually expressed in terms of a cjuantity called the

32

'Power Spectral Density (PSD).' The PSD at any particular

frecjuency, f, may be considered as the average fluctuating

wind power passing a fixed point when the wind as a random

process is filtered by a narrow band filter centered at f.

In the dynamic analysis of structures subjected to gust

loading, significant dynamic amplification of the response

may occur at a resonant frequency, i.e., when the natural

frecjuencies of vibration of the structure and of the wind

match (Simiu, 1985). Flexible structures like conductors

can have dynamic amplification of the response because the

fluctuating component of wind has a fair amount of power at

frecjuencies of structural vibration. On the other hand, if

the natural frecjuencies of vibration of the tower structures

are higher than 1 Hz, the dynamic amplification of the

response of the tower will be small because power in the

gust spectrum at those frecjuencies (see Figure 13, page 72)

is very small.

There are more than a dozen specific wind speed

spectrum ecjuations developed for meteorological and

engineering purposes. Some of these spectral ecjuations are

discussed by Kim (1977). For neutral atmospheric

conditions, wind turbulence is generated by the surface

shear stress. It follows that the magnitude of the PSD

should be proportional to the scjuare of the frictional

33

velocity. The analytical model for the gust spectrum used

in this study was developed by Kaimal (1978).

In general, the spectral energy of gusts is a function

of the wave-length, —, of the wind fluctuations and the

height, h, above the ground. The analytical form suggested

by Kaimal for the horizontal gust spectrum for height to

•f v> wave length ratios greater than one-half ( >0.5) is given

V

as:

f S„(f) ^ = A ^ *

f h

V '^ (2.22)

where S (f) = spectral density of gusts at frecjuency f.

u^= friction velocity, (^KV^Q)

K = surface drag coefficient,

(typical values are shown in Table 2)

V..Q= mean wind speed at 10 m height,

V = mean wind speed at height h,

h = height above ground, and

A, n = constants.

Constants A and n represent the amplitude and exponent

values of Kaimal's gust spectrum. For neutral atmospheric

34

conditions, A=0.3 and n=2/3 are suggested by Kaimal.

Ecjuation 2.22 is useful for describing the gust

spectrum in the high frecjuency range (low wave lengths) and

for heights limited to the first few tens of meters. Kaimal

also presents other forms of the gust spectrum which are

valid for lower frecjuencies and for unstable atmospheric

conditions.

Davenport Analytical Model

On the basis of Davenport's analytical model (1980),

the peak response, ft, of a conductor due to fluctuating

winds is represented by ecjuation 2.1. Davenport's model

provides an analytical expression for the RMS, cjp, and

suggests a peak factor value, g, in the range 3.5-4.0.

Validation and refinement of the expression for the RMS are

part of this study.

The mean scjuare fluctuating response of a conductor is

the area under the response spectrum. The area under the

response spectrum can be considered as the summation of the

background response, B , and the resonant response, R . The

c c

background response is caused by gust with various

durations, whereas resonant response is caused by gust

frecjuencies at the natural frecjuencies of the conductor.

The total mean scjuare fluctuating response of the conductor

is given as:

^R = ^c ^ ^c

35

(2.23)

where B = the mean scjuare background response, and

R = the mean scjuare resonant response.

The expressions for background and resonant responses

consider wind properties such as mean wind speed, turbulence

intensity and gust spectrum and structural properties such

as frecjuency of vibration and damping. Ecjuations for

background and resonant responses of the conductor response

have been developed by Davenport (1980) and are given below.

The ecjuations are:

B_ = e P E^ c 1 + 0.81(-ii-l

^s

(2.24)

2 — 2 R = 6 P E^ c c 0.323A h. 1 , o ^,-(n-H) (2.25)

1 f7 = mean wind pressure; —pV C^ where P

e = the influence coefficient which translates

E =

the force to response; for conductor transverse

response is the product of L and d,

exposure factor at the effective height of

36

the conductor; which is twice the turbulence

intensity,

Lg = transverse scale of turbulence,

C^ = conductor force coefficient,

p = mass density of air,

C = conductor aerodynamic damping as a ratio

to the critical damping,

V = mean wind speed at the effective height

of the conductor,

f^ = fundamental frequency of the conductor

(horizontal sway),

A, n = Kaimal's gust spectrum constants,

c = narrow band correlation coefficient of

turbulence, with a typical value of 8,

L = effective conductor span,

d = conductor diameter, and

h = effective conductor height.

The background and resonant terms for conductor

response are calculated using ecjuations 2.24 and 2.25. The

mean scjuare value of fluctuating response is obtained using

ecjuation 2.23. Some of the assumptions made by Davenport in

deriving the simplified expressions for conductor response

are discussed by Mehta (Criswell, 1987).

37

Conductor Damping Ratio

The energy gained by the conductors from the

fluctuating wind is dissipated by the conductor damping. In

general, three sources of damping can be identified for

conductors, which are material damping, structural damping

and aerodynamic damping. Material damping is due to

internal energy dissipation by the material of the

conductor. Structural damping is due to friction,

impacting, and rubbing of any two surfaces of the

conductors. Both material and structural dampings are very

small for conductors as compared to aerodynamic damping.

Aerodynamic damping is due to the retarding force which is

developed from the relative motion between the conductor and

the air. In the analytical model the value of damping

ratio, C/ which is defined as the ratio of damping

coefficient to the critical damping coefficient, is

determined using a theoretical expression. This expression

is based on the inertial force principle as conductor

movement displaces an ecjual volume of air (Davenport, 1980).

C = 0.000048 (^ ) C^ (2.26) o

where V = mean wind speed,

f = fundamental frequency of the conductor.

38

C^ = conductor force coefficient, and

d = diameter of the conductor.

Calculation of the aerodynamic damping ratio using field

response data is evaluated in Chapter V.

Conductor Fundamental Frecjuency

Conductor frecjuencies of vibration are analytically

estimated by modelling conductor as a conductor oscillating

from side to side and using the principles of dynamic

ecjuilibrium (Symon, 1961). Fundamental transverse frequency

of the conductor, f , in Hz, for a parabolic profile can be

obtained using the following ecjuation:

JG f = ^ (2.27) o 32 S ^ '

where G = acceleration due to gravity, and

S = conductor sag.

Ecjuation 2.27 is used in Chapter IV to calculate the

conductor fundamental frecjuency.

CHAPTER III

FIELD DATA

Full-scale data used in this study were collected by

the Bonneville Power Administration (BPA). Since 1976, BPA

has conducted several projects to study wind load response

of transmission line systems by collecting and analyzing

wind and response related data on test lines in the field.

Transmission tower and conductor wind response data were

collected on an energized 500 kV single circuit transmission

line. An instrumentation system was used to measure wind

speed, wind direction, insulator swing, insulator load, and

tower member stresses.

During the period of December 1981 through May 1982,

BPA collected twenty-three separate recordings of wind and

the transmission line response with twelve-minute duration.

Dates, times, and mean wind speeds and direction of these

twenty-three records are shown in Table 3. These recordings

are used in the study presented here. Each record is

numbered by Nxx, where xx is the secjuence number of

occurrence of high winds. Data utilized in this study are

limited to wind and conductor response data. The site

characteristics, the instruments and the data accjuisition

39

40

systems used for the collection of wind and conductor

response data are described in this chapter.

TABLE 3

Field Data Records During 1981-1982

Record Number

NOl N02 N03

N04

N05 N06 N07

N08 N09 NIC

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

Date

12/02/81 12/05/81 12/15/81

12/16/81

12/16/81 01/14/82 01/16/82

01/31/82 02/03/82 02/14/82

02/15/82 02/15/82 02/16/82

03/08/82

03/11/82 03/12/82 04/12/82

04/13/82 04/17/82 04/20/82

04/20/82 04/28/82 05/07/82

Time

01.31.57 06.42.45 16.11.48

08.30.32

16.10.28 10.57.52 19.04.51

01.36.32 14.11.35 13.05.30

23.26.40 10.29.05 00.38.56

16.03.42

14.52.13 15.13.50 01.10.26

15.29.13 17.57.11 22.03.35

11.50.38 12.28.43 14.19.51

Mean Wind Direction* (azimuth)

227 209 240

93 260 215

263 53

228

227 220 237

259 220 235

246 276 90

274 280 278

Mean Wind Speed* (mps)

18.8 16.4 14.8

zero wind

15.8 15.5 22.3

21.4 9.6 18.5

21.0 14.0 20.3

zero wind

15.7 10.0 18.8

21.8 15.8 18.5

18.4 19.3 15.2

* measured at 34.7 m on Tower 16/4

41

Description of Test Site

The test site is located at Moro, Oregon, 56 kilometers

southeast of Dalles, Oregon (east of the Cascade Mountains).

The general topography in the vicinity of the test lines is

shown in Figure 7. As observed from the contours in the

figure, the deep Deschutes river canyon is just west of the

site. The test line is essentially located on the rim of

the canyon. To the east of the test lines is a flat terrain

which is uncultivated land containing grass with some

shrubs. The test lines are approximately 1500 m to the east

and 500 m above the elevation of the Deschutes river.

As indicated in Figure 7, the test site includes three

lines running almost north-south and approximately 8 degrees

west of true north (Kempner, 1981). The first 2 lines east

of the Deschutes river are 500 kV energized lines, called

John-Day Grizzly (JD-G) 1 and 2. A non-energized mechanical

test line (also referred to as the Moro test line) parallels

the other lines. JD-G lines 1 and 2 are 45.7 m apart,

whereas the distance between JD-G line 2 and the mechanical

test line is 38.1 m. Towers at the site (on JD-G line 2)

are numbered from 1 to 5, Tower 1 being the tower at the

northern end of the site. The instrumented tower is Tower

4, part of the John Day-Grizzly (line 2) system. It is

referred to as Tower JD-G 16/4 or simply Tower 16/4 because

it is located on mile 16 of JD-G line 2.

42

Figure 7: Topography of Site and Orientation of Power Lines

43

Tower 16/4 is a delta configuration lattice tower

structure as illustrated in Figure 8. The 33.4 m tall tower

supports three twin Chukar conductors (west, east, and

central) and two overhead groundwires (west and east). Each

Chukar conductor has an outer diameter of 40.7 mm and weighs

3.1 kg m~ . The Chukar conductors have 84 aluminum and 19

steel strands having ciiameter of 3.7 mm and 2.2 mm,

respectively. In order to reduce subconductor oscillation,

one conductor of each twin conductor is 229 mm lower than

the other.

The conductor span to the north of Tower 16/4 is 252 m

to a similar delta configuration suspension tower. The span

to the south is 450 m (refer to Figure 9) to a horizontal

configuration suspension tower. The change in tower

configuration causes both east and west twin conductors to

hang in a non-vertical (outward) position at Tower 16/4.

The west twin conductor is ecjuipped with dampers which make

it heavier than the east twin conductor.

Instrumentation

Three different types of instruments were used to

measure wind speed and wind direction. A climatronics mark

III model anemometer was located on top of Tower 3. Two

three-blade propeller-vane anemometers were mounted on Tower

44

West conductor

Anemomeler (34.7 m)

Load cell and Swing angle indicators

41 mm diameter

Figure 8: Schematic of Tower 16/4

N -^

45

effective half span lengths of conductors

Figure 9: Elevation Along the Test Line (Vertical Scale Exaggerated)

46

16/4, as shown in Figure 8. One was located on top of the

tower at a height of 34.7 m. The second unit was located on

the northwest tower leg at a height of 10 m and it projected

out 2.3 m north of the tower. Two four-blade propeller-vane

anemometers were installed on top of Towers 4 and 5. The

three-blade propeller-vane anemometer has a threshold speed

of 1.7 mps and a distance constant of approximately 4.6 m.

The threshold value is the stall speed of the unit. The

distance constant is the wind passage recjuired for 63%

recovery from a step change in wind speed. The wind

instruments ecjuipped with internal heaters to allow for

winter operation. Wind direction is indicated by azimuth

readings in degrees referenced to true north. The wind

direction is such that a zero degree reading corresponds to

true north and a clockwise rotation represents an increase

in the wind direction reading.

The load cells and swing angle indicators measured the

magnitude and direction of the conductor and overhead ground

wire loads that transfer to the tower structure. The

instruments were installed in the linkage between the

insulator string and the tower. All conductors and ground

wires were instrumented with one axial load cell and two

swing angle indicators. The swing angle indicators measured

longitudinal and transverse swings of the insulators.

47

Baldwin-Lima-Hamilton (BLH) strain gage load cells were

used to measure axial loads. BLH Type T3P1 load cells,

rated at 5000 pounds, were used for the overhead ground

wires and BLH Type T2P1 load cells, rated at 20,000 pounds,

were used for the conductors. Humphrey pendulum swing angle

indicators, model CP17-0601-1, were used to measure the

longitudinal (along the line) and transverse (perpendicular

to tihe line) swings of the insulator string. These units

measure up to ±45 degrees of swing from the zero (vertical)

position with a resolution of 0.2 degrees (Kempner, 1977).

Valid data from these units are restricted to 2 Hz or less

because of the unit natural frecjuency of 3.2 Hz, as

suggested by Kempner (1980). The load cells and swing angle

indicators were calibrated in the laboratory and checked in

the field after installation.

Data Acquisition

Data was collected by the Moro UHV mechanical test

program data accjuisition system (Kempner, 1979). The data

accjuisition system consisted of a PDP-11/10 mini-computer

with 12 K memory, an ADAC Model 600-11 Data Accjuisition

System, a Digi Data Controller/Formater, a 7-track magnetic

tape unit, and a teletype. The data accjuisition system was

housed in an instrument trailer located 30 m southwest of

48

the Tower 16/4. The system was set up to record from 256

channels of instrumentation.

RecordincT Procedure

Several selected channels constituted a recording mode,

which were selected to capture a static or dynamic

phenomenon of interest. The data used in the present study

were recorded in Mode 22. Mode 22 was designed for

recording wind and the response of Tower 16/4. It consisted

of 38 channels of instrumentation. The instrumentation on

Tower 16/4 included 12 strain gages, 5 load cells, 10 swing

angle indicators, and 2 wind instruments. The remaining

channels provide readings from the anemometers mounted on or

adjacent to Towers 2, 3, 4, and 5 on the Moro mechanical

test line. A complete list of the channels with a

description of each is given in Table 4 (Norville, 1985).

The PDP-11/10 computer was used to monitor wind

conditions and to initiate recordings when prescribed

conditions were met. Prescribed information was entered

into and stored in the computer prior to placing the data

accjuisition system on-line. This information included

channel identification, date and time of recording, number

of samples, mode number, calibration and offset factors for

each channel (Kempner, 1979). This information was written

TABLE 4

File Description of Mode 22

49

File

LCOl LC02 LC03 LC04 LC05 SAOl SA02 SA03 SA04 SA05 SAO 6 SA07 SA08 SA09 SAIO

SGOl SG02 SG03 SG04 SG05 SG06 SG07 SG08 SG09 SGIO SGll SG12

WDOl WD02 WD03 WD04 WD05 WSOl WS02 WS03 WS04 WS05 WS06

Channel

78* 79* 80 81 82 83* 84* 85* 86* 87 88 89 90 91 92

66* 67* 68* 69* 70* 71* 72* 73* 74*

. 75* 76* 77*

159 163 168 179 181 156* 158 161 167 178 180

Instrument,

Load Cell 1 Load Cell 2 Load Cell 3 Load Cell 4 Load Cell 5 Swing Angle Swing Angle Swing Angle Swing Angle Swing Angle Swing Angle Swing Angle Swing Angle Swing Angle Swing Angle

Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage Strain Gage

1 2 3 4 5 6 7 8 9 10

1 2 3 4 5 6 7 8 9 10 11 12

Wind Direction Wind Direction Wind Directi on Wind Direction Wind Direction Wind Speed Hot Wind Speed E Wind Speed E Wind Speed E

Location,

Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower

Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower Tower

Anem. Anem. Anem. Anem. Anem.

16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4

16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4 16/4

Height

East West West East Cent. East East West West West West East East Cent. Cent.

NW 1 NE 2 NW 2 SE 1 SE 2 SE 3 NW 2 SW 2 NW 2 NW 1 NW 2 NW 3

Tower 3 Tower 4 Tower 5

OHGW OHGW Cond. Cond. Cond. OHGW OHGW OHGW OHGW Cond. Cond. Cond. Cond. Cond. Cond.

Dia. Main Main Dia. Main Dia. Main Main Main Dia. Main Dia.

47.4 m 41.5 m 39.3 m

Tower 16/4 34.7 m Tower 16/4 10.0 m

Wire Anem. Near Tower 2 'rop. Anem. Tower 3 »rop. Anem. Tower 4 »rop. Anem. Tower 5

47.4 m 41.5 m 39.3 m

Wind Speed Prop. Anem. Tower 16/4 34.7 m Wind Speed E »rop. Anem. Tower 16/4 10.0 m

recordings of channels not used in this study

50

as a heading on the magnetic tape preceding a strong wind

recording.

Triggering of this recording mode was automatic when

the wind speed was equal to or greater than 18 mps for one

minute and the temperature was ecjual to or greater than 4

degrees Celsius. Once triggered, the recording mode sampled

the data for 10 or 12 continuous minutes, depending upon the

sampling rate. After a recording period, the mode had one

hour waiting period before another trigger was allowed.

Two data sampling rates were used in recording the data

for mode 22. The sampling and recording rates were limited

by the size of the computer memory and the speed of the

magnetic tape unit. Sampling rates of 10 samples per second

(sps) and 20 sps were used in collecting data. Channels 66

through 79 (LCOl, LC02 and SG01-SG12) were monitored at 20

sps; all other channels were sampled at 10 sps. All the

data utilized in the study presented here had sampling rates

of 10 sps and were collected for 12-minute durations.

Description of Recordings

The recordings that were collected in the field are

summarized in Table 3. The winds at the test site during

the period of data collection were predominantly from the

west, with only three records of east winds. Two zero wind

51

records were collected for initializing conductor and tower

response data. Mean wind speed values ranged from 9.6 to

22.3 mps. Mean wind directions varied from almost normal

(transverse) to transmission line to 55 degrees from the

normal. The terrain over which the wind traversed in each

recording segment depended on the wind direction. These

variations in wind speed, wind direction and terrain caused

inherent variability in the collected data. Field

experiments depend on the vagaries of nature; they cannot be

duplicated or repeated. The inherent variability in field

data and the inability to repeat the experiment suggest that

results from the analysis should be based on an ensemble of

data and that some scattering of results is to be expected.

CHAPTER IV

FIELD DATA ANALYSIS

Analysis of field data recjuires that the validity and

accuracy of the wind and conductor response data be checked.

It is expected that results of field data will have a

certain amount of scattering. This scattering can be due to

an inherent variation in field data as well as a variation

in the data measuring system. To the extent possible, wind

data and conductor response data are checked for consistency

and accuracy.

The analytical procedure to predict the response of

conductors recjuires knowledge of wind characteristics such

as the mean wind speed, power-law exponent of wind profile,

turbulence intensity, and Kaimal's gust spectrum constants.

These characteristics are determined from the field data and

are subsecjuently used in the analytical procedure.

Field recorded conductor response can be compared with

predicted values in terms of mean response and fluctuating

response. Effective force coefficients that relate to mean

response are determined for each record of the field data.

Fluctuating response, as indicated in Chapter II, is a

combination of background response and resonant response.

52

53

These responses can be assessed from response spectra of the

field measured values. The conductor response spectrum of

transverse loads recorded by load cells is discussed in this

chapter. The discussion relates to background response and

resonant response. Comparisons of recorded fluctuating

responses with predicted values from the analytical

procedure are presented in the next chapter. In addition,

an assessments of the peak factor, admittance function and

damping from the recorded data are presented in next

chapter.

Validity of Wind Data

Wind speeds and directions were recorded at five

locations, namely on top of Towers 3, 4, 5, and 16/4, and at

10.0 m on Tower 16/4. Tower 4 is within 38 m of Tower 16/4,

while Towers 3 and 5 are several hundred meters from the

other two towers (see Figure 9).

The mean azimuth and root mean scjuare (RMS) of wind

direction for each record for these five locations are shown

in Table 5. Data were recorded at 10 sps for 12 minutes in

each record. One observation in Table 5 is the consistency

in mean azimuth value for the three instruments located at

the tops of Towers 4, 5, and 16/4. The differences in mean

values for these three instruments are less than 12 degrees.

54

and in most cases less than 6 degrees. The two instruments,

one at 34.7 m height on Tower 16/4 and one at 41.5 m on top

of Tower 4, give mean wind directions within 4 degrees for

winds from several different azimuths. This consistency in

mean azimuth values provides credibility to the wind

direction instruments and the data accjuisition system.

The mean azimuth values for the instrument at 10.0 m

height on Tower 16/4 are erratic when compared with the

values from the other instruments for records NOl through

N13 (see Table 5). The difference in mean azimuth value is

as high as 44 degrees. Such a large difference in wind

direction between records for instruments on the same tower

located at 10.0 m and at 34.7 m cannot be explained from

physical phenomena. Ekman's spiral suggests that a

deviation in wind direction occurs close to ground, but the

deviation in a 25 m difference in height should be less than

6 degrees (Simiu, 1985). This large difference in wind

direction for the instrument at 10.0 m on Tower 16/4 casts

doubt in its wind data for records NOl through N13; these

records are not used in further analysis. The remaining

records (N15 through N23) for this instrument appear to give

reasonable results. The wind instrument on top of Tower 3

also gives mean azimuth values higher by 10 to 20 degrees

than the other instruments on a consistent basis. Wind data

from Tower 3 are not needed in the analysis.

TABLE 5

Mean Azimuth and RMS in Degrees of Wind

55

Instrument WDOl Location Height

NOl* N02 N03

N04

N05 N06 N07

NOB N09 NIC

Nil N12 N13

N14

N15 N16 N17

Twr 3 47.4 m

Mean RMS

197 216 246

Zero

101 * *

•k-k

276 64

243

241 234 250

Refe

274 232 246

22 7 4

Wind

2 * *

*-k

6 9 7

7 9 8

rence

5 9 5

WD02 Twr 41.5 Mean

224 205 239

4 m RMS

9 12 5

record

94 259 211

263 54

225

225 218 235

Zero

260 218 231

2 7 9

8 7 8

7 9 9

Wind

7 8 8

WD03 Twr 39.3

Mean

229 207 246

94 266 219

268 56

228

230 222 236

5 m RMS

8 12 6

2 8 10

6 8 7

9 12 8

WD04 Twr 16/4 34.7 Mean

227 209 240

93 260 215

263 53

228

227 220 237

Speed Record

267 228 235

8 8 8

259 220 235

m RMS

9 12 5

2 7 9

8 8 8

7 10 9

8 9 8

WD05 Twr 10.

Mear

212 199 224

81 300 258

300 97

268

266 258 278

267 230 244

16/4 0 m L RMS

9 12 5

4 7 9

9 9 9

8 10 9

8 9 8

N18 N19 N20

N21 N22 N23

259 292 100

285 295 296

5 12 1

13 9 9

244 277 87

273 279 279

5 12 1

10 11 12

256 276 90

283 283 282

10 15 1

10 13 14

246 276 90

274 280 278

5 12 1

11 10 12

252 283 101

280 286 283

7 11 2

12 10 12

* Record Number ** error in data

56

The mean wind speed value for each record for the five

wind speed instruments is tabulated in Table 6. One

observation in Table 6 is the consistency in mean wind speed

value for the four instruments located on the tops of Towers

3, 4, 5, and 16/4. The values among these instruments are

within 15%. The variation in mean speed values may be due

to a combination of two reasons. First, the instruments are

at slightly different heights above the ground; wind speed

is expected to be higher as height increases. The second

reason may be that the distances between the towers are

several hundred meters; so the terrain over which the wind

travels can be different. This combination of variation in

terrain and height of instrument can account for variations

in mean wind speed values. The instrument at 41.5 m height

on Tower 4 recorded 4% higher mean wind speed than that at

34.7 m height on Tower 16/4 on an average (see Table 6).

Since Towers 4 and 16/4 are only 38 m apart with no terrain

modulations, it is reasonable to assume that a height

difference accounts for the small difference in mean wind

speed values. This consistency in mean speed values

provides credibility to the wind speed instruments and the

measurements.

The difference between the mean wind speeds at 10.0 m

and at 34.7 m on Tower 16/4 is very small for records NOl

TABLE 6

Mean Wind Speed in mps

57

Instrument WS02 Location Twr 3 Height 47.4 m

WS03 Twr 4 41.5 m

WS04 WS05 Twr 5 Twr 16/4 39.3 m 34.7 m

WS06 Twr 16/4 10.0 m

NOl* N02 N03

N04

N05 NO 6 N07

NOB N09 NIC

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

18.7 18.9 15.2

19.4 16.7 16.0

19.8 16.4 14.7

18.8 16.4 14.8

Zero Wind record

15.0 * *

* *

21.8 9.6 19.0

21.2 13.5 20.2

15.3 15.4 21.8

21.5 9.8 18.8

21.4 13.9 20.7

15.3 14.9 19.1

22.6 9.7 17.3

19.5 13.0 21.0

Reference Zero Wind Speed Record

16.4 9.0

20.4

22.1 16.1 18.3

15.4 9.9 19.1

22 15 18

1 5 4

15.5 8.6 17.3

17.4 16.2 17.7

15.7 10.0 18.8

N21 N22 N23

16, 19, 16.

0 1 6

18.4 19.4 15.2

18 18 16

0 0 6

18.5? 17.5? 14.8?

15.8 15.5 22.3

21.4 9.6 18.5

21.0 14.0 20.3

15.6? 15.3? 23.3?

20.4? 8.8? 18.3?

21.0? 14.3? 20.0?

12.1 8.0 15.5

21.8 15.8 18.5

18.4 19.3 15.2

18.2 13.0 14.4

15.5 16.5 13.2

* Record Number ** error in data ? these values are cjuestionable along with wind direction; they are not used in analysis

58

through N13. In a few records, the instrument at 10.0 m

height shows higher mean wind speed values than that at 34.7

m height. Wind speed at 34.7 m on Tower 16/4 is more

consistent with the other instruments on top of Towers 3, 4,

and 5. This casts some doubt on the wind speeds of records

NOl through N13 measured at 10.0 m height on Tower 16/4;

these records are not used for further analysis.

To further validate the field data, time histories are

plotted for all wind speed records. A typical time history

plot for wind speed recorded at 34.7 m on Tower 16/4 for

record NOl is shown in Figure 10. Each record has 7200

points, which corresponds to a recording rate of 10 sps for

the duration of 12 minutes. The time history plot in Figure

10 uses averages of 10 consecutive points; thus the points

are at one second intervals. Visual observation of time

history plots of wind speed records for any discontinuities,

trends, and noise shows the records to be good.

An additional check of stationarity was performed to

assess consistency of the wind data. Stationarity checks

are done to verify that the statistical properties are time

invariant. Stationary tests were accomplished by Levitan

(1987) as part of an earlier BPA contract. Most of the wind

records were found to be stationary with 95% confidence

limits as checked by the run and trend tests suggested by

59

Time (minutes)

Figure 10: Time History Plot of Wind Speed for Record NOl at 34.7 m on Tower 16/4

Bendat and Piersol (1966). These stationary test are not

discussed in this study. Finally, as a part of the

validation of the data, a power spectrum plot for each

record was generated. These plots would reveal, in the

frecjuency domain, any electrical or other sources of noise

in the data. Details of these power spectrum plots are

discussed in subsecjuent sections. In general, most of the

wind speed and wind direction records appear to be valid.

60

The mean wind speed and mean wind direction for each

record at 34.7 m height on Tower 16/4 are shown in Figure

11. There are eighteen west wind records, three east wind

records and two zero wind records. Wind directions vary

from almost normal to the transmission line to as high as 55

degrees of yaw angle (see Figure 11). The mean wind speed

values are 10 mps or higher for all records, except for the

zero wind records.

Two zero wind records (N04 and N14) were collected in

near calm or zero wind conditions. Based on a review of the

data of these two records, N14 is selected as a reference

record to assess conductor response.

Wind speed on top of Tower 16/4 (at 34.7 m height) is

used as the reference wind speed record. The reference wind

speed record is important to determine the wind speed at the

effective height of the conductor. The wind speeds at the

effective heights are determined using the reference wind

speed and vertical wind profile.

Power-Law Exponent for Wind Profile

The wind speed profile is assessed using mean wind

speed records of instruments located at 10.0 m and 34.7 m

heights on Tower 16/4 and at 41.5 m height on Tower 4.

Since Towers 16/4 and 4 are only 38 m apart with no terrain

6 1

• Record Number N_ • Mean Wind Speed and Direction

Figure 11: Mean Wind Speed and Direction Recorded at 34.7 m on Tower 16/4 of 23 Records

62

modulations between the towers, it is reasonable to assume

that the wind environment is the same at both the towers.

The other wind speed instruments are located far away; it

would be inappropriate to use their data to assess the wind

speed profile. As mentioned earlier, wind speed records NOl

through N13 at 10.0 m height on Tower 16/4 are cjuestionable.

Hence the wind speed profile is assessed using records N15

through N23 only. The power-law expression given in

ecjuation 2.20 can be written as

V^ Z, ln(-f.) = a ln(-^) (4.1)

Vl

where V, = wind speed at height Z^

V2 = wind speed at height Z2 and

a = power-law exponent.

Ratios of wind speeds and ratios of heights for record

N15 are plotted in Figure 12. Three data points in the

figure represent data collected at heights of 10.0 m, 34.7

m, and 41.5 m. A straight line using the least squares

method is fitted to the three data points. The slope of the

straight line is the value of the power-law exponent.

Power-law exponent values obtained for records N15

through N23 are given in Table 7. The flat land on the east

63

CVi >

ln(Z2/Zi)

Figure 12: Power-Law Plot for Record N15

and the valley on the west give similar power-law exponent

values for the limited data collected in this phase of the

project. These values are consistent with results reported

for this site by Kempner (1982).

Calculated power-law exponent values range between 0.11

and 0.18; this wide range corresponds to open farmland and

64

TABLE 7

Power-Law Exponent and Kaimal's Gust Spectrum Constants

Record Number

NOl N02 N03

N04

N05 NO 6 NO 7

N08 N09 NIO

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

Power 1 aw exponent

_

-

-

Zero Wind

_

-

-

.

-.

-

^

-

-

Reference

0.18 0.16 0.15

0.14 0.13 0.18

0.13 0.12 0.11

record

Zero Wind

A*

0.327 0.591 0.269

0.009 0.133 0.254

0.151 0.182 0.242

0.209 0.294 0.296

Speed Record

0.209 0.157 0.267

0.151 0.304 0.005

0.152 0.265 0.342

n**

1.355 1.721 1.571

1.686 1.494 1.463

• 1.579 1.925 1.786

1.200 1.696 1.420

1.516 2.052 1.573

1.383 1.681 1.565

1.375 1.628 1.695

* constant representing amplitude ** constant representing exponent

Note: Kaimal's gust spectrum constants are calculated for wind speed recorded at 34.7 m on Tower 16/4.

65

suburban terrains (see Table 1 in Chapter II). An average

power-law exponent value of 0.14 is used to determine the

wind speed at the effective height of the conductors (points

of effective pressures). A small change in exponent value

does not drastically change the mean wind speeds at the

effective heights in the present study.

Turbulence Intensity

Turbulence intensity represents the level of turbulence

present in the wind. It was defined in Chapter II as a

ratio of RMS to mean wind speed. Values of turbulence

intensity calculated using data from all five instruments,

are shown in Table 8 for all records considered.

The turbulence intensity values among the instruments

for each record are fairly consistent, though there is some

scatter. The instruments located on top of Towers 4 and

16/4, which are at similar heights and are 38 m apart, give

turbulence intensity values almost the same in most of the

records. Seventeen records for these two instruments give

turbulence intensity values within 0.02; only four records

show a higher difference. Turbulence intensity values for

the instrument located at 10.0 m on Tower 16/4 are

consistently higher than those at 34.7 m on Tower 16/4.

Winds near the ground are expected to contain more

TABLE 8

Turbulence Intensity

66

Instrument Location Height

WS02 Twr 3 47.4 m

WS03 Twr 4 41.5 m

WS04 Twr 5 39.3 m

WS05 Twr 16/4 34.7 m

WS06 Twr 16/4 10.0 m

NOl* N02 N03

0.14 0.13 0.08

0.16 0.22 0.07

0.15 0.24 0.11

0.18 0.17 0.11

N04 Zero Wind record

N05 NO 6 N07

NOB N09 NIO

Nil N12 N13

0.03 * *

* *

0.11 0.10 0.11

0.12 0.18 0.12

0.03 0.11 0.16

0.13 0.09 0.15

0.14 0.16 0.13

0.03 0.12 0.22

0.11 0.09 0.15

0.13 0.22 0.14

0.03 0.11 0.16

0.14 0.13 0.17

0.18 0.17 0.15

* * *

* * *

* * *

* * *

* * *

* * *

* * *

* * *

* * *

N14 Reference Zero Wind Speed Record

N15 N16 N17

N18 N19 N20

N21 N22 N23

0.10 0.12 0.12

0.09 0.17 0.02

0.18 0.17 0.18

0.13 0.11 0.14

0.08 0.17 0.02

0.13 0.15 0.21

0.14 0.14 0.16

0.16 0.17 0.02

0.15 0.13 0.16

0.12 0.12 0.14

0.11 0.17 0.02

0.11 0.15 0.21

0.14 0.18 0.14

0.13 0.21 0.04

0.17 0.19 0.23

* Record Number ** error in data *** cjuestionable da ta ; see Tables 5 and 6

67

turbulence than the winds at higher elevation. Therefore,

the fluctuating wind data can be considered valid.

The turbulence intensities of records for the

instrument at 34.7 m on Tower 16/4 (reference wind

instrument) vary between 0.02 and 0.21. The turbulence

intensities for winds from west and southwest show cjuite a

bit of scatter; they range between 0.11 and 0.21 for the

reference wind instrument. The west winds traverse over

hills and valleys (see Figure 7) before approaching the

transmission line. The variations in turbulence intensity

are random. Little correlation between turbulence intensity

and mean wind speed and direction (for west winds) was

found. Turbulence intensities of winds from the east, where

the winds traverse flat terrain (see Figure 7), are low.

Two east wind records N05 and N20 show very low turbulence

intensities, 0.03 and 0.02, respectively, compared to the

third east wind record N09 (turbulence intensity of 0.13).

The turbulence intensity can be related to a terrain

classification. In general, low values of turbulence

intensity are related to flat terrain and high values to

rough terrain. The west wind records, show a wide variation

in turbulence intensity. The turbulence intensity values

range from terrain classification of open farmland to

forest/suburban terrain (refer to Table 2). This suggests

68

that a terrain of canyons and hills is unpredictable for

characterization of wind. The power-law exponent values for

the profile, as well as turbulence intensity values, vary

from flat to suburban terrain, but there is no trend between

the two.

Turbulence intensity is one of the major parameters in

the analytical procedure to predict response. The exposure

factor in ecjuations 2.24 and 2.25 is twice the turbulence

intensity. A parametric study of the analytical procedure

conducted by GAI consultants (1981) and by Twu (1983) showed

that the turbulence intensity is the most significant wind

characteristic in predicting the response. Turbulence

intensity values recorded at 34.7 m on Tower 16/4 (reference

wind instrument) are used in predicting response of the

conductors in the next chapter.

Kaimal's Gust Spectrum Constants

Gust spectra were plotted for all wind speed records

collected at 34.7 m on Tower 16/4 (reference wind

instrument). The gust spectrum for record NOl is shown in

Figure 13. Gust spectra were obtained utilizing the

International Mathematics and Statistical Libraries (IMSL,

1982) program FTFREQ. The program uses an autocorrelation

function to calculate the spectral density function

69

estimates, S(f). Calculation of spectral density function

involve standard expansions (Jenkins, 1968) not involving

fast Fourier transformation. The highest frecjuency obtained

is the Nycjuist frecjuency, 5 Hz, which is one half the

sampling rate of 10 sps. The lowest frecjuency is restricted

by the number of lags selected for correlation. Here the

lowest frecjuency is 0.0025 Hz because of the selection of

2000 lags (200 seconds) from autocorrelation plot. The

program applied the Hamming window for smoothing. The plot

2 in Figure 13 uses (f S(f)/a ) on the ordinate to give a

2 normalized linear scale, where a is the mean scjuare of the

time series, and f is frecjuency in Hz. The abscissa in

Figure 13 is the log of the frecjuency, f.

The gust spectrum shown in Figure 13 is typical of the

gust spectra of other wind speed records. Fluctuations in

the spectral density values are partly due to statistical

methods employed in the program. A general trend of

reduction in ordinate (spectral energy) with increase in

frecjuency is noticeable in the figure. The ordinate becomes

negligible for frecjuencies greater than 1 Hz. Calculation

of the area under the gust spectrum (which is ecjual to the

mean scjuare value) indicates that only 1% of the spectral

energy is in the frecjuency range above 1 Hz. The gust

spectrum in the figure shows a reasonable amount of spectral

70

o 00

O C/1

CO • U.

u 3 rt

> ^^ C3 ^ o 1) Q .

CO

O P^

o

o CO

o

o in

a

LJ 3 *

. o

o en o

o CM

, o

a

o

o o

C . O O l

S (f) - spectral density value at f f - frequency SIGSQ - mean square value

Frequency (Hertz)

Figure 13: Gust Spectrum Plot for Record NOl Recorded at 34.7 m on Tower 16/4

energy in the frequency range of 0.1 to 0.4 Hz; natural

frecjuencies of the conductors are in this range.

The analytical procedure to predict response of the

conductor developed by Davenport (1980) utilizes an

analytical form of the gust spectrum. The specific

71

analytical form used in the procedure is the one proposed by

Kaimal, as described in Chapter II (ecjuation 2.22). Gust

spectra obtained from the field data are used to determine

Kaimal's gust spectrum constants A and n. Constants A and n

represents the amplitude and slope of the analytical gust

spectrum.

The curvilinear regression procedure (Miller, 1977) is

used to obtain suitable values for constants A and n from

the field wind data. Ecjuation 2.22 is rewritten to separate

dependent and independent variables:

A u^ f h '^. (4.2)

Ecjuation 4.2 is nonlinear. It is made linear by taking

natural logarithms on both sides.

In S^(f) = ln(A nl)+n ln(^)-(n+l) ln(f). (4.3)

Ecjuation 4.3 can be written in the form

Y = C - (n+1) X (4.4)

where

C = ln(Au2) ^ ^ ln(-^), and

X = ln(f).

72

In equat ion 4 .4 , Y i s dependent on X. To solve for A

and n, normal ecjuations of t h i s ecjuation are u t i l i z e d

(Mi l l e r ,1977) . The normal equat ions a re ,

^Y = N C - (n+1) j ; x (4.5)

^X Y = C (£X) - (n+1) p 2 ^4 gj

where N = total number of data points.

Values of n and C are found by solving ecjuations 4.5

and 4.6. Substitution of n and C in ecjuation 4.3 gives the

value for A for each record. Table 7 shows the calculated A

and n values considering gust spectrum values between the

frecjuencies of (—'—- ) and 1 Hz. As indicated in Chapter h

I I , ecjuation 2.22 i s v a l i d for f g rea t e r than ( f^ ^ ) . The

h

gust spectrum values for frecjuencies greater than 1 Hz are

neglected because spectral energy (the ordinate of the gust

spectrum) is very small. The value of cionstant A is least

affected by the selection of maximum frecjuency range,

specially beyond 1 Hz. Values of constant A range between

0.005 and 0.591 (refer to Table 7). The gust spectrum (see

Figure 13) shows that the slope of the spectrum curve decreases between 0.1 and 1 Hz, and remains almost constant

beyond 1 Hz. Hence, n values are sensitive to the maximum

73

frequency range selected. Values of n range between 1.200

and 2.052.

The suggested range of values for A and n are 0.15-0.60

and 0.33-0.67, respectively (GAI, 1981). Field measured

values of A are generally within the suggested range.

However, field measured n values are much higher than the

suggested range. The parametric study of the Davenport

model for gust response factor of transmission line

structures conducted by (Twu, 1983) showed that the response

is not sensitive to the value of n.

Kaimal's analytical form of the gust spectrum (ecjuation

2.22) is valid for values of f greater than (-^LLL_Y.). For

h

the field measured data obtained at 34.7 m, this low end of

frecjuency f would be 0.3 Hz, if the mean wind speed were 20

mps. Since the conductor natural frecjuencies are in the

range of 0.1 to 0.4 Hz, near or below the low end of this

frecjuency range, Kaimal's gust spectrum constants obtained

from the field data are not appropriate for use in obtaining

response. For this reason, values of A and n of 0.3 and

0.67, respectively, as suggested by Davenport, are used to

predict the response utilizing the analytical procedure in

the next chapter.

74

Validity of Conductor Response Data

The conductor response data comprise measurements from

the transverse and longitudinal swing angle indicators and

load cell transducers. These instruments were placed at the

attachment of the insulator to the Tower 16/4 (see Figure

8). The conductors are suspended at the bottoms of the

insulators. The load cells measure axial load in the

insulator and the swing angle indicators measure swing of

the insulator from the vertical. Appropriate combinations

of values of load cells and swing angle indicators provide

vertical, longitudinal and transverse components of loads

applied on the insulators by the conductors. Interest in

this study is restricted to transverse load components since

the primary influence of wind is in the transverse

direction. The transverse load component at any instant is

calculated by the following ecjuation:

F = P cos (p sin G (4.7)

where P = axial load measured by the load cell,

(p = longitudinal swing angle, and

9 = transverse swing angle.

Twenty-one records of response for each west, east, and

central twin conductor are available. Record N14 is used as

a reference zero wind speed record to initialize all

75

records. Static loads are transferred from the conductors

to the tower structure even before being subjected to wind

loads. These initial static loads are due to the self

weight of the conductors and the inclined orientation of the

insulators. The inclined orientation of the insulators is

caused by the differences in elevation of the supporting

tower structures and by the different configuration of the

tower (refer to Chapter III). Load cell and swing angle

readings of record N14 give the weight of the conductors and

initial transverse loads. These initial transverse loads

are subtracted from the transverse loads obtained in each

record. Initialization with the zero wind speed record is

necessary to obtain conductor response due to wind only.

A time history plot of the response of the west

conductor for record NOl is shown in Figure 14. Similar to

wind records, each response record has 7200 points, which

corresponds to data collection at the rate of 10 sps for a

time duration of 12 minutes. The time history plot in

Figure 14 uses averages of 10 consecutive points; thus the

points are at one second intervals. Mean, standard

deviation and one second interval peak values are shown on

the time history plot. The purpose of these plots is to

provide a graphical display of conductor response versus

time, thereby illustrating the overall quality and trends in

the data.

76

- Imer\-al Peak - 4.61

Time (minutes)

-, Mean+Sigma » 3.43

Mean - 2.97

Mean-Sigma = 2.51

Figure 14: Time History Plot of West Conductor Response for Record NOl

Similar to the wind data, stationarity checks are done

for the load cell and conductor swing angle data. Most of

the conductor response data (load cell and swing angle data)

are found to be stationary with 95% confidence limits as

checked by the run and trend tests (Bendat, 1966). The

response spectra are plotted (in subsecjuent section) to

check for noise in the data. In general, the conductor

response data are valid and can be used for analysis.

77

The mean and RMS values of response for the three

conductors (west, east, and central) are shown in Table 9.

The negative values in Table 9 indicate wind from the east.

The choice of sign for conductor response is arbitrary. The

values in the table show that response recorded at the

central conductor is higher than that for the west and east

conductors. This is expected since the central conductor is

suspended at 26.7 m above ground level, which is 8.4 m above

the west and east conductors at Tower 16/4 (see Figure 8).

The response of the west conductor is slightly higher

than that of the east conductor in many records, even though

they are suspended at the same height above ground. The

range of the difference in values is from -0.08 to 0.33 kN;

less than 10% of the recorded values. There could be one of

several reasons or a combination of reasons for this

difference. These reasons and observations based on the

data are given below.

One set of twin conductors (west or east) could shield

or intensify wind load effect on the other set of twin

conductors. This is not likely because the west and east

conductors are 13.4 m apart at Tower 16/4. In addition,

data do not show a specific trend in response for winds

coming from east or west.

78

TABLE 9

Mean and RMS Values of Conductor Response (Transverse Load Component)

Record Number

NOl N02 N03

N04

N05 NO 6 N07

NOB N09 NIO

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

West Conductor Mean RMS

(kN)

2.97 1.36 2.48

Zero

-2.84* 2.25 2.82

5.14 -1.10 2.81

3.35 1.27 3.70

0.46 0.43 0.21

Wind reco

0.13 0.20 0.89

0.76 0.10 0.52

0.61 0.41 0.66

Reference Zero

2.61 0.72 3.32

4.67 2.48 -2.74

2.99 3.09 2.39

0.31 0.12 0.61

0.78 0.52 0.12

0.51 0.60 0.50

East Conductor Mean RMS

(kN)

2.82 1.34 2.43

rd

-2.62 2.19 2.80

4.83 -1.13 2.76

3.28 1.34 3.51

0.41 0.40 0.19

0.09 0.18 0.82

0.76 0.11 0.46

0.56 0.41 0.59

Wind Speed Recc

2.47 0.80 3.10

4.27 2.33 -2.78

2.79 2.90 2.20

0.28 0.14 0.57

0.72 0.17 0.02

0.44 0.53 0.43

Central Conductor Mean

(kN)

3.13 1.34 2.56

-2.77 2.32 2.94

5.47 -1.16 2.91

3.53 1.35 3.85

>rd

2.68 0.81 3.34

4.94 2.58 -3.00

3.12 3.22 2.44

RMS

0.53 0.43 0.20

0.10 0.22 1.00

0.79 0.14 0.54

0.66 0.39 0.69

0.34 0.12 0.67

0.80 0.55 0.14

0.57 0.65 0.52

* negative sign indicates swinging to the west

79

Measurement of weight of the conductor during reference

zero wind condition can cause discrepancy in calculation of

transverse loads. The west conductor weight is recorded

higher than the east conductor by 2.4 kN (a difference of

9%). The information obtained from the field indicates that

detuner dampers are added to the west conductors. These

dampers account for approximately 0.9 kN of extra weight.

Discrepancy in dead weight of 1.5 kN in west conductor is

not reconciled. This extra dead weight could be a reason

for higher response of west conductor compared to east

conductor in most of the records (see Table 9).

Accuracy in swing angle measurements is important for

estimation of the transverse load component. The swing

angle indicators have a resolution of 0.2 degrees.

Discrepancy in swing angle reading of 0.2 degree could cause

error in response by 2-4% for the range of readings obtained

in these data. Accuracy of the swing angle indicator can be

one of the factors causing a difference in recorded

response. Notwithstanding the differences in recorded

responses of conductors mentioned above, the data for the

three conductors are fairly consistent. Some variations and

scattering in the field data are inherent; they cannot be

avoided. In general, conductor response data are considered

valid for use in analysis.

80

Effective Conductor Force Coefficient

As discussed in Chapter II, the mean response of the

conductor is related to a nondimensional force coefficient.

To calculate the force coefficient based on field measured

data, it is necessary to use measured values of the mean

wind speed, V, and the mean transverse load component, F.

The field measured effective conductor drag coefficient is

obtained by ecjuating the measured mean transverse load

component to the 'stagnation pressure' load.

4 p V L d 2

where C^ = conductor effective force coefficient,

F = mean transverse load measured in kN,

-3 p = mass density of air in kN m

( = 12.02x10"^ kN m"^)

V = mean wind speed at conductor effective

height in mps,

L = effective conductor span in meters, and

d = conductor diameter in meters

(= 2 times the diameter; for present study)

81

The diameter is multiplied by 2 because of twin

conductors in each conductor bundle. The effective span of

the conductor depends on the attachment heights, and the

conductor span. In addition, temperature, and conductor

tension have some effect on the effective span. As

indicated in Figure 9, effective spans are determined to be

376 m for west and east conductors, and 402 m for the

central conductor. For determination of effective spans,

the horizontal tension in the conductor is taken to be 45 kN

at 60 F, based on data obtained from the BPA.

The conductor effective force coefficient is a function

of wind direction (Potter, 1981). Computing the transverse

wind speed component from the wind speed and yaw angle to

relate to the transverse load component, the C^ values are

believed to be realistic. Wind tunnel tests show that the

force coefficient essentially remains the same for a yaw

angle up to 22 degrees (Potter, 1981). Therefore the

effective force coefficients computed from the data are

limited to records with yaw angles of 22 degrees or less.

Calculated effective conductor force coefficient values for

eleven records are shown in Table 10.

The effective force coefficient values are scattered

between 0.47 and 0.74. The values for the west conductor

are higher than those for the east conductor because the

82

TABLE 10

Field Measured Conductor Effective Force Coefficients

Record Yaw West East Central Number Angle Conductor Conductor Conductor

(degrees)

NOl 35 -N02 53 -N03 22 0.74 0.72 0.63

N04 Zero Wind record

N05 NO 6 N07

NOB N09 NIO

Nil N12 N13

11 2 47

1 29 34

35 42 25

0 0

0

.74

.61 -

.73 -

-

-

-

0.68 0.60

-

0.68 -

-

-

-

0.60 0.53

-

0.64 -

-

-

-

N14 Reference Zero Wind Speed Record

N15 3 0.69 0.65 0.59 N16 42 -N17 27 -

N18 16 0.64 0.58 0.57 N19 14 0.65 0.61 0.56 N20 8 0.52 0.53 0.48

N21 12 0.57 N22 18 0.54 N23 16 0.68

0.53 0.51 0.63

0.50 0.47 0.58

83

mean response of the west conductor is larger than that for

the east conductor as discussed in the previous section (see

Table 9).

There are three possible reasons for the scatter in the

values of effective force coefficient. The first reason may

be the accuracy with which transverse loads are measured.

Field measured transverse load components could vary by

about 10%, (see the previous section on validity of

conductor response data). The second reason may be the

modification of the wind speed for different heights above

the ground. In calculating the effective height of each

conductor, the conductor shape is assumed to be parabolic

and the site topography is taken into account. More

specifically, each effective height is calculated as the

average of the vertical distances between the ground and the

conductor at all points along the half spans on both sides.

The resulting values are 15.7 m for the east and west

conductors, and 23.8 m for the central conductor. The wind

profile is taken as having the same properties with respect

to the ground no matter how the ground elevation varies.

This raises some uncertainties, especially with regard to

the valley shown between Towers 16/4 and 16/5 in Figure 9.

Wind speeds are modified from the 34.7 m height to the 23.8

m effective height for the central conductor and to 15.7 m

84

effective height for west and east conductors using a wind

profile exponent value of a = 0.14. The modification in

wind speed is smaller for the central conductor than for the

west and east conductors. This could be the reason for the

force coefficients for the central conductor being smaller

than for the west and east conductors (refer to Table 10).

The third reason may be the wind characteristics of

turbulence and fluctuations in wind direction. As shown in

Table 5, many records show RMS of wind direction

fluctuations to be in the neighborhood of ten degrees.

These data in Table 5 suggest that the wind direction may

have fluctuated as much as 40 degrees within a record. In

addition, the turbulence intensity values shown in Table 8

are different between the records. Wind characteristics may

have significant effects on effective force coefficients.

The values of effective force coefficient obtained from

the field data are fairly consistent with the values shown

in Figure 3. The Reynolds Number for the field data is in

4 4 the range of 3x10 to 6x10 . For this range of Reynolds Number, Figure 3 suggest effective force coefficients in the

neighborhood of 0.6.

85

Response Spectrum

The conductor response spectrum represents the

fluctuating response about the mean response in the

frecjuency domain. Response spectra of west, east, and

central conductors for all records were obtained. A typical

response spectrum for the west conductor for record NOl is

shown in Figure 15. The response spectra were obtained

using the same procedure as for the gust spectra discussed

earlier. Similar to the gust spectra, the lower and upper

limits of the frecjuency range in spectral calculation are

0.0025 and 5 Hz, respectively. The response spectrum in

Figure 15 shows fluctuations in frecjuencies below 1 Hz, but

does not show a spike at a specific frecjuency. The range of

frecjuencies of interest for the present study is based on

natural frecjuencies of conductor vibration. The conductor's

natural frecjuencies are in the range of 0.1-0.4 Hz,

depending on its configuration. At frecjuencies between 2

and 4 Hz, the spectrum shows several very high peaks. These

peaks in the figure are somewhat misleading because the

spectral density function in the plot is multiplied by the

frecjuency. Generally, the spectral density function is

multiplied by frecjuency to get enhanced results in the

higher frecjuency ranges.

86

a CO O teiH

CO " » O >»,' CO

« V M

1

O 3 •3 >

1 ii CU

CO

Q 3 *

a

n CD

• a

o CM

• O

o .—1

f

o

a o

G.OOL

f - Frequency (in Hertz) S(f) - Spectral Density at f SIGSQ - Mean Sqaure Value

Zone Zone

^^Vr-n.. O . l

Frequenc:y, f (Hertz)

l . O

Zone III

Figure 15: Response Spectrum Plot for West Conductor Response for Record NOl

As noted in Chapter II, the area under the response

spectrum gives the mean scjuare of response. The response

spectrum is viewed in terms of area representing extent of

response. Peaks in the spectrum represent response at

specific frecjuencies. In order to discuss background and

resonant response, the spectrum is divided into three zones

for determination of relative response: (1) frequencies

87

less than 0.1 Hz, (2) frequencies between 0.1 Hz and 1.0 Hz,

and (3) frequencies greater than 1 Hz. Each zone of the

spectrum is reviewed in light of the conductor response to

wind.

The spectral area in the zone of frecjuencies less than

0.1 Hz is in the neighborhood of 75% for all records. This

response is primarily due to the background turbulence where

wind has significant turbulent energy (see Figure 13). The

peaks in the spectra in this zone do not represent dynamic

amplification of the response.

The spectral area in the zone between frecjuencies of

0.1 and 1.0 Hz is close to 15% for all records. Fundamental

transverse frecjuency of the conductor, f , in Hz for a

parabolic profile can be obtained using ecjuation 2.27. The

conductor sag depends on conductor span, horizontal tension

and temperature. The conductor span on the south side of

the Tower 16/4 is 450 m and toward the north it is 252 m

(see Figure 9). The sags, calculated using conductor

horizontal tension of 45 kN at 60°F, are 6.3 m and 21.0 m

corresponding to the 252 m and 450 m spans, respectively (a

parabolic profile is assumed). Using ecjuation 2.27 with

estimated sags, the calculated natural frecjuencies in the

transverse direction are 0.12 Hz and 0.22 Hz. The spectrum

in Figure 15 shows small but distinguishable peaks in the

88

neighborhood of these frequencies. The two unequal spans

are expected to respond in the transverse direction

independent of each other due to the low stiffness of the

conductor. The two peaks are judged to be due to resonant

response at natural transverse frecjuencies of the conductor.

At frecjuencies close to the conductor natural

frecjuencies the gust spectrum in Figure 13 shows that wind

has a fair amount of energy. Presence of gust energy at

natural frecjuencies of the conductor can cause significant

resonant amplification of the conductor response.

The spectral areas in the zone of frecjuencies greater

than 1 Hz are less than 15% in most of the records. High

spectral peaks (Figure 15) in the 2-4 Hz frecjuency range are

believed to be due to vibration of the tower and the

frecjuency of the swing angle indicators. These frecjuencies

of the tower and instruments enter into the conductor

response records since the conductor is connected to the

tower through load cells and swing angle indicators.

Natural frequencies of Tower 16/4 are determined using

MSC/NASTRAN version 63 software. The tower structure is

modelled as a space frame, without conductors and overhead

ground wires. Natural frecjuencies of the tower are 2.88,

3.01 and 4.92 Hz, corresponding to longitudinal, transverse,

and torsional modes of vibration, respectively. In

89

addition, the frequency of vibration of the swing angle

indicator is 3.2 Hz (Kempner, 1980). These frecjuencies of

the tower and the swing angle indicator are believed to

cause high peaks in the response spectrum in the frecjuency

range of 2 to 4 Hz. Even though spectral peaks are high in

the frecjuency range above 1 Hz, the amount of energy is

relatively low.

Since the amount of gust spectral energy above 1 Hz in

Figure 13 is very small and since peaks in the response

spectra can be justified as above, it is believed that the

response spectral energy in the frecjuency range above 1 Hz

is not due to response of the conductor to wind. This

spectral energy is neglected in further consideration of the

response of the conductors. The background and resonant

responses assessed from the response spectra are compared

with the analytical procedure in the next chapter.

CHAPTER V

COMPARISON AND REFINEMENT OF THE

ANALYTICAL MODEL

The goal of the present study is to compare and to

refine the analytical model using the results of field data.

The analytical model proposed by Davenport (1980) to

determine peak response of transmission line structures was

presented in Chapter II. The key elements in determining

the fluctuating response of conductors in the model are the

background response and the resonant response (see ecjuation

2.23). In addition, the model recjuires establishment a

value for the peak factor to determine peak response (see

ecjuation 2.1).

The field data analysis yields the peak response as

well as the fluctuating response. Background and resonant

responses are determined from response spectra of field data

utilizing the procedure indicated in Chapter IV. Peak

factors are obtained from the field peak responses utilizing

the upcrossing rate procedure.

Background and resonant responses assessed from the

field data are compared with values calculated using the

analytical model. Joint acceptance function coefficients

90

91

are obtained from the field data to refine the background

response of the analytical model. Also, aerodynamic damping

ratios are recovered from the field data to improve the

resonant response of the analytical model. Since the

fluctuating response depends on many parameters, total

fluctuating responses from field data are not compared with

total fluctuating response predicted using the Davenport

model (1980).

Comparison of Analytically Predicted Mean Square Response With Field

Measured Values

One of the two parts of the fluctuating response

component in ecjuation 2.1 is the mean scjuare value. In the

frecjuency domain analysis, the mean scjuare value of response

is computed as the area under the response spectrum. The

mean scjuare value can be considered as a summation of

background response, B , and resonant response, R (ecjuation

2.23). Background response is due to the wind turbulence at

low frecjuencies, and can be considered as quasi-static

response. Resonant response is due to coincidence of

conductor natural frecjuencies with gust frecjuencies. This

resonant response is the area under the response spectrum at

frecjuencies close to conductor natural frecjuencies.

92

Field Measured Mean Scjuare Response

The gust spectrum of Figure 13 shows that the wind

turbulence has energy up to 1 Hz, and energy beyond 1 Hz is

negligible. Conductor natural frecjuencies are in the range

of 0.1 to 0.4 Hz, hence the resonant peaks should dominate

above the background response in this frecjuency range. It

is difficult to separate background and resonant responses

in the field data. In Figure 16, which is the same response

spectrum as Figure 15, there are peaks at the natural

frecjuencies of the conductors: 0.12 and 0.22 Hz. However,

these peaks are diffused. As discussed in Chapter IV, the

spectral areas between the frecjuencies of 0.1 and 1.0 Hz for

most records were less than 15% of the total area under the

response spectrum (total mean scjuare value). At the risk of

being on the high side, the total spectral areas between

frecjuencies of 0.1 and 1.0 Hz are assumed to be resonant

response. The error introduced by this assumption is small

because the response in this frecjuency range is a small

portion of the total fluctuating response. This resonant

response is indicated by R in Figure 16.

It is reasonable to assume that the area under the

response spectrum below 0.1 Hz is the background response.

This area is close to 75% of the total area (total mean

scjuare value) in most of the records as discussed in Chapter

93

a 3* CJ

o en a

CO

O d CO

CO

«M CM

u o a •a > 1 a ° CO o

Q.OOL

f - Frequency (in Hertz) S(f) - Spectral Density at f SIGSQ - Mean Sqaure Value

0.1

Frequency, f (Hertz)

10.0

Figure 16: West Conductor Response Spectrum Plot for Record NOl

IV. The area designating background response is shown as B

in Figure 16.

The area under the response spectrum for frecjuencies

above 1 Hz is not considered to be response due to extreme

wind effects. This was discussed in some detail in Chapter

IV

94

Delineation of background and resonant responses in the

field response spectra permits assessment of responses in

each of the three conductors, west, east, and central

conductors, for all twenty-one field records. Field

measured values for background and resonant responses for

west, east, and central conductors are tabulated in Tables

11 through 13, along with mean scjuare values, for each

conductor. These values are compared with values predicted

by the analytical model as described in the next section.

The mean scjuare values of the three conductors are

fairly consistent, but have scatter for a given record.

This is not surprising, since there was variation (by about

10%) in the mean response of the three conductors (refer to

Table 9). It is reasonable to expect larger variation in

fluctuating response between the conductors. The variation

in the field data suggests that results will have scatter,

and that it is important to use an ensemble of data for

appropriate interpretation of the results.

Analytical Model Predicted Mean Scjuare Value

The analytical model used to predict the background and

resonant response contains a number of wind and conductor

related parameters (see ecjuations 2.24 and 2.25). These

parameters can be separated into fixed and variable

95

TABLE 11

West Conductor Response Spectrum Data Analysis

Record Number

NOl N02 N03

N04

N05 NO 6 N07

2 2

2 M

OO

O

vD

00

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

2 2

2 to

to

to

CA

) to

M

Mean Scjuare

0.211 0.184 0.043

Background Response

Field Analytical Measured Model*

0.141 0.132 0.028

Zero Wind Record

0.017 0.039 0.790

0.582 0.009 0.270

0.366 0.165 0.441

0.003 0.028 0.609

0.447 0.007 0.199

0.269 0.151 0.300

Reference Zero Wind

0.098 0.013 0.377

0.606 0.273 0.014

0.258 0.361 0.250

0.077 0.011 0.309

0.486 0.236 0.005

0.190 0.243 0.194

0.201 0.038 0.052

0.005 0.043 0.143

0.365 0.014 0.161

0.256 0.033 0.217

Resonant

Field Measured

0.033 0.024 0.009

0.008 0.006 0.087

0.057 0.001 0.037

0.043 0.010 0.077

Speed Record

0.069 0.005 0.152

0.186 0.125 0.002

0.076 0.151 0.177

0.015 0.002 0.041

0.066 0.024 0.002

0.042 0.060 0.036

. Response

Analytical Model*

0.254 0.043 0.056

0.006 0.048 0.203

0.502 0.012 0.200

0.348 0.034 0.288

0.077 0.004 0.192

0.259 0.140 0.003

0.095 0.194 0.194

* Davenport, 1980

96

TABLE 12

East Conductor Response Spectrum Data Analysis

Record Number

rH

CM

C

O

o o

o

2 2 2

N04

N05 NO 6 N07

2 2

2 M

OO

O

vD

00

Nil N12 N13

N14

N15 N16 N17

2 2

2 to

M M

O

VD

00

2 2

2 to

to

to

CA)

to M

Mean Scjuare

0.168 0.162 0.037

Background Response

Field Analytical Measured Model*

0.115 0.128 0.023

Zero Wind Record

0.009 0.033 0.677

0.572 0.013 0.207

0.310 0.165 0.348

0.002 0.021 0.541

0.450 0.009 0.157

0.231 0.153 0.248

Reference Zero Wind

0.078 0.020 0.321

0.512 0.205 0.008

0.193 0.276 0.187

0.059 0.017 0.268

0.420 0.178 0.002

0.147 0.199 0.153

0.181 0.037 0.050

0.004 0.041 0.141

0.322 0.015 0.155

0.245 0.037 0.195

Resonant Response

Field Measured

0.034 0.024 0.011

0.002 0.006 0.080

0.061 0.001 0.034

0.047 0.011 0.062

Speed Record

0.062 0.007 0.133

0.155 0.110 0.002

0.066 0.133 0.150

0.016 0.003 0.041

0.067 0.023 0.002

0.036 0.055 0.027

Analytical Model*

0.229 0.042 0.054

0.005 0.045 0.200

0.443 0.012 0.193

0.333 0.038 0.259

0.069 0.005 0.167

0.216 0.124 0.003

0.082 0.171 0.164

* Davenport, 1980

97

TABLE 13

Central Conductor Response Spectrum Data Analysis

Record Number

NOl N02 N03

NO 4

NO 5 NO 6 N07

NOB NO 9 NIO

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

Mean Scjuare

0.284 0.183 0.040

Background Response

Field Analytical Measured

0.175 0.119 0.026

Zero Wind Record

0.011 0.047 1.004

0.618 0.020 0.287

0.432 0.148 0.471

0.004 0.036 0.714

0.467 0.009 0.206

0.294 0.136 0.308

Reference Zero Wind

0.117 0.014 0.443

0.638 0.300 0.019

0.320 0.420 0.269

0.094 0.012 0.365

0.524 0.261 0.005

0.236 0.271 0.210

Model*

0.211 0.035 0.053

0.005 0.043 0.147

0.391 0.015 0.163

0.269 0.035 0.222

Resonant

Field Measured

0.038 0.021 0.008

0.005 0.007 0.085

0.046 0.010 0.035

0.038 0.009 0.065

Speed Record

0.069 0.006 0.146

0.197 0.128 0.002

0.078 0.155 0.175

0.016 0.001 0.038

0.054 0.024 0.013

0.049 0.061 0.032

Response

Analytical Model*

0.207 0.031 0.044

0.004 0.037 0.162

0.417 0.009 0.158

0.283 0.028 0.228

0.060 0.004 0.142

0.212 0.111 0.002

0.076 0.155 0.148

* Davenport, 1980

98

parameters. Some of the parameters depend on the geometry

and physical characteristics of the conductors. These

parameters such as conductor sag, diameter of the conductor,

effective height, etc. are fixed parameters. On the other

hand, other parameters such as mean wind speed, turbulence

intensity and aerodynamic damping ratio vary with each wind

record; these parameters are considered as variable

parameters.

The fixed and assumed parameters used in the analytical

model to calculate background and resonant responses are

tabulated in Table 14. Each conductor bundle consists of

two Chukar conductors, hence the effective diameter used in

the model is twice the diameter of the individual conductor.

The mean wind speed at the effective conductor height is

determined using the recorded mean wind speed at 34.7 m and

the power-law exponent, a = 0.14.

Table 15 shows the mean wind speeds calculated for the

west, east, and central conductors at their effective

heights. The exposure factor, E, in ecjuations 2.24 and 2.25

is twice the turbulence intensity. The exposure factor

values in Table 15 are obtained from the turbulence

intensity recorded at 34.7 m on Tower 16/4 (refer to Table

7). The exposure factors used in the model at the effective

heights of the conductors are assumed to be the same as at

99

TABLE 14

Fixed and Assumed Parameters Used in the Analytical Model

Parameters Typical Values Used

0 . 0 8 m

376 m 402 m

15 .7 2 3 . 8

m m

Values based on physical characteristics

(1) Conductor diameter (d)

(2) Effective span (L) east and west conductors central conductor

(3) Effective height (h) east and west conductors central conductor

(4) Conductor fundamental frecjuency (f ) 0.12 Hz

Assumed values*

(5) Conductor force coefficient (C^) 1.0

(6) Coherence exponent (c) 8

(7) Scale of turbulence (L ) 65 m ^ ' s

(8) Kaimal's gust spectrum constant (A) 0.28

(9) Kaimal's gust spectrum constant (n) 0.67

*assumed values are recommended by Davenport (1980)

34.7 m on Tower 16/4. The conductor aerodynamic damping

ratio is calculated using ecjuation 2.26. Since the

aerodynamic damping ratio depends on the mean wind speed, it

is different for each record.

100

TABLE 15

Variable Parameters Used in the Analytical Model

Record Number

NOl N02 N03

N04 Ze

Exposure Factor

0.36 0.34 0.22

ro Wind Rec<

Mean

@ 15.7

16.9 14.7 13.3

ord

Wind

m*

Speed (mps)

@ 23.8 m**

17.9 15.5 14.1

N05 NO 6 NO 7

222

MOO

O v

D 00

Nil N12 N13

N14

N15 N16 N17

2 2 2

to M M

O V

D 00

2 2 2

to to to

CO to M

0.06 0.22 0.32

0.28 0.26 0.34

0.36 0.34 0.30

Reference

0.24 0.24 0.28

0.22 0.34 0.04

0.22 0.30 0.42

14.2 13.9 20.1

19.3 8.6 16.6

18.9 12.6 18.2

Zero Wind Speed Record

14.1 9.0 16.9

19.6 14.2 16.6

16.6 17.3 13.6

15.0 14.7 21.2

20.4 9.1 17.5

19.9 13.2 19.2

14.9 9.5 17.8

20.7 15.0 17.5

17.5 18.3 14.4

* effective height for west and east conductors ** effective height for central conductor

101

Instead of calculating the mean wind pressure, P, and

the influence coefficient, 0 (which translates the pressure

to response), the field measured mean transverse load

components are used for each record (refer to Table 9).

Calculated background and resonant responses using the

analytical model are shown for the three conductors in

Tables 11 through 13. The values are obtained using

ecjuations 2.24 and 2.25. The majority of the records show

that the background response calculated from the analytical

model is smaller than the background response measured in

the field. Field measured values versus analytical model

values of background response of the three conductors are

plotted in Figure 17. The figure shows that the background

response predicted by the model is an underestimation of the

measured value. Since the background response accounts for

75% of the mean scjuare response value, refinement of the

analytical model of this part is desirable.

The analytical model predicts higher resonant responses

than the field measured values (refer to Tables 11 to 13).

Field measured values versus analytical model values of

resonant response of the three conductors are plotted in

Figure 18. The figure clearly shows that the predicted

values are very much higher than the field measured values.

One of the significant variables in the analytical model is

102

0.8

U

s •a >

•a

•a

o West Cond. East Cond. Central Cond

0.6-

0.4-

1 1 * r—

0.4 0.6

Field Measured Values

0.8

Figure 17: Analytical Model Background Response Versus Field Measured Background Response

the aerodynamic damping ratio. If the damping ratio is

higher than predicted by equation 2.26, the calculated

resonant response will be smaller. Field measured data are

used to assess a possible damping ratio for each record.

103

Even though a large scatter in evaluation of the damping

ratio is expected, it can lead to a better prediction of

resonant response.

0.60

(/)

3 •a >

•a o •a c <

0.45-

0.30-

0.15-

0.00

Q

• •

B a

• ° y Q n /

a • /

- s°" v y v n / y y^

Q / y y^ a ^9 /yy

\™ 1

" /

1

V

Q

a •

V

West Cond. East Cond. Central Cond.

_ ,

0.00 0.15 0.30 0.45 0.60

Field Measured Values

Figure 18: Analytical Model Resonant Response Versus Field Measured Resonant Response

104

Refinement of the Analytical Model

As noted in previous sections, the analytical model

underestimates the background response and overestimates the

resonant response. Refinement of the analytical background

expression using the field measured data is attempted. The

resonant response prediction is improved by recovering a

damping ratio for the conductor from the field data.

Background Response

The mean scjuare value of fluctuating response can be

calculated as the area under the response spectrum,

al = J Sj (f) df (5.1) 0

2 where a„ = mean scjuare value of response,

So(f) = spectral density value of response, and

f = frecjuency.

Utilizing ecjuations 2.10 and 2.11, equation 5.1 can be

expressed as

_2 2 = iL- J y^^(f) |H(f)|2 S^(f) df (5.2)

- 0 V

where S (f) = gust spectral density,

F = mean transverse force,

V = mean wind speed.

105

2 Z (f) = aerodynamic admittance function, and

2 |H(f)| = mechanical admittance function.

The area under the response spectrum is a summation of

background and resonant responses. Ecjuation 5.2 can be

written in a simple form as:

_2 2 4F R = — T B ^ Aj,| (5.3)

V

where Ag accounts for background response, and A„ accounts

for resonant response. Davenport (1977) developed ecjuations

for Ag and Ap as

Ag = j x^(4r) S (f) ^ (5.4) 0 V

and

f L AR = X^(-^) S (f ) J |H(f)|^ df (5.5)

V 0

where f is the fundamental frecjuency of the structure and

other terms are defined above. The background response due

to wind turbulence can be obtained using ecjuations 5.3 and

5.4, if the gust spectrum is defined by some appropriate

analytical function.

The aerodynamic admittance function is a relationship

between the gust spectral density function and the force

106

spectral density function in the frequency domain. It is a

measure of the effect that the wind turbulence has on the

transverse forces. The shape of the conductor and the sizes

of gusts relative to the size of the conductor influence

this function. A large gust, totally enveloping the

structure, is well correlated, while a small gust, acting on

a portion of the conductor, is uncorrelated.

The aerodynamic admittance function is usually

2 f L expressed in a nondimensional form as x (- - )/ where L is V

the conductor span, and the ratio -F ^ designated as the

scale of turbulence (L ). The aerodynamic admittance

function is termed as a 'joint acceptance function (JAF),'

if it is modified to account for the mode shape. In other

words, the important link between the gust fluctuations,

(described by the gust spectrum) and the modal force

fluctuations is provided by the the JAF. This function

depends on the mode shape and the velocity field, which

varies widely from structure to structure. Davenport (1977)

reduced the JAF to a simple form as below:

|JAF|2 = -^—^ (5.6) ' ' 1 + m (p

where m is a constant to account for the mode shape and

107

<p = cfL/V. The quantities c, f, L, and V represent the

coherence exponent, frequency, conductor span and mean wind

speed, respectively.

The theory described for computing conductor response

in Chapter II is based on the conventional assumption of a

constant force coefficient. For conductors with a

cylindrical shape the force coefficient, C^, depends on the

Reynolds Number (refer to Figure 3). This change in force

coefficient affects the fluctuating component of response.

The analytical model of fluctuating response should account

for changes in the force coefficient at Reynolds Numbers

corresponding to the mean wind speed (Davenport, 1980). To

account for this effect the numerator of ecjuation 5.6 is

replaced with an unknown constant Q. The resluting

ecjuation, which is a product of JAF and Q is simply termed

as JAF in this study, as shown below:

|JAF|2 = 2 _ ^ _ . (5.7) 1 + M(-^)

V

In addition to introduction of coefficient Q in ecjuation

5.7, the coefficient M is used to account for mode shape and

the coherence exponent, c (transverse correlation of

turbulence). The available data are not able to provide

separate coefficients for the mode shape and correlation of

108

turbulence. Equation 5.7 is in the same form as equation

2.24 given in the analytical model developed by Davenport

(1980). In 1:he model, Davenport uses approximate values of

1 and 0.81 for the coefficients Q and M, respectively. Here

the field response data are used to evaluate these two JAF

coefficients.

Determinincr the JAF Coefficients

The frecjuency transfer function (FTF) is a transfer

function between the spectral densities of fluctuating wind

turbulence and conductor response. The FTF can be

considered as the product of the aerodynamic admittance

function and mechanical admittance function. The FTF can be

written as

^SR(f) v^ 2 2 5: 1-^- = 4 |H(f)r IJAFr. (5.8) _ 2 fS (f) I V /I I I V /

F

The FTF can be obtained by plotting the ratios of the

nondimensionalized response spectral values to the

nondimensionalized gust spectral values (refer to equation

5.8). As explained in Chapter IV, the IMSL program FTFREQ

is used to compute the spectral density values of wind

turbulence and conductor response fluctuations. FTF plots

for all 21 records of west, east, and central conductors

109

were obtained. A typical FTF plot for the west conductor

for record NOl is shown in Figure 19. The spectral density

values of wind turbulence above 1 Hz are very small (refer

to Figure 13), and use of very small values in the

denominator of the FTF would be inappropriate. Hence, the

FTF values are plotted up to 1 Hz only. It was noted in

Chapter IV that wind gust and response spectra show cjuite a

bit of fluctuation because of the computational procedures

used in obtaining the spectral densities. The FTF plot in

Figure 19 is obtained from the ratios of two spectra, so

large fluctuations in ordinates are not surprising.

The mechanical admittance function, also known as the

dynamic amplification factor, depends on the structural

dynamic properties such as frecjuencies and damping ratios.

This factor amplifies the response spectrum (resonant

response) at the natural frecjuencies of the conductor. The

JAF related to background response can be obtained by

removing the resonant peaks at the natural frecjuencies of

the conductor from the FTF plot. Equation 5.7 was fitted to

the field measured JAF plot by regression analysis to obtain

values of the coefficients Q and M. Equation 5.7 is a

nonlinear ecjuation, hence nonlinear regression needed to be

applied. The SAS procedure NLIN (SAS, 1982) was used to fit

the nonlinear equation to the computed JAF field response

110

28.0 -,

I I I I 1 1 1 1 '

0.001 10.0

Frequency (Hertz)

Figure 19: Frecjuency Transfer Function of West Conductor Response for Record NOl

data. The procedure NLIN is used to fit ecjuation 5.7 to the

field data of all 21 records of west, east, and central

conductors.

Procedure NLIN implements iterative methods that

attempt to find least squares estimates for the nonlinear

equations. Parameter names and starting values, expressions

for the model, and expressions for derivatives of the model

Ill

with respect to the parameters need to be specified. Based

on expectations, the specified ranges for the coefficients Q

and M were 0.4-1.0 and 0.1-0.4, respectively. The NLIN

procedure first examined the starting value specifications

of the parameters in the specified search grid. The NLIN

procedure then evaluated the residual sum of scjuares at each

combination of values to determine the best values to start

the iterative algorithm. A modified Gauss-Newton iterative

method (SAS, 1982) was used, which involved regressing the

residuals on the partial derivatives of the model with

respect to the parameters until the iterations converged.

Some variation in data was expected, since the data were

measured in the field.

To find the best coefficient values, which in general

satisfied most of the records, a contour of lowest residual

sum of scjuares was plotted in the specified search grid.

Any combination of coefficients Q and M within the lowest

residual sum of squares contour is acceptable. A typical

contour plot for west conductor response for record NOl is

shown in Figure 20. Combining all plots of three conductors

(west, east, and central), there are 63 contour plots of

residual sum of scjuares. The sixty three contour plots show

some degree of dispersion of lowest residual sum of scjuare

contours over the search grid. The contour plots were

112

e (J

i

0.1 0.2 0.3 0.4

Coefficient, M

Figure 20: West Conductor JAF Coefficients Contour Plot for Record N15

113

overlapped to find the best values of the coefficients Q and

M, which were 0.45 and 0.2, respectively. These values are

significantly lower than the values of 1.0 and 0.81 used in

tihe Davenport model (equation 2.24). For better

visualization the JAF with Davenport model values and

refined values are plotted on the same graph as shown in

Figure 21.

The low value of coefficient M obtained from the field

data may be because of a low value of coherence exponent, c,

due to the long span of the conductors. Also, when

-=- >> 1, the joint acceptance function is independent of ^s

the mode shape and is proportional to the ratio of the

correlation length to the conductor span. These comments

are based on the results of wind tunnel experiments

conducted on a rod (Blevins, 1977).

The expression for background response of the

conductor, ecjuation 2.24, with the new coefficients is

— ^ 2 2 B^ = P Ol E^

0.45

1 + 0.2(-^) ^s

(5.9)

The parameters in ecjuation 5.9 are defined below equation

2.24. The background response calculated using the refined

analytical model for all 21 records of west, east, and

central conductors are tabulated in Tables 16 through 18.

o CJ C

£ o c

o CJ

<

c 'o

o CM

O O o

o LT)

o o in

o

o o o -r 1 1 — I — I I M

0.001 0.0 ]

114

1 - Davenport Model 2 - Refined Model

- I 1 — I — r i l l -I 1 1—I—I I I "T 1 1—I—I I I I

0. 1.0 10.c

Reduced Frequency - fLA^

Figure 21: Joint Acceptance Function Plot

These tables also show the ratios of the refined analytical

model values to the field measured values and the ratios of

the analytical model values to the field measured values.

In general, these nondimensional ratios show a comparison

between the analytical model and the field response values.

For better visualization the field measured values versus

115

the refined analytical model values are plotted in Figure

22. The refined analytical model gave slightly better

predicted than the analytical model when Figures 22 and 17

are compared. In Tables 16 through 18, means and

coefficient of variations (COV) of the ensemble of the

ratios are shown. In each table, the mean of the ratio is

closer to 1 for the refined analytical model. However, the

COV for each conductor did not change. This improvement in

mean value of the ensemble and insignificant change in COV

value are due to inherent scatter in the field data.

Resonant Response

As noted earlier the analytical model overestimates the

resonant response. One of the reasons may be the use of low

damping ratio values as determined by ecjuation 2.26. Here

field measured resonant response data are used to estimate

damping ratios for the conductors.

Determining the Aerodynamic Damping

Ratio

As noted in Chapter II, three types of damping are

noted for conductor response, namely, material, structural,

and aerodynamic damping. For conductors aerodynamic damping

is very much higher than material or structural damping.

Therefore, both material and structural dampings are

TABLE 16

Background Response of West Conductor

116

Record Number

rH

CM

CO

O

O

O

22

2

N04

N05 NO 6 NO 7

22

2 M

OO

O

vD

00

Nil N12 N13

N14

N15 N16 N17

22

2 to

M M

O

VD

00

rH

CM

C

O

CM

CM

C

M

2 2 2

Mean Scjuare

0.211 0.184 0.043

Field Measured

0.141 0.132 0.028

Zero Wind Record

0.017 0.039 0.790

0.582 0.009 0.270

0.366 0.165 0.441

0.003 0.028 0.609

0.447 0.007 0.199

0.269 0.151 0.300

Reference Zero Wind

0.098 0.013 0.377

0.606 0.273 0.014

0.258 0.361 0.250

mean value coefficient of

0.077 0.011 0.309

0.486 0.236 0.005

0.190 0.243 0.194

variation

Refined Model

0.239 0.037 0.062

0.006 0.051 0.170

0.432 0.017 0.191

0.304 0.039 ,0.257

Speed Rec

0.082 0.006 0.180

0.220 0.148 0.003

0.090 0.179 0.210

Ratio (1)*

1.695 0.280 2.214

2.000 1.821 0.279

0.966 2.429 0.960

1.130 0.258 0.857

ord

1.065 0.546 0.583

0.453 0.627 0.600

0.474 0.739 1.083

1.003 65.5%

Ratio (2)*

1.426 0.288 1.857

1.667 1.536 0.235

0.817 2.000 0.809

0.952 0.219 0.723

0.896 0.455 0.492

0.383 0.530 0.400

0.400 0.621 0.912

0.839 65.3%

(1)* ratio of refined model value to the measured value (2)* ratio of analytical model value to the measured value

TABLE 17

Background Response of East Conductor

117

Record Number

NOl N02 N03

N04

Mean Scjuare

0.168 0.162 0.037

Field Measured

0.115 0.128 0.023

Zero Wind Record

Refined Model

0.215 0.043 0.060

Ratio (D*

1.870 0.336 2.609

Ratio (2)*

1.574 0.289 2.174

N05 NO 6 NO 7

N08 N09 NIC

Nil N12 N13

0.009 0.033 0.677

0.572 0.013 0.207

0.310 0.165 0.348

0.002 0.021 0.541

0.450 0.009 0.157

0.231 0.153 0.248

0.005 0.022 0.168

0.382 0.018 0.184

0.291 0.043 0.231

2.500 1.048 0.311

0.849 2.000 1.172

1.260 0.281 0.932

2.000 1.952 0.261

0.716 1.667 0.987

1.061 0.242 0.786

N14 Reference Zero Wind Speed Record

N15 N16 N17

N18 N19 N20

N21 N22 N23

mean

0.078 0.020 0.321

0.512 0.205 0.008

0.193 0.276 0.187

value coefficient of

0.059 0.017 0.268

0.420 0.178 0.002

0.147 0.199 0.153

variation

0.073 0.008 0.157

0.184 0.131 0.003

0.079 0.158 0.178

1.237 0.471 0.586

0.438 0.736 1.500

0.537 0.794 1.163

1.078 63.8%

1.051 0.412 0.496

0.369 0.618 1.000

0.449 0.668 0.980

0.940 64.3%

(1)* ratio of refined model value to the measured value (2)* ratio of analytical model value to the measured value

TABLE 18

Background Response of Central Conductor

118

Record Number

Mean Scjuare

Field Measured

Refined Model

Ratio (D*

NOl N02 N03

N04

NO 5 NO 6 NO 7

NOB N09 NIO

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

0 .284 0 .183 0 .040

0 .175 0.119 0 .026

0 .256 0 .042 0 .064

1.463 0 .356 2 .462

Zero Wind Record

0.011 0.047 1.004

0.618 0.020 0.287

0.432 0.148 0.471

0.004 0.036 0.714

0.467 0.009 0.206

0.294 0.136 0.308

0.006 0.052 0.178

0.472 0.018 0.197

0.325 0.042 0.269

1.500 1.444 0.249

1.011 2.000 0.956

1.105 0.309 0.873

Reference Zero Wind Speed Record

0 .117 0 .014 0 .443

0 .638 0 .300 0 .019

0 .320 0 .420 0 .269

0 .094 0.012 0 .365

0 .524 0 .261 0.005

0 .236 0 .271 0.210

0 .083 0 .008 0 .176

0 .238 0 .155 0 .003

0 .095 0 .188 0 .211

0 .883 0 .667 0 .482

0 .454 0 .594 0 .600

0 .403 0 .694 1.005

mean value coefficient of variation

0.929 62.0%

Ratio (2)*

1.206 0 .294 2 .039

1.250 1.194 0 .206

0 .837 1.667 0 .791

0 .915 0 .257 0 .721

0 . 7 3 4 0 . 5 0 0 0 . 4 0 0

0.376 0 .490 0 .400

0 .331 0 .572 0 .833

0 .763 63.4%

(1)* ratio of refined model value to the measured value (2)* ratio of analytical model value to the measured value

119

0.8

_3

>

•s c

0.6-

0.4-

" West Cond. D East Cond. • Central Cond

0.2-

0.0 0.2 0.4 0.6 0.

Field Measured Values

Figure 22: Refined Model Background Response Versus Field Measured Values

neglected in this study, and the computed total damping is

assumed to be aerodynamic damping.

The response of the conductor at its natural frequency

of vibration is a function of excitation force and damping.

120

The magnification factor method is used here to estimate the

aerodynamic damping ratio. For a single degree of freedom

system subjected to wind turbulence, the resonant peak is

amplified at a fundamental frecjuency of the conductor. The

height of this peak is controlled by the damping for the

conductor.

The frecjuency transfer function (Figure 19) described

in the previous section is used to establish damping. As

noted earlier, the FTF is a combination of aerodynamic

admittance function and the mechanical admittance function.

The mechanical admittance function amplifies the response at

the natural frecjuency of vibration of the conductor. The

expression for the mechanical admittance function is defined

in ecjuation 2.13. At the fundamental frecjuency of the

conductor, the ecjuation for the mechanical admittance

function simplifies as follows:

|H(f^)|^ = - V (-^ 4 C

where C = aerodynamic damping ratio, and

f = fundamental frecjuency of conductor.

The damping ratios of the conductors are determined

using the heights of the resonant peaks in the FTF plots.

As noted in Chapter III, the conductor spans on two sides of

Tower 16/4 are different. The natural frecjuencies of

121

vibration of the conductors are 0.12 Hz and 0.22 Hz,

corresponding to spans of 450 m and 252 m, respectively.

The aerodynamic damping ratios estimated to be related to

these two natural frecjuencies of the west, east, and central

conductors from the FFT are tabulated in Table 19.

Conductor aerodynamic damping ratios for east wind records

(N05, N09 and N20) are not calculated because the peaks in

the FTF plots for these records are highly erratic. The

east wind records give poor results for the FTF because of

low turbulence intensities in tihe records. The estimated

aerodynamic damping ratios in Table 19 vary between 18 and

91%. This large scatter in establishing damping ratios is

expected because computational technicjues used to obtain the

spectra cause large fluctuations in FTF. In addition only

two specific peak values, closest to the frecjuencies of 0.12

and 0.22 Hz, are used in each FTF plot. The peak values in

the FTF plots are not expected to be highly accurate. Most

of the estimated damping values in Table 19 fall between 30%

and 60%.

The aerodynamic damping values predicted by ecjuation

2.26 are between 5% and 11% based on the mean winds recorded

in the field. These values used in the analytical model are

significantly smaller than the ones estimated from the field

data. In recognition of this discrepancy, a conservative

TABLE 19

122

Estimated Aerodynamic Damping Ratios in Percentages

Record Number

NOl N02 N03

N04

N05 NO 6 NO 7

NOB N09 NIO

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

West Conductor (1)*

45 66 43

Zero Wind

_

50 18

52 -

34

32 50 39

Reference

50 88 47

33 42 -

45 58 58

(2)*

43 66 91

Record

_

30 33

35 -

38

58 52 67

Zero Wind

41 75 69

35 46 -

29 33 29

East Conductor (1)*

43 32 38

^

41 18

50 -

27

32 22 44

Speed

45 29 44

30 45 —

45 60 60

(2)*

41 34 67

^

32 33

35 -

45

58 54 60

Record

40 28 75

41 42 —

29 32 34

Central Conductor (1)*

46 27 62

^

65 47

55 —

30

38 28 47

50 50 54

37 54 —

41 56 67

(2)*

44 34 66

^

37 22

35 —

41

50 50 63

37 50 52

51 42 —

27 33 56

(1)* corresponding to (2)* corresponding to

resonant peak at 0.12 Hz resonant peak at 0.22 Hz

123

ensemble average value of 40% aerodynamic damping ratio is

suggested for conductors. Resonant responses are calculated

with this suggested aerodynamic damping for comparison

purposes.

Resonant Response with Suggested Damping Ratio

Resonant response values for all three (west, east, and

central) conductors are calculated using an aerodynamic

damping ratio of 40% in the analytical model. The values

are tabulated in Tables 20 through 22. The tables also show

the field measured resonant response and the total mean

scjuare value for each record.

In addition, ratios of resonant responses obtained from

the analytical model with 40% damping to field measured

values and from the analytical model with damping from

ecjuation 2.26 to field measured values are shown in the

tables. Use of 40% damping improves the prediction of

resonant response significantly. Mean values of the ratios

for 40% damping are close to unity. The COV of the ratios

in the tables are not effected significantly, though this is

misleading. The mean values of the ratios of responses from

the analytical model with damping from ecjuation 2.26 are a

little more than 4; hence associated COV values of 45%

reflect a large variation.

124

TABLE 20

West Conductor Resonant Response With 40% Damping

Record Number

NOl N02 N03

NO 4

NO 5 N06 NO 7

NOB N09 NIO

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

Mean Scjuare

0.211 0.184 0.043

Zero Wind

0.017 0.039 0.790

0.582 0.009 0.270

0.366 0.165 0.441

Reference

0.098 0.013 0.377

0.606 0.273 0.014

0.258 0.361 0.250

mean value

Field Measured

0.033 0.024 0.009

Record

0.008 0.006 0.087

0.057 0.001 0.037

0.043 0.010 0.077

Zero Wind

0.015 0.002 0.041

0.066 0.024 0.002

0.042 0.060 0.036

coefficient of variation

Analytical Model

0.053 0.008 0.009

^

0.008 0.051

0.119 —

0.041

0.081 0.006 0.064

Speed Reco

0.014 0.001 0.039

0.062 0.025

-

0.019 0.042 0.033

Ratio (1)*

1.606 0.333 1.000

1.333 0.586

2.088 •

1.108

1.884 0.600 0.831

rd

0.933 0.500 0.950

0.939 1.042

-

0.452 0.700 0.917

0.989 48.5%

Ratio (2)*

7.697 1.792 6.222

8.000 2.333

8.807 .

5.405

8.093 3.400 3.740

5.133 2.000 4.683

3.924 5.833

-

2.262 3.233 5.389

4.886 45.8%

(1)* ratio of analytical model values with 40% damping ratio to the field measured values

(2)* ratio of analytical model values with damping ratio from ecjuation 2.26 to the field measured values

125

TABLE 21

East Conductor Resonant Response With 40% Damping

Record Number

NOl N02 N03

N04

N05 NO 6 N07

NOB N09 NIC

Nil N12 N13

N14

N15 N16 N17

N18 N19 N20

N21 N22 N23

Mean Scjuare

0.168 0.162 0.037

Zero Wind

0.009 0.033 0.677

0.572 0.013 0.207

0.310 0.165 0.348

Reference

0.078 0.020 0.321

0.512 0.205 0.008

0.193 0.276 0.187

mean value

Field Measured

0.034 0.024 0.011

Record

0.002 0.006 0.080

0.061 0.001 0.034

0.047 0.011 0.062

Zero Wind

0.016 0.003 0.041

0.067 0.023 0.002

0.036 0.055 0.027

coefficient of variation

Analytical Model

0.047 0.008 0.009

^

0.008 0.050

0.105 -

0.039

0.078 0.006 0.057

Ratio (1)*

1.382 0.333 0.818

.

1.333 0.625

1.721 •

1.147

1.660 0.550 0.919

Speed Record

0.012 0.001 0.035

0.052 0.022 -

0.017 0.036 0.027

0.750 0.333 0.854

0.776 0.957

-

0.472 0.665 1.000

0.905 45.4%

Ratio (2)*

6.735 1.750 4.909

7.500 2.500

7.262 _

5.676

7.085 3.455 4.177

4.313. 1.667 4.073

3.224 5.391

-

2.278 3.109 6.074

4.510 42.6%

(1)* ratio of analytical model values with 40% damping ratio to the field measured values

(2)* ratio of analytical model values with damping ratio from ecjuation 2.26 to the field measured values

126

TABLE 22

Central Conductor Resonant Response With 40% Damping

Record Number

rH

CM

CO

O

O O

2

2 2

N04

N05 NO 6 NO 7

22

2 M

OO

O

vD

00

Nil N12 N13

N14

N15 N16 N17

2 2

2 to

M M

O

VD

00

rH

CM

CO

CM

C

M

CM

2 2 2

Mean Scjuare

0.284 0.183 0.040

Field Measured

0.038 0.021 0.008

Zero Wind Record

0.011 0.047 1.004

0.618 0.020 0.287

0.432 0.148 0.471

0.005 0.007 0.085

0.046 0.010 0.035

0.038 0,009 0.065

Analytical Model

0.045 0.006 0.008

0.007 0.043

0.105

0.034

0.070 0.005 0.054

Reference Zero Wind Speed Rec

0.117 0.014 0.443

0.638 0.300 0.019

0.320 0.420 0.269

0.016 0.001 0.038

0.054 0.024 0.013

0.049 0.061 0.032

mean value coefficient of variation

0.011 0.001 0.032

0.054 0.020

0.017 0.035 0.026

Ratio (D*

1.184 0.286 1.000

1.000 0.506

2.283

0.971

1.842 0.556 0.831

ord

0.688 1.000 0.842

1.000 0.833

0.347 0.574 0.813

0.920 53.0%

Ratio (2)*

5.447 1.476 5.500

5.286 1.906

9.065

4.514

7.447 3.111 3.508

3.750 4.000 3.737

3.926 4.625

1.551 2.541 4.625

4.223 45.8%

(1)* ratio of analytical model values with 40% damping ratio to the field measured values

(2)* ratio of analytical model values with damping ratio from equation 2.26 to the field measured values

127

For better visualization, a plot of the resonant

response predicted by the analytical model with 40% damping

versus the field measured resonant values is shown in Figure

23. This figure, when compared with Figure 18, shows that

the analytical model predicted better resonant values with

40% aerodynamic damping ratio. Figure 23 also illustrates

the inherent scatter in the field data.

Peak Factors

Another important component of the analytical model for

fluctuating response is the peak factor, g. It is used to

predict the peak response value that can occur in a time

segment. The peak factor is defined as the number of root

mean scjuare values by which the peak value exceeds the mean

value. The peak factor for field measured response data is

calculated using the ecjuation

g = JLJL_R (5.11) ^R

where ft = peak response value,

R = mean response value, and

<Tp = root mean scjuare of response.

The peak factor values vary depending on the averaging

time interval; the smaller the peak averaging time interval

the higher the peak factor value. The peak factors from the

128

0.12

u 3 •a >

"8 :z "a u •c c

<

0.09-

• West Cond. a East Cond. • Central Cond

0.06-

0.03-

0.00 0.00 0.03 0.06 0.09 0.12

Field Measured Values

Figure 23: Analytical Model Resonant Response With 40% Damping Versus Field Measured Values

field data for response of the three conductors (west, east,

and central) are calculated and tabulated in Table 23.

The values computed are based on 0.1 and 1 second time

averaged peak values. As expected, peak factors calculated

129

TABLE 23

Peak Factors for Conductor Response

Record Number

NOl N02 N03

NO 4

West Conductor

5.62* 4.00** 3.97 3.03 4.54 3.94

Zero Wind Record

East Conductor

6.60 4.54 3.99 3.16 5.12 4.32

Central Conductor

6.91 3.26 5.37 2.69 5.53 4.54

N05 NO 6 NO 7

NOB N09 NIC

Nil N12 N13

4.77 4.07 5.14

3.48 3.28 4.29

6.67 3.76 5.45

2.07 3.30 3.20

2.74 2.88 3.34

4.52 3.60 4.20

4.18 4.02 3.74

3.42 3.30 4.33

6.78 3.37 7.09

2.31 3.00 3.12

2.72 2.73 3.38

5.11 3.15 4.69

3.71 4.75 6.40

3.59 3.52 4.32

6.06 3.95 4.45

2.80 4.28 3.04

2.52 3.07 3.43

4.02 3.58 3.36

N14

N15 N16 N17

NIB N19 N20

N21 N22 N23

Reference Zero Wind Speed Record

* based on 0.1 second peak values (instant peaks) ** based on 1 second average peak values

4.15 3.88 5.68

3.89 4.01 3.75

4.02 5.62 3.32

3.75 3.69 4.61

3.18 3.52 2.01

3.29 4.43 2.42

4.65 3.69 5.71

4.19 4.63 4.37

4.83 5.99 3.16

3.85 3.34 4.90

3.05 3.88 2.15

3.85 5.01 2.51

4.76 4.12 4.67

3.94 4.41 4.02

4.01 4.10 4.09

3.81 3.68 3.86

3.13 3.51 3.04

3.27 3.34 2.47

130

for 0.1 second peaks are higher than those calculated for 1

second time average. The peak factors for 0.1 second time

averages range between 3.16 and 6.78, and for 1 second time

averages they range between 2.01 and 5.11. The suggested

range for peak factors in the analytical model is 3.5 to 4.0

(Davenport, 1980). Many records show that the peak factors

measured in the field are higher than 4.0. This is not

surprising in view of the fact that wind speed fluctuations

tend to be Gaussian, while the response fluctuations

associated with flow separation are often highly

intermittent, thus giving rise to large peak factors. The

peak factors vary from record to record and their

distribution functions are recjuired in order to establish

peak values with a specified probability of being exceeded.

Peak factors as function of the probability distribution of

upcrossings, or ecjuivalently, a specified number of

occurrences in a given interval of time, are obtained from

the field data.

Probabilistic Peak Factors from Field Data

For a stationary Gaussian process the cumulative

probability distribution in terms of upcrossings can be

stated as (refer to ecjuation 2.21).

131

2 P(>x) = exp - { - ^ 1 (5.12)

2 G^ X

where P(>x) = probability of upcrossing.

X - threshold level specified (=gCT ), and

CT - root mean scjuare.

The Rayleigh distribution function in ecjuation 5.12 can

be expressed in Weibull distribution form as

k P(>x) = exp -(-2.) (5.13)

where g = X /

and c , k = c o n s t a n t s .

Ecjuation 5 .13 can be expanded as

I n ( - I n P(>x) ) = k l n ( g ) - k l n ( c ) . ( 5 . 1 4 )

A graph of P(>x) versus x on an appropriate log scale

will yield k as the slope of the straight line and c as the

zero intercept.

Upcrossing rates were calculated for different

threshold values (multiples of RMS) of response of all three

conductors for each record. The field data used were the

0.1 second time interval responses. A linear regression

line was fitted for each record to assess trends with

132

respect to wind speed, wind direction, and turbulence

intensity. The conductor response data did not indicate any

specific trend for wind speed, wind direction or turbulence

intensity (terrain roughness). The upcrossing rates for

eighteen west wind records are plotted on the same graph, as

shown in Figure 24. Even though values have scatter, there

is a specific trend. A linear regression line is fitted to

the ensemble of data, as shown in Figure 24. The regression

line has a correlation coefficient of 0.93. This

correlation coefficient lends credence to the use of data as

an ensemble. Values for k and c in ecjuation 5.14 are 0.580

and 0.136, respectively.

For the Rayleigh distribution, values of k and c are 2

and 1.414, respectively. A line representing a Rayleigh

distribution is shown in Figure 24. The regression line

fitted to the field data is quite different from the

Rayleigh distribution line, indicating that the conductor

response data has a non-Gaussian distribution. It is

observed in Figure 24, that a Rayleigh distribution

underestimates upcrossing rate as compared to the field

data. The upcrossing rate plot can be used to determine the

peak factor for a desirable probability of upcrossings. The

peak factors obtained from the plot are based on 0.1 second

peak values.

133

0.000000001 T

0.000005 •

o JO CO CO

9 D "o ! »

(0 n p

0.0006

0.0111 r

0.066"

A

0.2

0.4

1.7 2.7 4.5 7.4

Peak Factor, g

Figure 24: Cumulative P r o b a b i l i t y D i s t r i b u t i o n of Upcross ings for Conductor Response

CHAPTER VI

CONCLUSIONS

The purpose of this study was to compare and refine the

analytical model to predict dynamic responses of electrical

transmission line conductors to extreme winds using field

data. Wind and conductor response field data were obtained

from a full-scale field experiment. The field data were

collected by the Bonneville Power Administration (BPA) from

an instrumented single circuit 500 kV lattice tower on the

John Day-Grizzly line 2, which is located at the Moro site

in northern Oregon. The conductor spans 252 m and 450 m on

two sides of the instrumented tower. A total of

twenty-three twelve-minute duration records were utilized in

thi s study.

Based on the analysis of field data and the refinement

of the analytical model originally developed by Davenport

(1980), the following conclusions are made:

(1) The field measured wind and conductor response data

were found to be valid. The mean -responses of three

conductors were within 10% of each other for all

records. Fluctuating responses of the conductors

showed a significant amount of scatter.

134

135

(2) Winds traversing over the valley showed a wide

variation in profile and turbulence. Winds coming from

similar terrains of valleys and hills have a power-law

exponent range from 0.11 to 0.18 and a turbulence

intensity range from 0.11 to 0.21.

(3) The wind spectra showed 99% of spectral energy in the

frecjuency range below 1 Hz. The amplitude constant A

for Kaimal's gust spectrum was found to be within the

suggested range of 0.15 to 0.60; however, the exponent

constant n from the field data was much higher than the

suggested range of 0.33 to 0.67.

(4) The field measured effective conductor force

coefficient was found to vary between 0.48 and 0.75.

(5) Noticeable resonant peaks occurred in the frecjuency

range from 0.1 to 0.4 Hz. in the conductor response

spectra. Two of these peaks close to 0.12 and 0.22 Hz

were identified as corresponding to natural transverse

frecjuencies of the conductors associated with the two

unecjual spans. The resonant response energy level was

found to be low, less than 15% of the total energy in

most records.

(6) The majority of the records showed that the field

measured background turbulence response of the

conductors accounted for 75% of the fluctuating

response.

136

(7) The analytical model for background response was

refined by determining the joint acceptance function

(JAF) coefficients from the field measured data. The

best values for the coefficients (ecjuation 5.7) are

judged to be Q=0.45 and M=0.20.

(8) Damping of the conductors, found from the field

measured data, was much higher than the theoretical

aerodynamic damping ratio. A damping value of 40% is

suggested for the conductors.

(9) The refinement of JAF coefficients and use of

aerodynamic damping factor of 40% in the analytical

model gave a significant improvement in prediction of

background and resonant responses when results were

compared with the field measured data. However, ratios

of refined analytical model values to field measured

values showed a large scatter.

(10) Many records showed response peak factors (for 0.1

second response) to be higher than the range of 3.5-4.0

suggested in the analytical model. The upcrossing rate

principle was used to determine the peak factors on a

probabilistic basis. The Weibull distribution

satisfactorily describes the probability distribution

of the upcrossings rate.

137

It is recommended that additional field data be

obtained, particularly at reasonably predictable sites, to

further verify and refine the analytical model. The

computational procedures presented here are general and are

applicable to additional field data.

REFERENCES

ANSI, 1982: "Minimum Design Loads for Building and Other Structures, ANSI 58.1-1982," American National Standards Institute, Inc., ANSI, New York, NY.

ASCE, 1984: "Guidelines for Transmission Line Structural Loading," American Society of Civil Engineers, New York, NY.

Bendat, J.S., and Piersol, A.G., 1966: "Measurements and Analysis of Random Data," John Wiley and Sons, Inc., New York, NY.

Bendat, J.S., and Piersol, A.G., 1980: "Engineering Applications of Correlation and Spectral Analysis," John Wiley and Sons, Inc., New York, NY.

Blevins, R.D., 1977: "Flow-Induced Vibration," Von Nostrand Reinhold Company, New York, NY.

Brunt,. S.D., 1952: "Physical and Dynamical Meteorology," Second Edition, Cambridge University Press, England.

Chatfield, C , 1975: "The Analysis of Time Series: Theory and Practice," Chapman and Hall, London.

Criswell, M.E., and Vanderbilt, M.D., 1987: "Reliability-Based Design of Transmission Line Structures: Final Report," EPRI EL-4793, Project 1352-2, Vol. I, EPRI, Palo Alto, California, March.

Davenport, A.G., 1960: "Wind Loads on Structures," Technical paper No. 88 of the Division of Building Research, National Research Council of Canada, March.

Davenport, A.G., 1961: "The Application of Statistical Concepts to the Wind Loading of Structures," Proceedings of Institute of Civil Engineers, Vol. 19, pp. 449-471, August.

Davenport, A.G., 1967: "Gust Loading Factors," Journal of the Structural Division, Vol. 93, ST3, ASCE, New York, NY, pp. 11-34, June.

138

139

Davenport, A.G., 1972: "Approaches to the Design of Tall Buildings Against Wind," Theme Report, Proceedings of the ASCE-IABSE International Conference on Planning and Design of Tall Buildings, Lehigh University, pp. 1-22, August.

Davenport, A.G., 1977: "The Prediction of the Response of Structures to Gusty Wind," Proceedings of International Research Seminar, Safety of Structures Under Dynamic Loading, Vol. I, Norwegian Institute of Technology, Trondheim, Norway, pp. 257-284, June.

Davenport, A.G., 1980: "Gust Response Factors for Transmission Line Loading." Wind Engineering, Proc. of the Fifth International Conference on Wind Engineering, J.E. Cermak, Ed. (Fort Collins, CO; July 1979), Pergamon Press, New York, NY.

Ferraro, V., 1983: "Transverse Response of Transmission Lines to Turbulent Winds." Masters Thesis, The University of Western Ontario, London, Ontario, Canada, July.

GAI Consultants, Inc., 1981: "Transmission Line Wind Loading Research," EPRI Interim Report, RP 1277, EPRI, Palo Alto, California, April.

Ghiocel, D., and Lungu, D., 1975: "Wind Snow and Temperature Effects on Structures based on Probability," Abacus Press, Kent, England.

Gould, P.L., and Abu-Sitta, S.H., 1980: "Dynamic Response of Structures to Wind and Earthcjuake Loading, " John Wiley and Sons, Inc., New York, NY.

IMSL, 1982: International Mathematical and Statistical Libraries, Inc., Vol. 2.

Jan, Con-Lin, 1982: "Analysis of Data for the Response of Full-Scale Transmission Tower Systems Real Winds," Masters thesis, Texas Tech University, Lubbock, Texas, December.

Jenkins, G.M, and Watts, D.G., 1968: "Spectral Analysis and its Applications," Holden-Day, San Fransisco, California.

Kaimal, J.C, 1978: "Horizontal Velocity Spectra in an Unstable Surface Layer," Journal of the Atmospheric Sciences, Vol. 35, pp. 18-24.

140

Kancharla, V.S., 1987: "Analysis of Wind Characteristics from Field Wind Data," Masters thesis, Texas Tech University, Lubbock, Texas, May.

Kempner, L., Jr., 1979: "Wind Loading on a Latticed Transmission Tower," Report No. ETKA-79-1, Bonneville Power Administration, U.S. Dept. of Energy, November.

Kempner, L., Jr., and Laursen, H.I., 1977: "Response of a Latticed Transmission Tower," Report No. ETB-77-1, Bonneville Power Administration, U.S. Dept. of Energy, November.

Kempner, L., Jr., and Laursen, H.I., 1981: "Measured Dynamic Response of a Latticed Transmission Tower and Conductors to Wind Loading," Report No. ME-81-2, Bonneville Power Administration, U.S. Dept. of Energy, July.

Kempner, L., Jr., and Thorkildson, R.M., 1982: "Moro Test Site Wind Parameter Study." Report No. ME-82-3, Bonneville Power Administration, U.S. Dept. of Energy, December.

Kempner, L., Jr., Volpe, H.W., and Thorkildson, R.M., 1985: "Wind Loading Research on Transmission Line," Proceedings of the Fifth International Conference on Wind Engineering, Texas Tech University, Lubbock, Texas, November.

Kim, Soo-Il, 1977: "Wind Loads on Flat Roof Area Through Full-Scale Experiment," Doctoral Dissertation, Texas Tech University, Lubbock, Texas.

Lawson, T.V., 1980: "Wind Effects on Buildings: Vol. 1, Design Applications," Applied Science Publishers, Ltd., London, England.

Levitan, M., 1987: Unpublished Data, Civil Engineering Department, Texas Tech University, Lubbock, Texas.

Mehta, K.C., Norville, H.S., and Kempner, L., Jr., 1986: "Electric Transmission Structure Response to Wind," Proc. International Symposium on Probabilistic Methods Applied to Electric Power Systems, Samy Krishnasamy, Ed., (Ontario Hydro, Toronto, Canada; July 1986), Pergamon Press, New York, NY.

141

Melbourne, W.H., 1975: "Peak Factors for Structures Oscillating Under Wind Action." Proc, Fourth International Conference on Wind Effects on Buildings and Structures, Cambridge University Press, London, England.

Miller, I., and Freund, J.E., 1977: "Probability and Statistics for Engineers," Second ed., Prentice-Hall, Inc., Englewood Cliffs, New Jersey.

NESC, 1984: "National Electric Safety Code," ANSI C2, Institute of Electrical and Electronics Engineers, Inc., Wiley-Interscience, New York, NY.

Nigam, N.C., 1983: "Introduction to Random Vibrations," The MIT Press, Cambridge, Massachusetts, London, England.

Norville, H. S., K.C. Mehta, andA.F. Farwagi, 1985: "500 kV Transmission Tower/Conductor Wind Response: phase I," Institute for Disaster Research, Texas Tech University, Lubbock, Texas, May.

NRCC, 1980: "National Building Code of Canada," NRCC No. 17724, Associate Committee on the National Building Code, National Research Council of Canada, Ottawa, Canada.

Potter, J.L., and Boylan, D.E. 1981: "Aerodynamic Force Coefficients for the Prediction of Wind Loads on Electrical Transmission Lines," Sverdrup Corporation, Report submitted to EPRI, Palo Alto, California.

Pries, R.A., 1981: "Moro 1200 kV Structural/Mechanical Test Program," Report No. ME-81-1, Bonneville Power Administration, U.S. Dept. of Energy, March.

Racicot, R.L., 1969: "Random Vibration Analysis-Application to Wind Loaded Structures," Report No. 30, School of Engineering, Case Western Reserve University, Cleveland, Ohio, February.

Rice, S.O., 1945: "Mathematical Analysis of Random Noise," Bell System Technical Journal, Vol. 24, pp. 46-156.

Sachs, P., 1978: "Wind Forces in Engineering," Second Edition, Pergamon Press, New York.

Simiu, E., and Scanlan, R.H., 1985: "Wind Effects on Structures: An Introduction to Wind Engineering," John Wiley and Sons, Inc., New York, NY.

142

Statistical Analysis System (SAS), 1985: "SAS User Guide: Statistics," SAS Institute, Inc., Gary, North Carolina.

Symon, R.K., 1961: "Mechanics," Second Ed., Addison-Wesley Publishing Co., Inc., Reading, MA, May.

Twu, S.R., 1983: "Gust Response Factor for Transmission Line Structures," Masters Thesis, Texas Tech University, Lubbock, Texas, August.