Vous êtes sur la page 1sur 39

IEE REVIEW

Overhead linessome aspects of design and construction


J.F. Adam, J. Bradbury, B.Sc.(Eng.), C.Eng., M.I.E.E., W.R. Charman, C.Eng., M.I.Struct.E., G. Orawski, B.Sc.(Eng.), C.Eng., F.I.E.E., Sen.Mem.I.E.E.E., M.J. Vanner, M.A., C.Eng., M.I.E.E., M.lnst.P.
Indexing terms: Overhead lines, Cables and overhead lines Abstract: Overhead lines embrace so many engineering disciplines, that only some aspects of design and construction can be considered in the paper. The new trends in assessing design loads for the components are briefly outlined, together with their interrelation with the strength of the components. It is argued that ultimate loading conditions should replace 'working loads and factors of safety'. Although conductors have been fairly well standardised, new techniques for assessing their performance have recently been developed. However, with the advent of conductors incorporating optical fibres, some basic investigations are required. While there is a better understanding of wind induced phenomena, there is still no universal technique or device to solve galloping problems. After the conductors, the supports represent the second high-cost component of an overhead line, and much of the routine of design can now be performed by computers. Depending on requirements, loadings, ecological factors, economics etc., the supports can adopt many shapes and sizes, from the single member pole for low, medium or even high voltage lines, to the complex lattice types normally reserved for HV or EHV lines. The availability of fewer rolled sections and a limit on larger sizes tax the ingenuity of designers, who have to use, occasionally, compound sections or solid round members. Testing support designs is still a fairly common requirement. Considerable advances have been made in the knowledge of the electrical performance of air gaps, but the problem of insulation co-ordination still remains. Possibly, the greatest innovation in the field of insulators is the improving confidence in the use of composite insulators. The foundations of an overhead line are subjected to loadings seldom encountered with other civil engineering works, i.e. uplift. A review is made of the various shapes and sizes which can be used, together with up-to-date techniques of design. Site investigation is gaining more importance and more acceptance, in view of the need to match designs to actual site conditions with the dual purpose of economy and reliability. The influence of foundation movement is also discussed. Regarding installation, the equipment is adapted to the type of construction. It can be fairly simple for distribution lines and may become very sophisticated for transmission lines, especially when acceleration in the progress of work is desired. In addition, consideration of logistics, transportation and access have become exceedingly important. Work in highly inaccessible terrains, or opposition to access road building for ecological reasons, has focused attention on the application of helicopters for overhead transmission-line construction.

General introduction

Any reader assuming that this paper can replace a textbook on overhead-line design and construction is likely to be disappointed. On the other hand, those readers who may be wondering how overhead lines, which are now a common feature in any developed country, have been designed should hopefully find some answers to their questions. They will also discover that overhead lines require multi-discipline knowledge and training, and that the techniques of design and construction are not static. This paper may be considered as a companion to the review published earlier by the IEE [1]. The earlier paper describes lines up to 132 kV, but 132 kV is now a distribution voltage in the UK. Hence it can be considered that Reference 1 covers aspects of distribution, whereas this paper will be biased towards HV design and construction; with only short statements regarding distribution practice, for the sake of completeness. Unfortunately, it is impossible to cover fully and comprehensively all the aspects of transmission by overhead lines in such a review, hence it is only proposed to outline some aspects of the design and of the construction. It will be assumed that the system designer has done his work and that the stage has been reached when a specific overhead line is needed between two points. Notwithstanding the availability of codes, standards, specifications etc., the analysis of performance, the use of extrapolation of experience leads to a situation where most aspects, i.e. most components, are under constant review to ascertain that
Paper 3291C, (P8), first received 1st August 1983 and in revised form 21st March 1984 The authors are with the Power Transmission Division, Balfour Beatty Power Construction Ltd., 7 Mayday Road, Thornton Heath, Surrey CR4 7XA, England

reliability can be achieved within economic constraints, while drawing maximum advantage of the strength of the materials used. As overhead lines are exposed to all weathers, a continuing analysis is being made of the performance of lines, to assess whether they have been properly and reasonably designed for the environment in which they are located. Thus, it can be found that design bases, which had been judged adequate for many years, are reviewed from time to time. Such a situation has frequently arisen in many parts of the world, and not only in the UK, when in December 1981 and January 1982, many areas were deprived of electricity for varying periods of time [2]. As a result, most countries have set up teams of technical experts whose responsibility it will be to draft new codes and new specifications. Hopefully, these will be based on generally agreed uniform basic principles; although the quantitative values may vary from area to area. Those principles are most likely to be based on probabilistic techniques proposed by such international organisations as CIGRE or IEC, whose activities are followed by representatives from the corresponding British committees through BSI. The prime purpose of transmission lines is to transmit electricity from one place to another as cheaply and as reliably as possible; consequently, the conductors could be considered as the most important item; all the other components acting as a complex support system. Conductors represent 20 to 40% of the installed cost of a line, and it is not surprising that many designs have been evolved. A fair selection is thus available to satisfy both electrical and mechanical requirements. A solution to the vexing issue of inelastic elongation of aluminium-based conductors has now been found, and guesswork is almost entirely elimi149

IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

nated. Through the availability of computers, exact shapes of suspended conductors (catenary as opposed to parabola) can be readily analysed. However, manufacturers are still introducing new concepts, and one of the most important recent developments has been the incorporation of optical fibres in the core of conductors, thus providing new means of communications: be it data transmission or speech. While the principles of most windinduced phenomena (aeolian vibration, subconductor oscillation for bundle conductors, galloping) are now properly understood, there is, as yet, no universal solution to the galloping problems. Obviously, an adequate selection of conductor fittings has a critical importance so as not to impair the life of the conductor itself. For many years, porcelain and glass were the only materials used for the manufacture of insulators. These insulators are still used for the vast majority of installations, but a serious contender, referred to as 'composite insulators', has appeared on the market, and is slowly gaining acceptance. Special test methods are urgently being developed in the hope that such tests will be helpful in the assessment of the comparative long-term performance of the various materials entering into the manufacture of these new insulators. Composite insulators have many favourable mechanical characteristics, and they have already been used for the design of compact lines where they can become structural elements. In recent years, worldwide, there has been a considerable reduction in the availability of the range of steel sections, which is taxing the ingenuity of the designer. Another problem which the overhead-line engineer has to face is the increasing opposition of the public to the installation of structures in the countryside. Thus, new concepts and new silhouettes are constantly investigated. Occasionally, single-member supports have been installed, not only for low-voltage, but also for very-high-voltage lines, and a constant search is being made into the application of alternative materials to steel, for fabricating new supports. Many lines installed in ice-free zones must consider the loadings due to installation and maintenance, because they may prove critical for some structural members. Overhead-line supports of new designs are normally subjected to a type test, i.e. they are tested before bulk fabrication commences. The main change for the foundation designer has been the ever growing importance of ground investigations. Foundation loads have increased with the increase in transmission voltages: besides, the lines are frequently installed in inaccessible areas where unusual foundation problems can be encountered. It must be remembered that the vast majority of tower foundations are designed for uplift, a condition seldom met in civil engineering activities. There has been a swing away from the classical padand-chimney foundation towards augered shafts, small piles or rock anchors, when the ground is suitable. With the shift of industrial growth rate towards nations of the Third World, overhead lines have to be installed in new territories where there is little past experience. A period of adaptation was therefore necessary for the construction engineer when he had to accept the challenge of construction in areas where there is lack of communication, of adequate roads and bridges, and where transport is a problem. Thus, the construction engineer must not only develop special equipment to maintain required rates of progress, but has to evolve all the logistics of a complex operation involving the installation of very many disparate items, to ensure that they are all at the right place, at the right time and in the right quantities. In some cases, access
150

is so difficult that the choice lies between the construction of new roads or the use of helicopters. The later aspect has in fact been developed, and such a degree of sophistication has now been introduced that they can be used for the transport of labour, materials and also for the actual erection of towers or for the stringing of conductors, quite often at competitive costs. From its initial inception to its final completion, an overhead line will require the services of many organisations. In broad terms, the contractor will be working to specifications prepared by others, such as consultants or clients. These latter organisations will be responsible for determining the basic requirements, such as conductors, insulation levels, tower configurations, loads, combinations of loading cases, clearances etc. The overhead-line contractor will be responsible for translating these parameters into a proper design, and he must have sufficient technical knowledge and experience to be able to comply with the specifications and to fill the gaps, if any, in these specifications. These aspects of relationships are covered in a recent IEE paper [3].
2 2.1 Loadings Introduction

Every engineering construction is subjected to loads imposed by nature, or by man. A knowledge of the phenomena which give rise to such loads, and an assessment of the influences of the phenomena on particular constructions, or part of those constructions, is of paramount importance. Whereas loads imposed by man, such as floor load density in a building, can be defined and controlled, the natural influence, such as wind pressure on a building, can only be estimated. As engineering activities always involve at least two parties, the client and the contractor, early agreement on the design principles is very desirable. Hence, sets of statutory regulations are evolved in each country which form the technical basis of the relationship between the parties. In many cases, if questions are raised regarding the basic principles from which loads have been derived, answers are not readily forthcoming. It is a fact that most countries have been systematically collecting loading data and checking the performance of structures under the influence of various loads. Thus, the basic information should be available, but it must be accepted as a guiding principle that most engineering decisions are made in anticipation of worst probable and not worst possible climatic or other conditions. If the latter were used as a design basis, there would seldom be any failures. Such an approach would carry heavy financial penalties, and, in any case, who could fix a value to a worst possible event as a unit can always be added to any value? In theory, such a value could be increased to infinity. Overhead transmission lines, by their very nature, are built across the countryside and are therefore subjected to climatic (or meteorological) load influences, such as wind, ice and temperature, which can vary in time and in space. In addition, they can also be subjected to varying load influences due to geological effects, such as land slides, subsidence, earthquake, floods etc.
2.2 Statutory regulations in the UK

In the UK, the design of overhead lines is now based on 'Statutory instruments, 1970, number 1355' [4]. These regulations supersede the previous ones published in 1947 [E1.C.53 (1947 revised)] and incorporate the results of experience gained up to 1970. While there has been little
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

change in the loadings themselves, there have been some changes in other paragraphs [5] affecting 'margins of strength' and clearances. So far as loadings are concerned, the regulations differentiate between two types of construction, depending on conductor size and voltage level: (a) For conductors up to 35 mm2 and voltages between 650 V and 33000 V, an equivalent wind pressure of 760 N/m 2 is applied on conductors only (no ice to be considered, and wind influence on supports, insulators and equipment to be disregarded). (b) For all other constructions, an equivalent wind pressure of 380 N/m 2 is applied on conductors, supports and insulators, but, in addition, an augmented mass is specified together with an augmented diameter, which is deemed to represent the effect of icing on the conductors: (i) for voltages up to 650 V, this increase in diameter and in mass can be deemed to be equivalent to a radial ice thickness of 4.75 mm (ii) for higher voltages, the increases are deemed to represent a radial ice thickness of 9.5 mm. Young engineers please note: these are conventional loadings and there is no suggestion that the interstices in the conductors should be filled with ice! Besides, the ice deposits on conductors seldom conform to a perfect cylinder, and the regulations only attempt to help in mathematical modelling of a complex phenomenon. For distribution lines with small conductors, these design parameters have been accepted as the norm. However, for transmission lines, practice has led to a situation where these statutory regulations are considered as minimum requirements. Thus, the CEGB has adopted higher loadings on conductors and towers than those specified: e.g. the augmented mass and the augmented diameter of the conductor are equivalent to a radial ice thickness of 12.5 mm, and the wind pressure on 1.5 faces of the supports (lattice structures) is taken as 1200 N/m2. In the 'Statutory instruments' two other new features are worth mentioning: (a) The expression 'factor of safety' is not used. Instead, 'stress limitations' are given for materials, with reference to statutory loadings. (b) Attention is focused on the need for maintenance. The critical reader may well ask, logically, why is it that two lines built in the vicinity of each other should suffer different loading conditions? Nature is not aware of the fact that lines of different voltages and conductor sizes should not be loaded equally. The simple answer is that, implicity and intentionally, different reliability levels are artificially introduced by the specification writer, recognising the varying duties of the lines. Actually, the new trends in design, mentioned in the following text go some way in helping the specification writer to choose different loadings, by considering the present-day worth of a line including installation, repairs, loss of revenue etc. In fact, as a result of the Department of Energy/ Electricity Supply Industry Inquiry following the overhead-line failures, mostly on distribution lines, due to the blizzards in December 1981 and January 1982, the 'Statutory instruments' of 1970 will be revised, accepting, most probably, the concepts of ultimate loads and sequences of failures described in Section 2.3. 2.3 Statutory regulations in general The UK regulations on overhead lines, compared with those published in other countries, are exceedingly short and concise.
IEE PROCEEDINGS, Vol. 131, Pi. C, No. 5, SEPTEMBER 1984

In many countries, the regulations can be very extensive and can sometimes become almost equivalent to a designer's guide. It must be appreciated that regulations go beyond the question of loadings to be applied to the construction; they are conscious of the safety of the public, and give guidance on clearances, warning notices, precautions against access, earthing etc. In fact, one of the problems with regulations, in many countries, is that they were originally compiled many decades ago for overhead lines of relatively simple construction, and they are frequently lamentably out of date when applied to the present higher-voltage transmission lines. This comment applies to many of the world's major countries! In industrially developed countries, the regulations reflect many years of experience, when performance and costs could be analysed separately and as a function of each other. However, when constructing lines in new territories where experience on performance is either lacking or nonexistent, there is a need to formulate general design principles. Such work has been undertaken by the IEC (International Electrotechnical Commission) [6] based on probabilistic techniques. The randomness of natural phenomena has been emphasised recently, when many electricity supplies in Europe were interrupted, for varying periods of time, due to unusual ice loadings. The UK was not spared, and many readers may well remember the disruptions which occurred on the distribution system in December 1981 and January 1982 [2]. Such havoc had not been experienced for many years. 2.4 New trends in overhead-line design 2.4.1 Basic principle: Every engineering activity must accept the fact that no phenomenon can be described by a single value, whether dealing with loads or with strength or any other aspect. Until recently, the trend has been to assess a maximum value of load and to design a component with respect to its minimum strength. As there was some degree of uncertainty in both values, factors of safety were introduced. An interesting development resulted from this approach: the goods supplied to a client are normally subjected to testing, and the value accepted for testing is the minimum ultimate strength, which is compared with the specified load multiplied by a factor of safety. Why not specify ultimate loads in the first instance? Unless loads and factors of safety were perfectly understood, there would be hidden dangers in extrapolating data from one meteorological zone into another. Now, most engineering constructions are designed to resist loads. In simple terms, as long as the strength is greater than the load, there will be no failure. Thus, the problem is reduced" to that of assessing the probable load acting on a component and the probable strength of that component. Hence, probabilistic techniques can be used. 2.4.2 Loadings in ice-free areas: In areas not subjected to ice loading, the only meteorological influences to consider for the mechanical design of overhead lines are wind and temperature. The wind affects all the components of a transmission line, whereas the temperature (or rather temperature range) affects mostly the conductor behaviour. Measurements of wind speeds have been carried out for many years in many countries, and some measurements
151

are more reliable than others [7]. An analysis can, therefore, be performed with the reliable data, to evolve mathematical laws which could be used with confidence. It was found that a law of largest extreme distribution type 1, Gumbel, could be accepted, for the distribution function of extreme wind speeds, see Fig. 1 [8].

Table 1: Probability of wind speed exceeding at least once the maximum wind speed in period T Line life Return period 7", years
N 25 50 100 150

7"=25 f = 5 0 0.63 0.87 0.98 0.998 0.39 0.63 0.87 0.95

7=100 0.22 0.39 0.63 0.78

7"=150 0.15 0.28 0.49 0.63

7=500 0.05 0.095 0.18 0.26

of 0.63 that it may be subjected at least once to loads in excess of those which have been accepted for design. In fact, this is equivalent to stating that the yearly probability of a wind speed exceeding X in any one year is 2% (1/50). 2.4.4 Wind speeds: It is now accepted that Gumbel (type1) largest extreme distribution can be used to represent the cumulative distribution of extreme wind speeds recorded in any year. Mathematically, it is expressed as r {

Fig. 1 Probability density function of yearly maximum wind speeds (extreme values) f(v) = a exp {a(v u) exp [ a(v u)]}
where u = v - 0.45005<r

The averaging time for the measurement of the wind speed is important, in as much as it affects factors used for calculating wind loads [9]. Fig. 1 illustrates the probability density function of yearly maximum wind speeds, obtainable from meteorological offices, assumed to be expressed by a law of extremes. As pressure is proportional to the square of the wind speed, a probability density function can be prepared for each component, by means of a transfer function. In fact, it has been shown recently that the distribution function of the squares of maximum yearly wind velocities is closer to the Gumbel type-1 law than the distribution of wind speeds. 2.4.3 Return period: This concept is useful for selecting the wind load to be applied to a component. The return period can be defined as the average period of time between the recurrence of a phenomenon (e.g. wind speed), whose value will be equal to or exceeded once in N years, where N is the life of a line. Typical values for N would be N = 25 years for distribution lines = 50 years for transmission lines = 150 years for important lines and major links etc. Statistically, return period T and probability p are related by Tp = 1. Now, Reference 7 states: the probability that a value less than or equal to a specific wind speed will occur in any one year is q = (1 p). In a period of N consecutive years the probability that a value less than or equal to X will occur at least once is Q = qN. Hence the probability that a value greater than X will occur at least once in a period of N years is p = 1 - Q = 1 - q = 1 - (1 - pf
N

= exp - e x p j - f

C ( - fV - V

(2)

where Cx and C2 are constants depending on the number of years for which data are available, av is the standard deviation of statistical data. Now the probability that a wind speed V is exceeded at least once in any one year is p = \ - F(V) and if data are available for an infinite number of years, the law has the form: p = 1 - exp (V -V + 0A5av)
(3)

where V = mean value of recorded wind speed data V = calculated wind speed having a probability p of being exceeded av = standard deviation of recorded wind speed data Another useful concept for statistical analysis is the 'coefficient of variation', which is defined as the ratio between the standard deviation of recorded data and the mean value of that recorded data, i.e. oJV. For Europe, 0.12 < aJV < 0.2. Remembering the relation Tp = 1 where T is the return period (see Section 2.4.3), then p = 1/T. Thus, the above eqn. 3 can be solved for p = = 0.04 for 25 year return 25 p = = 0.02 for 50 year return etc. 2.4.5 Wind pressures and wind loads: Once a wind velocity is selected, the reference pressure is given by where PR is the reference wind pressure (in Pa or N/m2), p = air density (1.225 kg/m3 at 15C at an atmospheric pressure of 1013 mbar). VR = reference wind velocity (in m/s), which is equal to the product of ground roughness coefficient by the velocity V from eqn. 3. It is important to note that the averaging time of the wind velocity is critical, as most factors for calculating the
1EE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

(1)

If we accept a life for the line of N = 25, 50, 100 or 150 years, and different periods of return, the following probabilities of occurrence of a wind speed exceeding at least once the maximum wind speed in that return period are given in Table 1. Table 1 shows that if a component, (such as a tower) whose assumed life is 50 years, has been designed for a wind speed with a return period of 50 years, there is a probability of 0.63 that the wind speed will be exceeded at least once during its 50 year life; i.e. there is a probability
152

loads will be related to this averaging time: e.g. Reference 9 uses an averaging time of 2 s, whereas Reference 6 uses 10 min. The resulting design loads calculated by both References are the same. To transform the pressures into loads, they must be multiplied by suitable drag coefficients, gust factors and the areas affected. The general equation giving the load L on a component takes the form: L = PR CGA where PR = reference wind pressure as given in eqn. 4. C = drag or pressure coefficient depending on the shape of a particular element and on whether the air flow is supercritical or subcritical, e.g. for conductors and supercritical flow C = 1.2, but for subcritical flow C = 0.6 to 0.8. For convenience Reference 6 accepts C = 1.0 for conductors. G = gust factor, allowing for the turbulence of the wind. It depends on the dynamic response and the height of the element. Values are normally given by suitable graphs related to the averaging time of wind speed. A = area affected by wind. 2.4.6 Ice loads: While wind data (sometimes of dubious accuracy) can be obtained from most parts of the world, as, in addition to meteorological stations, all airports have anemometers, data on ice accretion is very scanty and highly unreliable. Observations of failed lines may give indications as to the minimum loads which caused failures, but do not give information on the actual loads, which may be well in excess of those calculated on the basis of the known strength of components which failed. In some parts of the world, ice loads are critical for the design of the lines: this is fully acknowledged, and IEC TC11 has formed a working group (in which there are two UK members), which has been given the task to draft recommendations for the assessment of: (a) ice loads in the absence of wind (b) loads resulting from a combination of wind and ice. Analysis of whatever data were available has led to the conclusion that ice overloads could be mathematically expressed by Gumbel's law (type-1) of largest extreme distribution. Ice loads will be assessed as an additional mass on the conductors. The problem is further complicated by the fact that the density of ice can vary between 0.4 and 0.9 g/cm3 depending on the process of ice accretion (a typical value for use in design suggested by the UK statutory regulations is 0.91 g/cm3). There are two processes of ice accretion: (i) Precipitation icing (a) Glazed ice is formed by supercooled rain or drizzle on objects at or below freezing. It is associated with temperature inversion. (6) Wet snow is snow which has begun to melt as it falls through a relatively warm layer, and will stick to obstacles in its path, (ii) Incloud icing Rime is formed by supercooled water droplets in suspension in the clouds. This is a characteristic of sites in exposed areas above the cloud base. 2.4.7 Ice in the absence of wind: In the absence of wind, ice loads have a dual influence on the designs of overheadline components: (a) they increase all vertical components on the structure
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

(b) they can give rise to unbalanced tensions in the conductors, which can influence the required strengths of the structure (torsional and longitudinal). Data on ice loads should preferably be collected by icemeasuring stations, and already some agreement has been reached as to their instrumentation and methods of operation. When such data are available, a reference ice load can be determined by accepting the 'law of extremes' for a selected return period. When statistical data are not available, the overheadline engineer can only attempt to assess the value of the maximum ice load and recognise the lack of accuracy of his information by introducing fairly large values of the standard deviation. So far as the conductor performance is concerned, the investigations should assume that ice loading occurs at a temperature of 5C, which represents a consensus of opinions, as lower temperatures can occur. It was concluded that this would give a reasonable design basis among the many loading conditions to be considered [6]. 2.4.8 Ice with wind: While this loading condition is that which comes almost automatically to mind, it is in fact the most difficult to define because of lack of correlated data. Very little thought is required to appreciate that a combination of maximum wind (albeit resulting from statistical analysis) and of maximum uncorrelated ice (also obtained by one of the methods in the preceeding Sections) is likely to lead to expensive overdesigns. Actually, it could occasionally lead to dangerous designs, because many members suffer stress reversals under different loading cases. What is really needed is information on wind during icing weather. This is not available on an international basis. However, to help designers to fulfil their duties some recommendations have been advanced by the IEC working group. Such recommendations are not entirely arbitrary as they represent the consensus of opinions of engineers who have seriously studied the problems, and have analysed the performance of actual lines. For the purpose of design, the additional mass of ice load is transformed into an equivalent annular accretion to calculate loads due to wind effects. Again, the design temperature is suggested as being 5C. One aspect, which has not yet been considered, is that of icing of supports. Cases have been recorded where ice can fill all the gaps in a lattice structure, transforming the latter into a solid obstacle whose drag coefficient and area would be different from those of a lattice tower. Fortunately, icing occurs mainly in those countries which already have some experience on line performance at most voltages. The engineer can then make inquiries and decide whether such loading conditions should be investigated. Especially with icing, the influence of 'microclimate' is most important. Hopefully, a situation can be reached where loadings will be known reasonably accurately, and the components of a line will be designed to suit; thus avoiding the temptation to which engineers have occasionally submitted, of modifying loadings to permit the use of lighter components. 2.5 Strength of components It is common practice when purchasing materials to specify minimum guaranteed breaking (or failing) strengths. All engineers will, however, agree that this is not
153

a unique value describing the strength of the component; hence there is a probability that many items of the same component will have greater strength than the minimum guaranteed and a few may have lower strength: hence, the concept of statistical distribution of strengths. For convenience, the distribution is assumed to be normal (or Gaussian) (see Fig. 2). Obviously, any manufacturer can

1.00 0.84 F(x)=y= J-oo 0.50

cumulative distribution function ^ "

ing components in such a way. However, a relative comparison of the areas will be helpful in assessing which component of a system is likely to fail first. While the subject of the calculation of the reliability of a line is still under discussion, there is some agreement on the principles of relative designs of components: (a) the load is selected for a predetermined return period (b) the strength distribution curve is so located that there is a selected (usually 90%) probability that the strength of the component will be greater than this value (see Fig. 3). The reliability of an actual line is much more difficult to calculate, because it is a function of many influences, such as the size of the loading event (e.g. one span or several spans) and the degree of utilisation of each component, i.e. the 'use factor'. The easiest way to visualise the use factor is to imagine a tower designed for a span / but used actually on a span / t < /; in which case the tower is not fully utilised. The risks of failure are different for a tower fully utilised and for a tower which is underutilised. The technique illustrated by Fig. 3 can be used for the design of any component. Hence, it should be possible to design the various components so that their strengths are co-ordinated to give a desirable sequence of failure, i.e. one component is intentionally stronger than another. If the mean strength R of each component were known, then the 90% withstand value could be estimated from

r
| /

0.1 6
0

probability density """"v function

strength /I / characteristic /

y
X

R90 = K(l-\.2XVe)
where R = mean value of strength (or resistance) (While this value can be easily found for, say, a batch of insulators, or individual steel bars, it is difficult to give a unique value to a compound structure such as a tower.) Vc = coefficient of variation = oR/R. Similarly
R95 = 1.645Ve)

Fig. 2

Normal distribution

choose a minimum guaranteed strength so far below the average strength that the probability of failure under test is negligible.

2.6 Design of components and strength co-ordination


The preceding Sections have shown that neither the load nor the strength are deterministic, clearly defined, values. Hence probabilistic techniques can be used for comparisons of relative performances. Fig. 3 illustrates the basic approach. Let f(L) be the probability density function of extreme loads and P r (L) the cumulative distribution of strength (or resistance). It is accepted that the shaded area A is a measure of the risk. By designing components so that their cumulative strength distribution is well to the right, this area will obviously decrease, but there will be an economic penalty in design-

i.e. each engineer could modify the position of the strength distribution with respect to the load. Fig. 4 gives the calculated probabilities of failures (area A from Fig. 3) for different coefficients of variations of loads (assumed to be of the extreme type 1) and of resistances (assumed to be Gaussian) for the situation illustrated by Fig. 3. Typical values for coefficients of variations are given in Table 2.
Table 2: Coefficients of variation for OHL components (in %) Components Conductors Towers Wood poles Foundations Insulators

cumulative distribution function of strength (eg. tower) p ((J T probability density function ' of extreme loads
0.5

COV (V:c )

1 to 5

5 to 10 10 to 25

10 to 30

2 to 8

0.10 load corresponding to / selected return period (eg. 50 years) area proportional to risk = A A= fp(L)f(L)dL J 0 load and strength

Fig. 3 154

Principle of risk assessment

In this selection many aspects come into play, such as coefficients of variation, impact of failure on adjacent components, use factors, ease of repair and maintainance etc. However, for distribution lines, on wood poles or on light supports, it would be sensible to accept the conductor as the weakest component, provided that supports and foundations remain serviceable. Thus, in the case of an overloading event, storm or ice, the conductors would fail, and repair work could be performed soon after the unusual loading event. By applying the reasoning to its conclusions, in some cases, the use of 'mechanical fuses' may well be contemplated. These would act as failure limitation devices. Such
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

devices have been used in the past, to limit cascading failures. One example would be that of a crossarm connected

pf = probability of failure of that component under the same loading event, measured by the area A in Fig. 3. Hence, preference should be given to the expression 'reliability', which is. a quantifiable value giving the probability of survival.

0.04

Conductors and earthwires

0.03

0.02

0.01 conductors towers 1-5/. 5-10%. insulators 2 - 8% 0.05 0.1 0.15 0.2 0.25 0.3

0.005

foundations 10 - 30/.

3.1 Introduction While it is customary to use the above two expressions, they both describe the same physical component. They are normally differentiated by their function and their position. Economically, conductors represent between 20 to 40% of the total cost of a line, consequently their selection is of prime importance. Actually, the whole concept of electricity transmission or distribution revolves around these elements. In the early days of electrical power transmission, when voltages of 11 kV or 33 kV were considered high voltages, copper was mainly used as the material for overhead-line conductors. In the meantime, with the expansion of electricity networks, several factors, such as price, weight, availability and conductivity, have virtually compelled overhead line engineers to concentrate on aluminiumbased conductors, e.g. AAC = all aluminium conductor ACSR = aluminium conductor steel reinforced (which was also known as SCA, steel-cored aluminium) AAAC = all aluminium-alloy conductor ACAR = aluminium conductor alloy reinforced. The alloy used in the construction of AAAC and ACAR is a heat-treatable aluminium alloy containing magnesium and silicon. In areas of European influence, approximately 0.5% magnesium and silicon are added, while, in areas of North American influence, a slightly mechanically stronger alloy is produced by the use of up to 0.9% of these elements. All conductors suffer from corrosion, which is more pronounced with bimetallic construction when an electrolytic cell can be developed under the right conditions, between the aluminium and the zinc coating of the steel core. It is now fairly common practice to grease the inner strands of ACSR as a corrosion protection. Occasionally, the steel core can be manufactured from wires coated with aluminium (aluminium-clad steel wires), but, in highly corrosive areas, such as the UK, it is desirable to grease the inner layers. Particular care, however, must be exercised in the selection of the grease to be used. With monometallic conductors (AAC, AAAC), the problem of bimetallic corrosion does not arise, and cheaper fittings can be developed, especially when no grease is used. In the UK, earlier distribution lines were installed with copper conductors, but there has been a gradual shift towards aluminium-based conductors, and nowadays copper is rarely used. For the grid and the supergrid, the CEGB have standardised on three types of ACSR, but now a new family of CEGB designs based on AAAC and ACAR are being introduced. Table 3 gives an indication of the conductor types used on CEGB lines. A recent worldwide inquiry (Table 4) gives an indication of the quantities of conductors used in a year. Table 5 lists some of the possible constructions covered by various international specifications, giving some of the mechanical characteristics which are relevant in calculating their performance.
155

-2R

Fig. 4 Probability of failure as function of coefficient of variation of components For P(R >Q) = 90%
COV = coefficient of variation aR = COV of resistance of components R
a

Sr = C O V of loads

to a tower through a shear pin which would fail before the tower body. This is equivalent to the selection of a sequence of failure of elements in a structure. References 10, 11 and 12 treat the problem of loadings and strengths in greater detail. The concepts described in the preceding text are not new. They have been used for some time in the design of concrete structures. Although the equations appear difficult, most of them can be solved easily by any calculator with mathematical functions. Computers make their manipulations even easier. The reader should appreciate that there is a tremendous difference between the probability of failure of a component and the probability of failure of a line. In an ideal situation, if all components were equally loaded, the probability of failure of the line would be that of the weakest component. Finally, the concept of failures is anathema to most engineers. It must be remembered that the reliability of the line is the complement to 1 of the probability of failure, i.e.
Ps = 1 - Pf

where ps = probability of survival of a component subjected to a loading event


IEE PROCEEDINGS, Vol. 131, Pi. C, No. 5, SEPTEMBER 1984

Table 3: Conductors used by CEGB


a Conductor construction Conductor code Equivalent cross aluminium section mm 2 70 175 400 500 700 425 Type of conductor Wire diameter mm Number of wires steel aluminium aluminium alloy

Table 5: Conductor characteristics a ACSR Number of conducting wires 6 8 18 37 61 37


9

Number of reinforcing wires 1 1 1


3

Breaking length UTS/weight/ unit length km 9.0 9.3 7.3 10.1 8.5 11.6 12.4 8.7 9 . 1 9.5 7.5 7.5 7.8 8.3 7.2

Young's modulus kN/mm 2 79.0 98.0 66.0 88.0 75.0 105.0 108.0 72.6 75.5 80.0 58.8 61.0 60.8 69.0 60.0 108.0 117.7 78.5 66.7 66.5

Coefficient of linear expansion


*10-6/OC

Horse Lynx Zebra Collybia Araucaria Totara

ACSR ACSR ACSR ACAR AAAC AAAC

2.79 2.79 3.18 3.37 4.14 4.14

7 7 7

12 30 54 24

19.1 16.9 21.2 17.1 19.8 15.3 15.0 19.6 18.9 17.8 21.2 20.9 20.9 19.3 21.7 15.0 14.2 18.0 19.4 18.3

b Use of conductor Line voltage kV


132

Phase conductors conductors/phase 1 1 2 2 Lynx Zebra Lynx Lynx

Earthwire Horse Lynx Lynx Lynx Lynx Lynx Lynx Lynx Zebra Zebra Zebra Totara Lynx

6 12 14 24 26 30 42 45 48 54 72 14 16 30 54 66 19 19 19 19 19

275

2 Lynx 1 Araucaria 2 Zebra 2 2 4 2 2 2 Zebra Zebra Zebra Araucaria Araucaria Collybia

275/400
400

12.5 12.8 9.7 8.6 14.9

Note: Breaking length is defined as the length of conductor which when hung vertically will cause failure due to its own weight. It is a useful guide to the strength of the conductor. b AAC and AAAC Breaking length UTS/weight/unit length Toung s AAAC modulus AAC km kN/mm 2 km 5.7-5.8 5.6-6.1 10.5 10.4 10.4 10.3 10.3 61.8 59.0 56.0 56.0 53.9 53.9 Coefficient of thermal expansion

Table 4: Types of conductor used throughout the world Conductor type Total length of conductor km/year 145512 16655 2310 2290 845 620 168232

Number of wires in the construction 3


7 19 37 61 91

xio-6/C

ACSR (aluminium conductor steel reinforced) AAAC (all aluminium alloy conductor) AAC (all aluminium conductor) AACSR (aluminium-alloy conductor steel reinforced) ACAR (aluminium conductor alloy reinforced) other types total

5.6-6.7
5.6-6.4 5.5-6.4 5.4-6.2

23 23 23 23 23 23

Extracted from the results of CIGRE (Working Group 05 Study Committee 22) questionnaire on stringing problems. To be published in Electra.

Note: The variation in breaking length shown in the above Table for AAC is due to the variation in strength of the wires due to the degree of work hardening received during drawing. In the case of AAAC manufactured to ASTM specifications, the breaking length would be slightly longer, as the alloy used is slightly stronger than that called for by IEC and European specifications.

As conductors and earthwires are exposed to all the weather conditions they are subjected to many windinduced phenomena. The advantages of transmission of data by optical fibres have been recognised for some time, and, recently, they have been incorporated in several lines.
3.2 Phase conductors

The simplest expression for surface gradient is Q 2ne0 r where E = surface gradient, V/cm Q = electrical surface charge per unit length, C/m r = equivalent smooth conductor radius, cm e0 = permitivity of free space = l/(367i x 109) F/m. It will be seen that an increase in radius means a reduction in gradient. Actually, an increase in equivalent radius also leads to a reduction in inductance and an increase in capacitance, which is advantageous from stability considerations. This explains why designers have considered expanded conductors (i.e. conductors with artificial voids), and bundle conductors, a common feature of lines above 220 kV; the latter being the preferred solution. The breakdown strength of air is approximately 31 kV/cm peak, or 22 kV RMS/cm. Thus a useful guide for the selection of a conductor diameter is that the voltage
1EE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

3.2.1 Basic selection parameters: The conductors fulfil an electromechanical function, hence both the electrical and mechanical aspects come into play [13]: (a) Electrical parameters: The most important parameter affecting the choice of a conductor is its resistance, because it influences voltage regulation, power loss and rating, i.e. ability to transmit current. For AC lines, the diameter of a conductor affects the inductance and the capacitance. Up to a voltage of 132 kV, the above considerations are generally adequate; however, at higher voltages, the voltage gradient on the conductor surface may require the selection of a conductor on the basis of its diameter.
156

gradient must be less than 18 kV RMS/cm. This provides a margin for atmospheric pollution effects which may give rise to corona at lower levels, and for inaccuracies in mathematical expressions. Corona is a phenomenon closely related to the breakdown strength of air, which can become conductive owing to radiation under high-voltage stress in the immediate vicinity of the conductor. Quite apart from losses, corona may be objectionable because of the luminosity in the dark. Essentially, a corona is a source of electromagnetic waves which causes audible noise and radio interference. Techniques for the calculation of radio interference are now well documented [14]. They can be classified into two groups: (a) analytical methods (b) empirical and semiempirical methods The latter are generally preferred, because they are related to performance of existing lines on which measurements were made [15]. The parameters affecting radio interference include: (a) gradient of the conductor or conductor bundle which is itself dependent on an equivalent radius defined in Reference 16. (b) diameter of the conductor. (c) number of conductors in a bundle. (d) distance from the line at which the interference is measured. (e) frequency (normally 1 MHz is accepted as a reference). (/) altitude of the line. (b) Mechanical parameters: Generally, a minimum size is specified and the UK practice is fairly typical, e.g. 16 mm2 for copper conductors and 25 mm2 for aluminium-based conductor. Larger conductors are selected on the basis of their mechanical characteristics. Reference 13 shows the variety of constructions which are possible. Possibly the largest conductor ever manufactured in the UK is that which has been used for the River Thames Crossing. It illustrates the versatility in design of conductors to suit specific requirements (Fig. 5).

corrosion. It is easy to appreciate that a wrongly selected grease can exude from the conductor, collect at the underside and become a source of corona. 3.3 Earth wires Earthwires differ from conductors only by their function, whatever can be stated for conductors is valid for the earthwires. They fulfil a dual function: (a) they intercept direct lightning strikes (b) they can provide a low-resistance path to the source of fault current. In addition, they have a beneficial effect, through coupling, on the insulation performance, as, in case of a fault, the earthwire or tower-top potential does not appear across the line insulation because a voltage of the same polarity is induced on the conductors. From the point of view of lightning, the material is of no importance; what matters is the surge impedance. Hence, many lines use steel as an earthwire material. For practical reasons, it is undesirable to use steel earthwire of smaller diameter than 6 mm, as at least one case of melting has been reported. Steel earthwires are still commonly used in many parts of the world, with diameters up to 13 mm for 500 kV lines. Earthwires are subjected to electromagnetic induction, thus their diameter must be checked to ensure a reasonable voltage gradient to reduce corona. With the development of networks, the shortcircuit levels have so increased that earthwires are quite often expected to carry short-circuit currents from 10 to 50 kA. For values above 10 kA, it is worth checking the temperature reached in the earthwire during fault conditions. It is now quite common in areas where high shortcircuit currents are encountered to include at least one conducting layer on the earthwire, to reduce its resistance. The rating of a conductor/earthwire under fault conditions is calculated on the assumption of an adiabatic phenomenon; i.e. no heat loss during the temperature rise. 3.4 Optical Fibres (in conductors or earthwires) Communication by overhead lines (data transmission, control, speech ...) has always been a great temptation to the overhead-line engineer. Thus, throughout the years, power-line carrier systems, conventional telephone or coaxial cables have been developed and used. Unfortunately, those solutions suffer from the dangers of electromagnetic induction. Optical fibres, on the other hand, being nonmagnetic, do not suffer from these drawbacks, and the last few years have seen the development of conductor systems incorporating optical fibres (Fig. 6).

:$=

Fig. 5

Thames Crossing conductor

1289 mm 2 aluminium area ACSR as used on the 400 kV River Thames Crossing. The conductor comprises 1/3.4 mm plus 90/3.1 mm diameter galvanised steel wires of 190 kg/mm 2 quality, surrounded by 74/3.4 mm diameter aluminium wires in two layers, and 36/17.16 mm 2 wedge-type aluminium segments, to give a finished overall diameter of 56.24 mm with a guaranteed minimum breaking load of 132 t.

As already indicated, nowadays aluminium-based conductors are the rule. The advantageous mechanical properties of aluminium alloys have been recognised for a long time, but AAAC has always been more expensive than ACSR, for equivalent conductivities. However, there are cases where such advantages take precedence over cost. One of these is the resistance to corrosion, since, being monometallic, the risks of bimetallic corrosion between the aluminium and the zinc on the steel core are nonexistent. Nowadays, it is customary to grease the inner layers of ACSR with a type of grease which must meet very stringent requirements as regards drop point and drainage properties, as a means of protection against this type of
1EE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

optical fibre in polymer tube (Q) polymer tension member polyethylene aluminium alloy steel
Fig. 6 Typical cross-section of conductors containing optical fibres
a 28.62 mm diameter and equivalent to Zebra ACSR b 14.88 mm diameter and equivalent to half-inch-steel earthwire

157

For the time being, most solutions consider the incorporation of the optical fibres in the earthwire of HV transmission lines which is at earth potential. Obviously, the design of the system must take into account the possibilities of faults, and adequate precautions must be taken. Such systems have already been tested and their performance has proved so satisfactory that they are being installed on actual lines [17]. This solution is not suitable for all lines, particularly at low voltages, as these are frequently strung without an earthwire, consequently consideration has been given to incorporating fibres into the phase conductors. The only objection to this is the difficulty in making the optical-fibre connection between earth and line potential. Already, suitable sealing ends for this purpose have been developed for up to 33 kV, and, hence, these special conductors can be used as phase conductors. At present, several trial installations on distribution lines are being planned or operated. 3.5 Theoretical calculations

are of the order of 5O-60C [22], but the temperatures of 75-80C are used to permit the transmission of more power through the same right of way. Hence, the selection of the operating temperature of a line is a matter of judgment based on the statutory regulations, if any, the anticipated power utilisation, the difficulty in obtaining wayleaves for a line and economic considerations. Within the UK, up to the introduction of the new statutory regulations [4] in 1970, all lines had a maximum operating temperature of 50C. Nowadays, with lines operating at 275 kV and above, there is a tendency to increase the operating temperature up to a maximum of 75C, but distribution lines are still usually operated at 50C. So far as the earthwires are concerned, the calculations are exactly the same, but an additional constraint may be imposed, that at a particular temperature the sag of the earthwire should be up to 20% less than that of the conductor. This arises in an attempt to reduce midspan flashovers due to lightning strikes (see Section 4.6.2).

3.5.1 Sag and tension calculations: For the uninitiated,


these calculations are nothing more than an anticipation of the conductor behaviour under varying conditions of loads and temperatures. Prior to the widespread use of computers, this was done by simulating the shape of the hanging conductor by that of a parabola [18]. Such a simplified approach is very satisfactory when the line is erected over flat or slightly undulating terrain, as in the UK. Nowadays, with the availability of computers, the shape of the conductor can be simulated by that of a catenary, which represents more accurately the hanging conductor [19]. The catenary equations have the added advantage of being easier to program. Experience has shown that, as the terrain becomes progressively more mountainous with greater sloping spans, a more rigorous treatment is necessary [20, 21]. There are generally, but not exclusively, two sagging bases. This expression defines a set of constraints which must be met: (a) The so-called EDS condition, when, in still air, at ambient temperature the every day stress (EDS) must not exceed a given percentage of the ultimate tensile stress (UTS). Typically, 20-25% are used for ACSR, and 15-20% for AAAC. This constraint results from a worldwide analysis of conductor failures due to vibration. (b) The maximum tension condition which occurs, generally, at minimum temperature and maximum anticipated external loadings. With the working load concept and factors of safety, the tension would be typically limited to 40-50% of the breaking load, but, with the new trends of ultimate loads, this value would be 70-80%. Generally, it is accepted that these bases are valid after all the expected creep has occurred. Depending on local conditions, on the conductor selected and on the length of the equivalent span, either of the above conditions can become the starting point for the calculations of the change of state in the conductor. It is then possible to calculate the maximum sag in still air at the maximum conductor operating temperature, which is the condition under which statutory clearances to ground must be maintained. This maximum temperature is normally accepted as 75C or 80C, as at this temperature, or below, the risk of annealing the aluminium wires is negligible. Occasionally, higher temperatures can be accepted, provided that the times are monitored. What is not often appreciated is that the economic conductor temperatures
158

3.5.2 Creep calculations: Creep is a phenomenon which affects most materials subjected to high stresses. It manifests itself by an increase in conductor length with time; hence, as an increase in conductor sag with the attached risks of infringing statutory clearances. In the early days, in the UK, creep of conductors was either ignored or was compensated for by increasing the sagging tension by a fixed percentage, irrespective of span length. This resulted in some spans being given insufficient clearances to ground, when the final sag could exceed the design figure, or alternatively too generous a margin was provided. Much research effort has been given to this subject, culminating in an IEE paper [23] and a CIGRE report [24] summarising the current state of the art. The main points stressed in these papers are: (a) Creep can be expressed mathematically as a change in temperature in the change of state calculations. (b) Creep depends very much on the method of manufacture of the aluminium wires. Conductors manufactured from rods produced by the hot rolling method may have up to 60% more creep than those made from rods produced by either extrusion or by the continuously cast and hot rolled process. (c) As a first step in making an estimation of the creep of conductors subjected to field conditions in which both the loads and temperatures vary with time, it is necessary to carry out creep tests on conductors under constant load and constant temperature conditions. From these tests, a predictor equation is formed which will give the creep strain at any time for any regime of constant stress and constant temperature. By applying a creep theory known as strain hardening to the predictor equation, it is possible to estimate the creep of a conductor under actual line conditions. (d) References 23 and 24 do give suitable predictor equations for standard strandings, but any new constructions will require new laboratory creep tests which must be carried out in a controlled manner to ensure the validity and repeatability of the results.

3.5.3 Current carrying capacity: The calculation of current rating for conductors is based on a heat balance equation: I2R + Ha = Hr + Hc
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

where I2R = the power dissipated in a conductor. Ha = heat absorbed by the conductor due to solar radiation. Hr = heat loss by radiation, function of temperature difference between conductor and surroundings. Hc = heat loss by convection which is a function of wind speed. A little thought will lead to the conclusion that all factors affecting the rating are variable in time and space. This has led to the two means of calculation: (i) Deterministic: This technique has been used for many years [25], in which typical values are accepted for maximum solar radiation, for maximum ambient temperature and for minimum wind speed; typically 0.5 or 0.6 m/s. Clearly, all the worst conditions do not occur all at the same time, hence current ratings calculated by this method would tend to underestimate the actual rating; however, this technique provides a convenient tool at the design stage. (ii) Statistical: In this technique each variable is represented by a statistical distribution, and the rating is calculated by accepting random values (e.g. Monte-Carlo method) for each variable in the heat balance equation. This technique requires an accurate knowledge of local meteorological data and is probably more suitable for use by the client. 3.6 Aerodynamic phenomena There are three types of aerodynamic phenomena affecting the performance and the fatigue endurance of overhead conductors and earthwires: (a) aeolian vibrations, affecting all types of lines. (b) subspan oscillations, affecting only bundle conductor lines. (c) full-span galloping, affecting all types of lines. With regards to wire failure due to fatigue, a useful empirical rule [26] is that the bending strain of a vibrating conductor should be limited to 150 microstrains, and techniques have been developed for the actual measurement of such strains [27]. A surprising feature of vibration damage is that wire breakages need not necessarily be initiated in the outer layer; and, in most cases, they are likely to occur in the penultimate layer, making their detection much more difficult. 3.6.1 Aeolian vibrations: The main characteristics of this type of vibration are its high frequency (5 Hz to about 30 Hz, although isolated cases of higher frequencies have been reported) and low amplitude (one or two conductor diameters) mainly in the vertical plane. It occurs most frequently in relative laminar winds in the range 0.5 to 10 m/s. Several years ago, the energy balance principle [28] was advanced as a means of forecasting the amplitude of vibration. It equates the energy input into a conductor with the energy dissipated by that conductor and associated fittings. The application of this principle is still made difficult by lack of accurate data on the wind energy input function, on self damping of conductors and energy dissipation by fittings such as vibration dampers. The latter aspect is under review and it is anticipated that dampers may be compared on the basis of their ability to dissipate energy. 3.6.2 Subspan oscillation: This type of vibration is restricted to bundled conductors, when pairs of subIEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

conductors lie in the same horizontal plane. In such configurations, the leeward conductor experiences variations in lift and drag forces, because it is placed in the turbulent wake of the windward conductor. The oscillation is a standing wave, predominantly in the horizontal plane, having a low frequency (0.5 Hz-3 Hz), high amplitude (up to clashing amplitude has been recorded), with subconductors frequently oscillating 180 out of phase with each other. This type of oscillation is initiated by low turbulent wind flow of 5-24 m/s and a wind direction normal to the conductors. A critical factor in control of subspan oscillation is the ratio of the distance between horizontal subconductors to the conductor diameter. Experience has shown that, if the ratio is greater than about 15, little difficulty should be encountered in controlling the oscillation to acceptable limits. Having fixed the ratio of distance between subconductors and conductor diameter, control of subspan oscillation is achieved by selection of the spacer, or spacer damper system, to be used. This means consideration of: (a) the positioning of the spacers throughout the spans (b) the flexibility in the vertical and horizontal planes of the spacer (c) the mass of the spacer and the damping built into its design (d) fatigue endurance limits of the spacer. Of these factors, the most critical appears to be the inspan location of the spacers. A large proportion of the control of subspan oscillations is achieved by ensuring unequal subspan lengths between spacers, and hence ensuring a spacer is not located at a node. Considerable work has been done on the optimisation of the position of spacers and spacer dampers within a span [29]; and each spacer manufacturer has developed his- own methods of determining the positions for his products, taking into account practical erection tolerances. 3.6.3 Full-span galloping: This form of vibration can occur with both single and bundled conductors and is a low-frequency (0.1 to 1 Hz) high-amplitude ( several metres) motion, predominantly in the vertical plane having 1, 2 or 3 half wavelengths per span. The motion is initiated by two different mechanisms: (a) Non-ice-initiated galloping: This form of galloping is restricted to conductors of large diameter (> 40 mm) which have more than 30 round wires in the outer layer when the wind direction is coincidental with the direction of lay. The aerodynamic force initiating the galloping conditions arises through a difference in the forces between the wind travelling under and over the conductor, as in one case the wind travels with the direction of lay and consequently there is little friction effect, while in the other case the conductor presents a rough profile, with a consequential high frictional effect. This condition was well documented by the experience of the CEGB in operating the Severn Crossing [30]. The solution to this form of galloping is to ensure that all conductors have a smooth profile by using specially shaped wires in the outer layer (see Fig. 5). (b) Ice-initiated galloping: This form of galloping can occur on most, if not all, conductor constructions, and is caused by a slight build up of ice altering the aerodynamic shape of the conductor, so that an aerodynamic lift occurs. A necessary condition required to produce this galloping is that the temperature of the line should be within a few degrees of freezing, although, as the ice formation required
159

to initiate galloping is so small, it is frequently impossible to detect its presence on the conductor. A number of different types of dampers based on mechanical or aerodynamic arrangements [31, 32] are suggested, but the adoption of antigalloping dampers is considerably hampered by the lack of experimental results. At present, most research work is concentrated on lines of 220 kV and above. Even if suitable antigalloping devices were developed, it is not expected that such devices would be fitted to every transmission line, as the cost of such protection is estimated to be high, up to about 5% of the cost of the line. Consequently, some method of easily identifying the critical section is required [33]. Although the problem has been noted many years ago, overhead-line engineers have no definite answer, and, should a safe design be needed, recourse may be made to Lissajous' figures advocated in 1939, with the full knowledge that the solutions are likely to be over conservative [34]. 4 Insulators and insulation levels

clean unpolluted conditions, they have been used on lines at 66 kV. Their choice is dependent on the type of structure on which they are mounted and on the overall economics. As they effectively raise the conductor above the crossarm, they can occasionally permit the installation of shorter supports. 4.2.2 Cap-and-pin insulators: In this case (see Fig. 8), while the same formulation of glazed porcelain is used, the

Fig. 8
a Normal;

Cap and pin insulators


b antifog

4.1 Introduction The cheapest form of electrical insulation known to man is air, whose performance under electrical stress is fairly well documented. To maintain the conductors at adequate spacings, or adequate height above ground, supports are installed, to which the conductors must be attached. Manmade materials are therefore necessary to prevent flashovers to the supports under normal conditions (power-frequency voltage) or abnormal conditions (overvoltages due to internal causes such as switching, or external causes such as lightning). 4.2 Insulators for overhead lines Until recently, glazed porcelain and glass were the only materials used for the manufacture of insulators, but a new family termed composite insulators is now gaining acceptance and is increasing its share of the market. By definition, a composite insulator consists of at least two insulating materials: a core of oriented glass fibres in a resin matrix to provide mechanical strength, and an external housing to protect the core and to achieve suitable electrical characteristics. 4.2.1 Rigid insulators: These insulators (see Fig. 7) comprise an insulating body mounted on a pin, or spindle, for attachment to the structure. Glazed porcelain and annealed glass are used for their manufacture. They are commonly restricted to lines up to 33 kV, although, under

alternative in glass is treated to reach a toughened state which puts the external layer in tension and the inner material in compression. In both cases, the design of the metal fittings is such that, effectively, the insulating material is in compression. In addition, the metal parts permit a reasonable amount of flexibility if several units are installed in series, especially because of the ball ending of the pin, and the shape of the recess in the cap. Up to 11 kV, single cap and pin insulator units can be used. For higher voltages, series parallel connections, with suitable yoke plates, can provide an extensive range of mechanical and electrical properties (Fig. 9). This is the type of insulator most commonly used in most parts of the world. Their mechanical characteristics have now been standardised [37, 38]. 4.2.3 Long-rod insulators: This type of insulator (see Fig. 10) is manufactured from glazed porcelain and is extensively used in Central European countries (Germany, Austria, Switzerland etc.). It has not gained acceptance in the UK because British engineers dislike the thought of porcelain under tension. Besides, they are produced in a limited range of lengths, which can adversely affect tower designs in that they cannot readily be tailored to suit the electrical requirements as can an assembly of cap-and-pin insulators. There is also the difficult and costly problem of spares associated with long-rod insulators compared with cap-and-pin insulators which are standardised internationally and available worldwide. However, long-rod insulators have given good service and are still being used extensively by some countries. To achieve the characteristics required by an overhead line, they can be used also in series-parallel arrangements, but extra care is needed when designing the fittings. With the ever increasing population of insulator years, accidents are bound to happen; and the breakage of one string in a parallel arrangement can lead to a catastrophic failure of the adjoining string, unless elaborate precautions are taken in the design of all accessories. 4.2.4 Composite insulators: Through intensive testing and experimentation, composite insulators have been developed to a stage where they are considered seriously
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

Fig. 7 160

Rigid insulator

Work is now proceeding to evolve test procedures to eliminate the designs which may be prone to this type of failure. From an installation point of view, their light weight and high strength are a considerable benefit. (How many engineers realise that on a single-circuit angle tower at 765 kV the mass of the conventional cap and pin insulators would exceed 10 tonnes? Equivalent composite insulators could be handled by one man per phase.) As the mechanical strength is provided by parallel glassfibre filaments, any desired strength can be achieved. A problem which is slowly being solved is the fitting of end metal connections to ensure no unacceptable increase of localised stresses. The electrical performance is ensured by an external insulating housing which can be shaped to provide a much longer creepage path than is possible with conventional insulators. In the UK, composite insulators were tried in the early 1960s at 132 kV on wood poles [1], on an experimental line, at Connah's Quay. Their installation on operational lines has given mixed results. Modifications to the design of this type of insulator have led to their limited use being resumed, and satisfactory performance is now observed (Fig. 11).

Fig. 9 Suspension insulator assemblies on 400 kV Thames Crossing at West Thurrock


View from top crossarm

for installation on new lines or used for replacement of insulators on old lines. Their rate of application was slowed down following some mishaps on existing lines, and, in most cases, the trouble was identified as being due to brittle fracture.

3 3 3 3 3 3 3 3 3 3 3 3
3
Fig. 11 132 kV steel-bole construction with insulating crossarms in the UK (Eastern Electricity Board)

Although, in theory, the length of the insulator can be designed to suit exact requirements, economic pressures will lead to standardisation in strength and in length; but, in all cases, there will be fewer metal parts along the string and, consequently, the voltage distribution along the length of a string will be more linear than along a string obtained by connecting several cap-and-pin units in series.
4.3 Insulator characteristics

Fig. 10

Long-rod insulator Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

It is obvious that insulators must fulfil an electromechanical duty. It was necessary to evolve series of tests which
161

IEE PROCEEDINGS,

could give the user an assurance that the insulators are likely to perform satisfactorily: (a) The mechanical stresses to which an insulator is subjected are either bending, for pin insulators, or tension, for all other types. The tension is due to the weight of the conductors at suspension towers or to the mechanical tension of the conductors at tension towers. Seldom are insulators subjected to compression, except in certain designs of V-insulator assemblies. (b) The electrical stresses are due to the power-frequency voltage, to the externally induced overvoltages (lightning) or to internally induced overvoltages (switching). Thus, the testing procedures must reflect the service conditions. Throughout the years, on the basis of increasing experience, specifications have been improved, and, while some aspects are still being investigated, it is now common practice to accept insulators which comply with either national standards or international recommendations. Thus, BS 137 Parts 1 and 2 are available in the UK. IEC Standard 383 [37] prepared by the International Electrotechnical Commission is used internationally. The differences between BS 137 and IEC 383 are minimal. There is a ruling of the European Economic Community according to which any IEC document accepted by a National Committee must be used as a national standard, but slight modifications can still be incorporated in the latter. Insulators are divided into two types according to their construction: Type A: for which the length of the shortest puncture path through solid insulating material is at least equal to half the length of the shortest fiashover path through air outside the insulator. Type B: for which the length of the shortest puncture path is less than half the length of the shortest fiashover path through air outside the insulator. An insulator unit or a rigid insulator is characterised by the following values: (a) the specified dry lightning-impulse withstand voltage (b) the specified wet power-frequency withstand voltage (c) the specified electromechanical failing load (for type-B string insulator units only) (d) the specified mechanical failing load (for type-A string insulator units and those of type B, for which the electromechanical failing load is not applicable and for rigid insulators) (e) the specified puncture voltage (for type-B insulators only) (/) the specified significant dimensions, including the creepage distance. Special tests or reception tests have been devised to ensure that each individual unit satisfies the above requirements. From the point of view of the system, it is the performance of an insulator string or an insulator set which is of importance. In fact, nowadays, final tests are always performed on insulator sets; i.e. insulator strings complete with all their fittings. Type tests are carried out to check that the following quantitative characteristics are satisfied [37]: (a) the specified dry lightning-impulse-withstand voltage (b) the specified wet switching-impulse-withstand voltage: (Note that this test is only applicable for system voltages equal to, or greater than, 300 kV; and that the geometry of the test arrangement must represent as closely as possible the conditions of use) (c)the specified wet power-frequency-withstand voltage (d) when applicable, corona and radio interference tests.
162

4.4 Insulator selection As already stated, insulators must perform satisfactorily under service conditions. With a knowledge of the external loads applied to them, it is relatively easy to choose insulators developing clearly defined mechanical strengths. In addition, they must satisfy some electrical criteria which can be found in BS 137 [35], IEC-71 [38] and BS 5622 [39]. Most of the requirements have already been covered in the preceding Sections, but no mention has yet been made of pollution and of the influence of the creepage path, which is in fact the length of insulation. The longer the length of the creepage path, the better will be the ability of an insulator to withstand electrical stresses. Under pollution conditions, it has also been found that the length of the creepage path is the main factor controlling the insulator behaviour, although the insulator shape also plays an important role. Two methods of testing have been evolved for a comparative assessment of insulator performance under pollution: (a) the salt fog method [35, 40] (b) the solid layer method [40]. To determine either the equivalent salinity or the surface conductivity of the pollutant, a test insulator can be installed at a polluted site and the maximum leakage current can be measured over a long period. Attempts can then be made to reproduce these conditions in a laboratory to decide on the best insulator to be used. However, a considerable amount of work has already been done on the problem of pollution and guidelines are available (see Table 6) [41].
Table 6: Creepage length as function of pollution level
Pollution level Minimum specific nominal creepage length (between phase and earth) mm/kV (phase to phase referred to highest system voltage) 16 20 25 31

1 Light II Medium III Heavy IV Very heavy

The minimum creepage distance of an insulator (or insulator string) can then be calculated. Minimum creepage distance = minimum specific nominal creepage length (Table 6) x highest system voltage (phase to phase). Discussions are taking place at an international level on the introduction of a correction factor which would depend on the shape of the insulator. For cap-and-pin units this factor is likely to be less than 1.1. In the meantime, the above relationship can be used.

4.5 Insulation co-ordination at a tower


Having selected insulators to suit particular criteria of voltage, overvoltage, pollution, loads, altitude etc., an attempt must be made to ensure that air-gap clearances between live metal and support are adequate. While the solutions may be complex in detail, a guiding principle [42] could be stated as follows: the minimum distance between live metal and earthed steelwork should have a fiashover value approximately 10% greater than the flashover value of the insulator. It is the application of this principle which gives rise to problems because of the variability in the conditions, and statistical techniques would be
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

the best to calculate probabilities of flashovers. As statistical techniques are not yet available, for design purposes, deterministic values are preferred, bearing in mind that it would be economically unjustifiable to assume all the worst conditions at the same time (Fig. 12):

shapes should not be underrated. Fig. 13 illustrates the 50% fiashover voltage of a rod/plane gap for different wave shapes.
5000

4000

&3000 o

2000

1000

0 Fig. 12 Basis of clearance diagrams to define tower top geometry 9 = angle of insulator swing due to wind Cd = deflected clearance Cs = still-air clearance lg = length of gap between horns /, = insulation length

2 4 6 rod/plane air gap length, m

Fig. 13 Effect of wave shape and polarity on insulation strength of rod/ plane air gap
1 Negative 200/3000 /JS (switching surge) 2 Negative 1.2/50 (is (impulse) 6 Positive 1.2/50 fis (impulse) 9 50 Hz (power frequency) 10 Positive 200/3000 /is (switching surge)

(a) Under still air conditions, it is usual to select the distance Cs such that the electrical strength of the air gap is 10% greater than the impulse strength of the insulator set, taking into account the shapes of the electrodes. (b) Under wind conditions, the distance Cd should have an electrical strength 10% greater than that of the insulator under wet power-frequency voltage. While criterion (a) is relatively easy to apply, the difficulty with criterion (b) is to define the value of the insulator swing 0. Under the old principle of working loads multiplied by factors of safety, it was usual to calculate this angle on the assumption of the working pressure for an arbitrarily defined weight-span/wind-span ratio. Usually the pressure would be 40% of the ultimate (factor of safety = 2.5) and the weight-span/wind-span ratio would be 0.7. With the new trends of ultimate load designs, proper statistical analysis would be needed, when, in any case, a value should be given to the tower designer to help him in defining the tower geometry. A logical approach, similar to that which has been used so far, would be to use wind speeds, hence pressures, corresponding to shorter return periods reflecting more common happenings. Generalised techniques for the definition of clearances as a function of swing angles are not yet available, but this item is on the agenda of several international working groups. The influence of the wave shape and of the electrode
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

In fact, the rod/plane configuration is normally accepted as a reference. However, at a tower, many configurations of electrodes can exist, e.g. rod/rod, rod/plane, conductor/ plane etc. So far as lightning surges are concerned, the shape of the electrodes is less critical than for switching surges. On the basis of extensive testing [43], it was concluded that the electrical strength under switching surges of an air gap could be expressed by V = K x 500d06 where V = 50% flashover in kV d = gap in meters K = factor reflecting the influence of air gap (see Table 7 for some values)
Table 7: Influence of electrodes on F/O values for switching surges
Gap geometry Rod/plane Conductor/plane Conductor/window Conductor/structure (under) Conductor/structure (over & laterally) K 1.0 1.15 1.2 1.3 1.35

163

4.6 Lightning 4.6.1 Main causes of lightning outages: There are four ways in which lightning can affect overhead lines: (i) Discharge to nearby ground: In this case, the capacitance between the leader stroke and the earth is discharged quickly and a travelling voltage wave is set up in the line conductors. The impulse insulation level of lines above 66 kV is normally sufficient to withstand this effect. (ii) Discharge to a tower: When lightning strikes a tower, the voltage V impressed at the top of the tower is given by the vectorial sum of two components:

above 5 kA, flashover across insulators occurs almost in all instances of direct strikes. The' measure of lightning incidence is expressed as the Keraunic level (KL). This is the mean annual number of days that thunder is heard in an area. There is a direct relationship between the KL and the number of strikes to the ground or a line, but this relationship can vary from country to country or from area to area. 4.6.2 Lightning protection: The best method of protecting transmission lines from lightning-induced surges is to provide effective shielding from direct strikes, and to design the line with the highest impulse insulation withstand levels consistent with acceptable, practicable and economic limitations. Effective shielding can be provided by installing one or two overhead earthwires, which should be positioned to give, in most cases, a shielding angle not greater than 30 over the outer phase conductors. The shielding angle is defined as the plane angle between the vertical and a line between the earthwire and the phase conductor. While the shielding angle is not an exclusive design parameter, its effectiveness is illustrated by Fig. 15.
0.8r

where R = tower footing resistance / s = lightning strike current defined statistically (see Fig. 14) [44] L = tower inductance

98.00 95.00 90.00 80.00 60.00


(

40.00 20.00 10.00 5.00 1.00 20 30 40 angle of shade, degrees A 6 8 10 20 I kA 40 100 200 300 Fig. 15 50

0.1

Influence of angle of shade on lightning performance

Fig. 14 144]

Distribution of lightning strike currents as function of height

If the value of V is greater than the impulse level of the line, flashover will occur between the tower and the phase conductors. This is a very simplified statement and in a critical study the effect of electrostatic and electromagnetic couplings of the conductors, the effect of multiple reflections etc. should be taken into account. (iii) Discharge to earthwire at midspan: In this case, the voltage impressed at the point of strike is given by V = (IJ2) x Zs, where Z s is the surge impedance of the earthwire. Two lightning currents, each of value IJ2 travel in opposite directions from the point of strike, until each meets a tower and is discharged to the ground through the tower footing resistances. The line insulation is subjected to the various voltages arising from this sequence of events. (iv) Discharge to the conductor: This can happen where no earthwires are provided or due to shielding failure. As the surge impedance of a line conductor is between 250-350 Q and the majority of lightning currents are
164

a 220 kV lines, single-circuit steel towers b 110 kV lines, single-circuit wood poles c 110 kV lines, single-circuit steel towers d 110 kV lines, double-circuit steel towers From CIGRE paper 314, 1960 NB. Interruptions due to shielding failures

The tolerable effective footing resistance of transmission-line towers is related to the impulse level of the line insulation, and is often calculated on the basis of an acceptable rate of lightning outages. As a guide, typical values often quoted are approximately 10 fi for 132 kV lines up to approximately 25 Q for 400 kV lines, with variations in practice of up to 100%; especially when very high-resistivity soils are encountered. Economic considerations come also into play. To reduce the footing resistance, special steps have to be taken, such as earthing rods, counterpoises etc. In areas of high earth resitivity, e.g. rock, it may even be necessary to run a continuous underground earth counterpoise connecting the towers in a section of the line. On double circuit lines, with better earthwire shielding, there is an increased likelihood that back flashovers could affect both circuits. To prevent back flashovers affecting both circuits, attempts have been made to introduce differ1EE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

ential insulation on each circuit, on the premise that the higher insulation level provided on one circuit would reduce the probability of outage on that circuit. In view of the random nature of lightning, no guarantee can be given at the design stage, although satisfactory behaviour has been reported; the main problem being a decision on the value of relative insulation levels. To protect against midspan flashovers, it is also customary to reduce the sag of the earthwire in comparison with that of the conductor so that the spacing at midspan is more than at the tower (see Section 3.5.1). The impulse insulation level at the tower, can be increased not only by increasing the normal insulation but also by using wood crossarms. In this respect, it is worth referring to the 132 kV and 275 kV steel-tower transmission lines in Malaysia, where the crossarms for suspension supports are generally manufactured from local timber [45]. These lines have given very satisfactory lightning performance. For the design of such lines, an impulse withstand level for wood of approximately 160 kV/m is normally assumed. So far as distribution lines are concerned, they are affected not only by direct strikes, but also by indirect strikes (discharges to nearby ground) when their operating voltage is less than 33 kV. As a result of investigations in the UK, some insulation levels are recommended to ensure acceptable performance of distribution lines up to 132 kV, and these can be found in Reference 46.

lightning flashes from cloud to ground, which is also assumed to be proportional to the Keraunic level. Owing to lack of data, this proportionality is assumed to be the same throughout the world. A similar assumption applies regarding the waveform of lightning flashes and their frequencies. The validity of these assumptions could be questioned, but until new methods of recording lightning data are introduced in meteorological stations, the prediction techniques can only be considered as the best available for the time being. It must also be appreciated that lightning is a random phenomenon which can only be treated statistically; i.e. no discrete value can ever be quoted, only a mean value with a dispersion around this mean value. For quite a long time, there was no means of calculating the number of flashovers owing to shielding failures. Most theories assumed perfect shielding. However, when a line is installed on tall towers, there is, under the conductors and on the sides, a zone which is not really protected by the earthwire. Reference 47 provides a tool for so doing, and Table 8 illustrates the result of computer calculations using the latest techniques. The proportion of shielding failures is not negligible.
Table 8: Calculated outages per year for 132 kV line, in mountainous terrain. Keraunic level = 40
Case 1 Line conditions (footing resistance) Shielding failures 0.38 Back flashovers a 0.13] >.26 b 0.13j a 0.26] >.51 b 0.25] a 0.25] .47 b 0.22J a 0.25] b 0.22J 0.17
.47

Total

fl = 10fi
angle of shade = 10 R = 20fi angle of shade = 10

0.64

4.6.3 Predictions of lightning performance: The lightning performance of an overhead transmission line is a field in which much research work has been and is still being undertaken. Over the years, many theories have been put forward for predicting the lightning performance [47, 48, 49], generally based on model lines to which surges have been applied to assess whether flashover occurred. The testing was made with scale models of surge generators, analogue computers and mathematical models applying Monte-Carlo techniques [50]. As a result of such work, each author produced a series of curves which give the predicted flashover for the model considered. By reference to these curves, it is possible to make a prediction for a new line by interpreting, extrapolating and correcting the figures for the various line parameters. Attempts at validating these results by monitoring actual line performance have been made, and a reasonable correlation has been found. Each of these methods is internationally recognised as a valid method of predicting the lightning performance of an overhead transmission line. However, no one method has shown better results over the complete range of transmission lines. Indeed, no one method covers the complete range, thus the technique used for a particular line depends on the availability of a suitable model and the success in predicting the performance of a line at the line voltages considered. Hence, it is necessary to find the most suitable model and use it for a specific purpose. The prediction of the lightning performance of a line is not an exact science and to quote from Section 12.2 of Reference 48 'One should not be too surprised if a line, when built, has a tripout rate that differs by two or three times from a calculated rate. This holds true, particularly if the tripout rates calculated or observed are very low'. One of the main reasons for the wide disagreement between theory and practice is probably in the base data. The lightning flashover rate is proportional to the Keraunic level. The important aspect is, in fact, the number of
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

0.38

0.89

Mixed resistance* angle of shade = 10

0.38

0.85

Mixed resistances* angle of shade = - 4

0.22

0.69

Mixed resistances* angle of shade = -17

a 0.25") .47 b 0.22J

0.64

a represents calculated failures b represents correction for tower height *R = 10 fi (40%), R = 20 fi (40%), R = 60 fi (20%).

Supports

5.1 Introduction Support design is influenced by a considerable number of factors which vary from project to project, and, therefore, the overall design concepts are continually being affected. As far as the selection of the appropriate wind speed for design is concerned this has already been discussed in Section 2.4.4. Consideration has also to be given to ice loading if this is appropriate for the territory in which the line is to be built (see Sections 2.4.6-2.4.8). Basic span, wind and weight spans are selected, based on local factors and overall economics of the line under consideration. Having established these basic parameters, it is then necessary to consider the loading conditions for which the supports should be designed. 5.2 Loading cases The design loading cases can be treated under the following headings: (i) normal conditions (ii) torsion conditions
165

(iii) construction and maintenance conditions (iv) containment or anticascade conditions.


5.2.1 Normal conditions:

T = maximum transverse loads on conductors, insulators and tower body with wind at right angles to the line (consideration should also be given to an inclined wind blowing on the line) V = maximum vertical loads due to weight of conductors and insulators, and, as a separate condition, the minimum vertical loads which will be encountered due to the terrain of the line.
5.2.2 Torsion conditions:

T = as for normal conditions, but with a reduced value at the point where the torsion load is applied V = as for normal conditions, and again with a reduced value where the torsion load is applied L = longitudinal load related to the maximum calculated phase tension (with allowance for insulator swing along the line for suspension towers). There is still considerable argument as to the definition of the torsion condition. So far as single-circuit towers are concerned, it is generally agreed to consider the loads due to one severed phase (or earthwire). However, for doublecircuit towers there is, as yet, no general agreement as to whether one or two severed phases should be considered. This is further complicated with bundle-conductor phases, as to the number of conductors to be severed within the bundle.
5.2.3 Construction and maintenance conditions:

T zero, or, in some cases, loads based on a wind speed having a return period of 2 or 3 years V = depends on construction and maintenance, but should allow for lifting of conductor and insulators plus the weight of linesmen (it may well reach a value of between 2.5 and 3.0 times that for the normal condition) L = a percentage of the everyday tension of the phase conductors (see Section 3.5.1). These loads would only be applied to individual attachment points in isolation. 5.2.4 Containment or anticascade conditions: In most cases, it will be found that a reasonable torsion condition will provide adequate strength against cascading. However, it may be decided to check the tower for the following: T = zero, or loads based on a light wind having a return period of say 2 years V = as for normal condition L = a percentage of the everyday tension of the phase conductors. These loads would be applied to all attachment points simultaneously. 5.3 Major factors affecting support designs In addition to the factors already mentioned, it is necessary to consider the minimum electrical clearances to be maintained between live elements and earth (see Section 4.5), the earthwire shade angle (see Section 4.6.2). The effect of these two items is that it dictates the physical shape of the top of the supports from the earthwire peak down to the lowest conductor level, particularly when considered in
166

conjunction with minimum phase-to-phase and circuit-tocircuit centres, if appropriate. The overall height of the support, for level ground situations, is governed by the required clearance from live metal to ground. Having established all of the above parameters, then the final shape of the support has to be determined around these to try to establish the most economical design. However, there are many other influences which can greatly affect the designer's thinking, one of which is the availability of steel sections in the country where it is decided to fabricate the supports. Very often it is not possible, at the early stages of a tender, to decide where the supports will be fabricated, and it may be necessary to produce two or even three designs incorporating different sections to suit the various fabricators under consideration. This, of course, can be a nightmare for the designer, bearing in mind that he may be faced with several section lists to work from, with wildly varying sections. Most conventional towers are constructed from angle sections and are bolted together directly or through plates. Over the last few years, most countries have rationalised the rolling of steel angle section, so that today only about 60% of the sections may be available compared with 10 years ago. In addition to this rationalisation, the designer is faced with minimum rolling tonnages for particular section sizes. This minimum tonnage may vary from 50 to 250 tonnes and, obviously, can create problems with the lighter sections in ensuring that the overall contract requirements are such as to meet these minimum tonnages. If this is not possible then the section in question cannot be considered and the designer is forced into using a heavier section. A secondary factor related to sections is that of available steel qualities. In most countries there are two grades of steel available, a mild steel and a high-yield steel. However, in certain countries, only mild steel is available at the present time, whereas elsewhere an extra-high-yield steel can be obtained. High-yield and extra-high-yield steels are generally more expensive than mild steel, and this adds another factor into the equation of trying to produce the most economical design. Aspects related to construction of the line cannot be ignored, as there may be restrictions on member weights or lengths to satisfy the erection needs. For instance, if access to the line is difficult or impossible by conventional wheeled transport^ it may be necessary to use either helicopters or manpower, depending on the economics. Obviously, if member weights or lengths have to be restricted then this will introduce extra joints into the structure, which, in turn, increases the structure weight and hence cost. The overall objective is to attempt to produce a tower plus foundation at minimum cost. This obviously influences the selection of the base width of the tower. A small base width gives high foundation loads and hence large foundations, whereas a large base width reduces foundation sizes. It is necessary to obtain the costs for fabrication and erection of steelwork, and also the costs of excavation, concrete and reinforcement, to be able to select the most economic solution. It is rarely possible to obtain accurate prices at an early enough stage during a tender to make this possible, and so rates based on experience in particular territories are usually adopted for this exercise.
5.4 Supports

Having considered the major factors affecting support design, it is now appropriate to consider the various types of support that can be contemplated.
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

5.4.1 Towers: By the time the tower designer starts work, the decision will have been made on the type of construction: double circuit or single circuit, vertical or horizontal formation of phases etc. There will be instances where even triple or quadruple circuits may be required, particularly in countries where land is at a premium. Some of the more common types of towers are shown in Fig. 16.

5.4.3 Poles: Another category of structure is loosely defined as a pole. A pole is usually considered to be a

Fig. 17

Typical guyed structures


c Y-guyed tower d Guyed-single-member tower

a V-guyed tower b Guyed portal tower

Fig. 16

Typical types of towers


b Single-circuit towers: (i) triangular formation (ii) delta formation (iii) horizontal formation

a Double-circuit towers: (i) 110-150 kV (ii) 220-275 kV (iii) 400-500 kV

Generally speaking, most double-circuit lines are in vertical or semivertical (when ice is expected) formation, using a conventional double-circuit-type tower. Single-circuit lines can be in either vertical, triangular or horizontal formation. In the latter case, a horizontal-type tower is used, commonly referred to as a waisted tower. As a general rule, it is considered satisfactory to design the vertical formation towers using a statically determinate analysis; however, for the waisted-type tower, this type of analysis is far from satisfactory, and they need to be analysed considering the elastic properties of the members to obtain the correct distribution of loads in the structure. 5.4.2 Guyed structures: Where land is not at a premium and there are no problems with wayleaves, it is possible to consider the use of guyed structures for suspension towers. This type of structure would normally be considerably cheaper than the more conventional self-supporting tower, but suffers from requiring considerably more land to build it on, because of the spread of the guy ropes. Guyed structures can take several forms, the most common of which are guyed portals, guyed V-structures and guyed Y-structures. These structures are generally used where the route of the line is reasonably fiat. They are ideal for erection with a mobile crane, as they can be assembled completely on the ground and lifted into their upright position in one piece (see Fig. 17).
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

structure which consists of a single main member, but can also be a straight sided lattice type construction. Poles can be fabricated from steel, concrete, wood or other more sophisticated type materials, such as glass-reinforced plastic or laminated wood, and are normally erected on a monoblock foundation. The type of material to be used would depend on availability, transport and cost. Poles are now accepted for lines up to 220 kV and may be used on higher voltage lines where aesthetics are of prime importance, particularly in urban areas. The poles used for these higher voltages are much more sophisticated type structures and are generally fabricated from a foldedplate-type construction with flat sides in plan where the sides form a regular polygon. This type of construction tends to be very expensive and would normally only be used where environmental considerations prevail. Apart from environmental considerations, poles are normally considered where land is at a premium, as they are rarely more than one metre wide at ground level and generally considerably less. The most common type of pole these days seems to be manufactured from tubular steel, although there are instances where concrete poles of either solid or annular prestressed form may prove to be economical. Although poles can be used for suspension positions or for small-angle deviations as self-supporting structures, it is normal for large-angle deviations to guy the poles or to use the more conventional lattice construction tower. Guying of poles can become quite complicated at times, to take account of the various design conditions and the necessary electrical clearances. Pole structures can also take the form of an H-type structure, particularly for terminal positions, and once again these would normally be guyed. The basic one-piece-type pole is ideal where there is no problem with transport, access or erection, but where these are difficult it is preferable to think in terms of an all lattice construction pole, which can be transported 'piece small' and then assembled on site. It can be very cost effective even when compared with a tubular pole, as the fabrications costs for the lattice-type construction tend to be cheaper than those for a tubular pole.
167

5.5 Structure design

Whereas a few years ago very few designers used computers for design of transmission towers, today most of them have computer programs of some sort to assist with analysis and/or design. There are obviously many advantages to the design engineer in having computers to assist him in his work. Apart from the obvious advantage of being able to consider numerous design conditions which would have taken weeks to do manually, it also enables various structure shapes to be considered, to try to establish which is likely to be the most economic. This latter advantage is considerable when compared with the manual design, where the design engineer had to use his experience to produce the tower outlines, as time would not allow several alternative shapes to be considered. Another advantage is that it allows the designer to consider different section lists, which then allows a comparison of costs between different fabrications to see which solution is likely to prove the most economic. Unfortunately, no computer is so sophisticated that it can accurately assess the design of every member on the tower. As nearly all towers are constructed from angle sections, it means that almost every connection is eccentrically loaded, and the design engineer has to assess the strength of an individual member taking account of its axial load and bending moments created by eccentric connections. Most designers in the overhead-line business employ their own 'strut curves' which have been based on their own experience over the years, mainly gained from fullscale testing of towers. These strut curves would take account of the type of member (i.e. main leg or bracing) and will allow for normal end eccentricities and end fixity. The end fixity will take account of the number of bolts in the connection and also the type of connection. The type of connection is Very important, as this can have a considerable effect on the strength of the member; particularly where it is necessary to use plates to achieve the connection, in which case the strength of the member can be considerably reduced. The design of so-called unstressed (redundant) members (these are bars which are put into the structure to support the stressed members) has also become far more sophisticated today, compared with a few years ago when they were purely designed on a slenderness ratio basis. Nowadays, it is possible with the use of computers to consider these members in a nonlinear analysis, which shows that they can in fact carry quite high loads in performing their duty of supporting stressed members. This aspect is particularly important on the heavier towers on lines with voltages of 400 kV and above. The design of towers is made even more difficult when the designer has to contend with such things as rolling and fabrication tolerances and slippage of bolts. These factors can have a considerable effect on the analysis of a structure, but, even with the aid of a computer, it is not possible to consider all the possible combinations of fabrication tolerances and bolt slip, and this aspect is still left largely to the designer's experience. Where towers have to be designed to withstand wind loading in typhoon areas this has also created problems for the designer in that the largest angle generally available today is 250 mm x 250 mm, and for the main legs on tension towers this may not be large enough. There are then several alternative options open to the designer. One is to use a compound angle leg made up from two, three or four angles bolted together. This has a major disadvantage
168

in that it increases the effective wind area of the leg by a factor of at least two and complicates the detailing of tower members and connections. It also creates problems with fabrication, as the members forming the compound leg have to be matched in the factory to ensure a satisfactory fit on site. The second option is the use of tubes or solid round members which gives certain advantages but also creates disadvantages. The main advantage is that it gives an effective wind drag coefficient less than that for an angle section of the same size owing to its circular shape. It also allows concentric connections by utilising welded plates to connect to the bracings. The major disadvantages are that the basic material is more expensive and the workmanship required in the factory is very high because of the need for welded plates. Finally, there are two very important matters which the tower designer has to consider. The first of these is for him to consider whether the abnormal loading conditions, i.e. so-called broken wire or unbalanced conditions will give the towers sufficient longitudinal strength to avoid the possibility of a cascade failure in the field. If no broken wire conditions are specified, then, obviously, the tower will have virtually no longitudinal strength and a cascade would be a real possibility. If a complete phase is considered to be broken, then this probably gives the absolute minimum required longitudinal strength to a tower, but, if only a part phase is considered, then there may be a danger of a cascade. Where more than one phase is considered to be broken simultaneously, then this obviously gives increased longitudinal strength to the tower and significantly reduces the chances of any problems. The second matter relates to construction and maintenance loadings and is largely a question of ensuring that sufficient vertical loading capacity is given to the crossarm design. When towers were being designed for the home market they were designed for ice-loading conditions, which invariably meant that they had sufficient vertical loading capacity to cope with maintenance loading. However, when towers are being designed for areas in the world which are not subjected to ice loading, the vertical loading capacity is not normally sufficient to cope with construction or maintenance loadings, and, therefore, the designer has to ensure that special loading conditions, to cover these conditions, are considered. If he does not do this then it may be very difficult to string the conductors without very serious problems being encountered.
5.6 Detailing

The final strength of a tower can be greatly affected by the details of the connections and, therefore, it is extremely important that the draughtsman producing the member details works very.closely with the designer. The details for a tower can be produced in two ways, either on the template floor, where the structure is laid down full size, or alternatively on the drawing board, where all connections are drawn out to a large scale and physical lengths of members are calculated either by hand or by computer. In recent years, there has been a move away from the template floor approach, as not many fabricators have this facility. As a result, the task of detailing has fallen more and more to the drawing office. This has encouraged most drawing offices to look to the computer to aid them, at first in a small way to determine dimensions between node points of the structure and, for quite a few years now, towards complete detailing by computer. Although there are many structural detailing programs
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

on the market, primarily developed for buildings in either steel or concrete, these are of no use for tower detailing which requires a much more sophisticated approach. The reason for this is that most members in a tower are not in either vertical or horizontal planes, unlike buildings, and, therefore, generally a 3-dimensional approach is necessary. Although it is possible to consider large parts of the tower in two dimensions where they lie in a fiat plane, it is still necessary to check whether members on two adjacent faces may foul each other at a connection. Certain parts of the tower and, in particular, crossarms and hip bracings have to be treated completely 3-dimensionally. For some connections, it is necessary to clip bars or completely remove one flange of an angle, and, in these instances, it obviously reduces the strength of the bar and needs to be checked by the designer. Also, because of the 'eiffelised' (tapered) shape of the towers, the corner legs do not lie exactly in the planes of the tower faces; hence the bracing members do not lay flat on the main legs. It may, therefore, be necessary to bend the ends of the bracing members so that they do lie flat on the main leg. Obviously the methods employed by the fabricator will also affect the detailing. If the factory is highly mechanised and using numerical control machines, then any operations involving manual labour, such as bending, are going to be expensive and should therefore be avoided as much as possible. As mentioned previously, welding is also an expensive operation, and it may pay to look to alternative methods of providing a particular detail, even though it adds more weight. The final very relevant factor in producing connections is that of trying to save as much weight as possible by careful detailing. Bad detailing not only causes failures during tower testing but can also add as much as 10% to the weight of the tower.

during shipment or storage while it is bundled together, it is necessary to treat the bars to prevent storage stains. The accepted method is to dip the bars in a dichromate solution after galvanising but before bundling, and this should prevent the formation of white stains for about 3 months.

5.8 Tower testing

5.7 Tower fabrication

Some aspects of tower fabrication have already been considered, but, nevertheless, it should be stressed that tower design and detailing must consider likely fabrication and construction problems in deciding on the overall concept, to avoid expensive or time-consuming elements which may have an effect on the final cost or contract programme. As stated in the preceding text, more and more fabricators are using numerical control machines for cropping, punching, drilling and hard stamping of bar marks. These machines are not capable of bending bars or doing other press work, such as open or close flanging, flaring etc. All these operations require the use of a press and may also require the bars to be heated before the pressing operations to avoid cracking or hardening of the steel. Fabrication tolerances are of prime importance. It is normal practice to specify the tolerances required at the tender stage, so that the fabricator is fully aware of what is required of him. Unfortunately, these tolerances cannot apply to the rolled sections, as these are invariably covered by national or international standards. However, general fabrication tolerances are not, so it is necessary to ensure that these are fully understood by the .fabricator. These tolerances would cover such aspects as hole diameters, edge and end distances, hole centres, lengths of bars and straightness of bars. Most towers for transmission lines, these days, are galvanised using the hot-dip process after fabrication to provide protection against corrosion. If the galvanised steelwork is to be exposed to high-humidity conditions
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

A tower designer is one of the few structural engineers who has to put his designs to the test. The purpose of a tower test is to prove the adequacy of the design of the majority of the members in the tower. To do this, it is necessary to select a series of tests encompassing the design cases which prove critical for the various members. In this way, it is possible to ensure that the majority of the members are subjected to stresses as close as possible to their maximum design values. To achieve the above, it is usually necessary to subject a tower to six or more tests. It is often argued by engineers that the effects of a series of tests on one structure are cumulative and may, therefore, lead to a premature failure towards the end of the testing. For this reason, and because of the high cost of each test, let alone the costs of a retest due to 'test fatigue' failures, the number of tests should be the fewest necessary to prove the adequacy of the design. It is not unusual to complete the test series with a destruction test, which is generally carried out under normal loading conditions. It is a normal requirement to have to test at least one tower type on a contract, and sometimes every tower type has to be tested. For testing, the tower is built on a rigid foundation at a tower testing station, and the ultimate design loads are applied by means of ropes to the structure. The loads can be applied by 'scale pans', winches or hydraulic rams; in the latter two cases, a load cell is inserted in the rigging adjacent to the structure, to enable the loads to be recorded. Testing stations vary considerably throughout the world, from the very basic to the very sophisticated which are computer controlled. It would probably, be fair to say that most designers prefer to test their towers at the very basic type testing station, where the loads are applied in scale pans, so that they can be sure that the tower is not inadvertently overloaded. They are not so happy with the sophisticated stations where they have to rely on electronic gadgetry to inform them of the loads being applied to the structure. One of the drawbacks with the computercontrolled stations is that they continuously strive to keep all loads at the demanded level and, therefore, there is a continuous 'hunting' situation, brought about by the fact that if one load point is adjusted slightly it then affects many of the other load points which, in turn, try to adjust. The loads have to be applied for between one and five minutes at 100% loading, depending on the requirements of the contract, and most designers will admit that this period seems like a lifetime! Many of the testing stations offer the facility to strain gauge members in the tower, should this be required. Strain gauges have certainly improved over the last few years and can now be linked to a computer for a continuous readout. They have, however, proved to date not to be a great aid on testing; although they are quite accurate at measuring strain where the stresses are uniform across the bar, but, when this is not the case, i.e. for eccentrically loaded members, it is not possible to position them accurately enough to measure the average stress.
169

6 6.1

Foundations Introduction

Foundations for transmission-line towers are often very simple but they can, conversely, be sophisticated in concept, and they can be large. The early transmission lines were small in size, and so the foundations were also small and the design was uncomplicated. As the size of construction increased above ground, so did it also below ground, but there was very little modification to the basic design principles in use. The main change in the UK probably occurred in the early 1960s, when the 400 kV (L6) transmission system was introduced. The increase in loads meant that transmission tower foundations were becoming major works and, in addition, both the design and the construction of foundations were becoming complicated, owing to the routing of the line (perhaps for amenity reasons) through ground conditions which were often less than ideal. Reinforced concrete replaced mass concrete as the material used for foundation construction, and site investigations were carried out as an accepted preliminary to many designs rather than as something which would only be done in exceptional circumstances. Traditionally, foundations had been designed on what was considered a fairly conservative basis, so that the foundation would be suitable for the majority of soil conditions in which it was expected to be used. There was, thus, little requirement for soil mechanics' expertise on site, decision making could be left to field staff, unless something was obviously wrong. If, when the excavation was opened, water or soft ground were encountered, then the normal practice was to cast a pad over the complete base of the excavation, and to cast above that the normal foundation. This ad hoc practice effectively halved the bearing pressure that the foundation exerted on the soil, and provided, without calculations, a foundation suitable for submerged ground conditions. During construction of the CEGB 400 kV transmission system, although the specification only detailed one soil condition, the contractors tended to produce as an obvious corollary to the standard foundation, the formalised design for a submerged foundation. In addition, the design engineers would study the route maps of the lines in conjunction with ordnance survey and geological survey maps. Using the unpublished copies of the master survey maps available at the Geological Museum they could find details of bores taken to establish the survey. The designer had thus an indication of areas of ground along a route which were likely to be wet or soft, and he would arrange for the construction department to carry out penetrometer tests to check the conditions. If the ground was suspect, CEGB would employ specialist site investigation firms, who carried out standard civil engineering site investigations, on the basis of the results of which suitable foundation designs were prepared. These could be, for example, extended pads, rafts or piled arrays. Foundations for transmission-line towers differ from most other foundations because, in addition to having to resist compressive and laterally applied loads, they also have to resist uplift. There had been little investigation into the uplift resistance of foundations. The overhead-line industry used an empirical formula to design the foundations, but carried out very few, if any, tests to prove the designs. The general civil engineering industry, if it required foundations to resist uplift, would use concrete dead weight to counter the full tensile load, or would consider the use of tendons extending deep into rock, and test loaded to prove their
170

capacity. In the last two decades, in the UK, Europe and the USA, there have been an increasing number of studies carried out on tower foundations, and transmission-line engineers are beginning to realise that foundations may be more complex in their behaviour than originally had been considered. It is to be hoped that this resurgence of interest continues as, once foundations are constructed accurately in their correct locations, the rest of the transmission-line construction is above ground, and it should be able to proceed to a predetermined programme.
6.2 Foundation loads

Loads from the conductors and from wind on the tower enter the foundation through the stub which is an extension to the tower leg. The loads most frequently are resolved into vertical uplift, vertical compression and horizontal forces acting on the top of the stub parallel to the faces of the tower. The horizontal forces considered in the majority of UK foundations are those called 'bracing shears' (they are the horizontal components of forces coming down the bracings) and they are used to design the chimney or column of the foundation. When raft or piled foundations are being designed then the horizontal forces used are the total shears (they comprise all the horizontal forces in the tower occurring at any leg: the sum of the bracing shears and the horizontal component of the leg load). The range of foundation loads are shown in Table 9, for
Table 9: Typical values of foundation loads for UK standard constructions Construction Tower type Vertical compression
t

Vertical uplift
t

Horizontal shear
t

132 kV 0.175 in 2 L4(M) 132 kV 2 x 0.175 in 2 L7 275 kV 2x0.175 in 2 L3 400 kV 2 x0.4 in 2 L8 400 kV 4 x 0.4 in 2 L6

suspension 30 angle suspension 30 angle suspension 30 angle suspension 30 angle suspension 30 angle

40 101 54 134 71 170 97 215 143 312

36 97 38 115 44 139 57 164 99 235

1.5 3.0 2.6 6.2 1.9 6.3 5.2 12.2 9.0 22.4

Loads, for standard height towers.

suspension and 30 angle towers for a variety of UK standard constructions. For comparison, typical values of uplift and compression loads found for many constructions used worldwide are given in Table 10. Most towers used on a line will be suspensions, and so the majority of foundations will be designed for loads varying between 20 and 120 tonnes. Foundations in many overseas countries will
Table 10: Vertical components of leg load typical values
Construction voltage Single circuit and tower type uplift compression t t 132 kV suspension medium angle 220/275 kV suspension medium angle 400 kV suspension medium angle 500 kV suspension medium angle 400 kV heavy angle or river crossing 20 35 30 60 50 75 95 170 25 40 35 75 65 90 120 190 Double circuit uplift 35 65 65 95 90 185 165 355 415 compression 40 80 80 110 150 240 220 385 490 740

590

IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

be smaller than those used in the UK, but there are some constructions and some territories where loading conditions are more onerous, producing foundation loads greater than the British ones. Discussion above has been related to the self-supporting tower. A guyed tower, obviously, is different in that the foundations for the mast(s) are designed basically for compression and shear loading, while the uplift forces are resisted entirely by the anchorages provided at the base of the guys. The fact that these are set at a fairly steep rake angle from the vertical can affect the design concept used for the anchors, but the lack of shear forces on the guys means that only axial loading conditions have to be considered. There is no compression loading on the guy anchor, and so there is much less dead weight in an anchor foundation than in the average self supporting tower foundation. It is necessary, therefore, to ensure that the guy anchor is designed reasonably conservatively, as there is no inbuilt safety factor from a relatively heavy foundation.
6.3 Foundation design

There is no single internationally accepted procedure for designing overhead-line tower foundations, although there is, luckily, some measure of agreement between engineers from different countries. Design for compression load is generally as would be expected, and is as carried out for normal civil engineering. It is the first move when designing a foundation, as it defines the minimum base area required. Fig. 18 reproduces the UK (ESI) specification for design of tower foundations [51], where only one basic soil condition is listed. The maximum allowable earth pressure is quoted as 345 kN/m2. This may be seen to be a net value, as the effective weight of the concrete in the foundation is assessed to be its excess over the weight of soil displaced. (Other specifications, in particular the German VDE Standard [52], define bearing pressure as the gross value; in. this case the applied load on a stratum is calculated from a summation of loads from leg, from weight of concrete and from weight of soil above that stratum.) The method most frequently used to dimension foundations to resist uplift forces is to equate the ultimate load to

the weight of soil contained within a frustum, rising to the ground surface from the base of the foundation. ESI Standard 43-4 says that the half angle of the frustum shall be 30 and gives values for the densities of earth and concrete. Using the base area found from the compression calculations, the assessment of uplift resistance determines the overall depth of the foundation. In the UK the foundations most frequently consist of a concrete pyramid at the bottom, above which is a concrete chimney, or column, surrounding the stub up to and just above the ground surface. The shape of the pyramid is controlled normally by two angles, 45 and 70, measured between the faces and the horizontal. The faces are not nearer the horizontal than 45, to prevent the need for reinforcing the concrete, they are not further away than 70, to allow the uplift frustum to be assumed from the base of the pyramid. The size of the chimney depends on the size of the stub and on the shear forces assumed to be applied to it. All specifications require concrete cover to the stub to be at least 100 mm, which indicates a minimum chimney size. Then, either assuming tensile strength in a concrete column, or using reinforcing steel to provide the tensile reinforcement, the column size and the area and number of required steel bars can be computed. This possibly requires a larger concrete area than that needed for cover. ESI Standard 43-4 considers that the lateral earth pressure from the soil beside the foundation will provide some relief to the bending moment in the chimney; the value of the ultimate lateral earth pressure is restricted to 240 kN/m2. A cap providing rain runoff is provided at and above ground surface. Stubs are anchored into the foundation concrete usually by bolted-on cleats, although other methods can be used. The practicality of installation of the foundation is considered; any construction joints are detailed to avoid planes of weakness. Although the ESI Standard only describes one ground condition, it does indicate that, if necessary, for uplift foundations, consideration shall be paid to the effects of buoyancy. In this case, a soil density of 0.96 Mg/m3 and a concrete density of 1.25 Mg/m3 are used, but the other design parameters remain as in the specification. In actual

DESIGN PARTICULARS FOR TOWER FOUNDATIONS A1. STANDARD FOUNDATIONS A1.1 Pyramid Type Foundations Maximum allowable earth pressure on concrete foundations under specified maximum ultimate loadings, including factor of safety Ultimate lateral earth pressure Assumed weight of earth (see Note 1) Assumed weight of concrete (see Note 1) Assumed angle to the vertical of sides of frustum of earth resisting uplift for foundation design Maximum angle between base and side of concrete foundation for uplift conditions NOTES: 1. When assessing the suitability of normal uplift foundations the nett dead weights of concrete and earth shall be used in the calculations and an allowance made for buoyancy effects where applicable. For normal compression foundations ultimate loads shall be obtained assuming the added dead weight (over displaced earth) of concrete multiplied by the appropriate factor of safety. Fig. 18 Extract from ESI Standard 43-4
171

345 kN/m 2 (7200 Ibf/ft 2 ) 240 kN/m 2 (5000 Ibf/ft 2 ) 1.6 Mg/m 3 (100 Ib/ft3) 2.25 Mg/m 3 (140 Ib/ft 3 ) 30 70

The lateral earth pressure acting against the foundation is assumed to vary linearly with depth from zero at ground level to a maximum at the top of the foundation frustum. Ultimate design methods shall be employed and the stresses in the steel including factors of safety shall not exceed the elastic limit.

IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

fact, for such a submerged foundation, the base size of the foundation would normally be increased, effectively providing about double the area used for the normal foundation base. This came about because submerged foundation designs normally consisted of a pad cast below and reinforced into a normal concrete foundation. In Table 11 extracts are given from two international overhead-line specifications. The first, detailing four different sets of foundation design parameters, is perhaps more
Table 11 : Specified design parameters
Foundation or soil description Assumed angle to vertical of frustum of earth resisting uplift. degrees 30 15 10 5 45 33 34 30 25 20 10 Assumed soil density. kg/m 3 Assumed soil bearing pressure, kN/m 2 350 200 200 100 2000 900 700 500 300 100 100

may not have been the prime foundation when the PL1 (132 kV) Specification was written for the CEGB in 1928,

Normal Heavy Extra heavy Submerged Solid rock Shattered rock Dense sand and gravel Compact sand/firm clay Medium clay Loose sand Soft clay

1600 1600 1600 1000 2300 1800 2100 2000 1700 1500 1400

Fig. 19 Foundation formwork for pyramid and chimney foundation for one leg ofaD 30 tower

common in the extent of its requirements; the second, which has seven foundation types, is one of the more complex lists which has been specified. Using soil densities, bearing capacities and frusta of the half angles specified, it is quite possible to design a range of foundations of similar shape, but progressively increasing in size. This is not normally done; varying types of foundation are designed to suit both the type of ground conditions in which they will be used and the design parameters specified. These different types of foundations will next be described, but then it is necessary to determine how the decision will be made as to which foundation will be put in which location during construction.
6.4 Types of foundation

Certain forms of foundation design and construction are more suited to one particular condition than to another. The list of different foundations given in Table 12, while
Table 12: Types of foundation
Pyramid and chimney Pad and chimney Grillage Concrete ball ('Malone' anchor) Rock (pad and chimney) Piled Anchor arrayrock ground Augered shaftsmall diameter caisson (also hand dug etc.) Side bearing foundation (for single member supports) Although the above list may be considered to be of foundations for self-supporting structures, they can all, with some little modification, be made suitable for guyed towers.

but it certainly has been for all the more recent constructions. The foundation shape is neat, and there are no areas where doubts can be expressed about the design or the method of construction. It has an advantage when it is repeatedly used; provision can be made for convenient steel formwork, which is in any case required for small foundations where the minimum plan area of the excavation is greater than that of the concrete. The pyramid minimises the concrete volume and, because of its shape, does not require reinforcing steel. The formwork provides a control over the volume of concrete used and the concrete can be fair-faced. The chimneys can be parallel sided, or they can increase in width from the ground down to the top of the pyramid, thereby increasing concrete volume, but possibly saving on steel reinforcement in the column. The disadvantage of this type of foundation are twofold; the stripping of the steel formwork from the pyramid and chimney is time consuming, but, more important, the foundation is surrounded by backfilled material, and so its uplift resisting behaviour cannot be as good as a foundation cast against the undisturbed soil. 6.4.2 Pad and chimney: This foundation is used very frequently in Western Europe, either with single pads or with multiple ones in the 'wedding cake' form. It tends to contain more concrete than an equivalent pyramid foundation, but savings can be made by not having to provide pyramid-shaped formwork and also, possibly, in the time of casting blocks compared with pouring and vibrating concrete into the pyramid. The reinforced concrete pad also becomes economical if the base size is fairly large; then the excavation does not have to be oversized and the concrete can be cast directly against the soil. This form of construction is more efficient at resisting uplift forces than is one cast within formwork. Further, by undercutting into the virgin soil, additional advantage can be gained. The German VDE Specification is one of the very few specifications, if not the only one, which allows quantification of the benefits.
1EE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

not comprehensive, covers many of the types which are in use. 6.4.1 Pyramid and chimney: As stated earlier this is the foundation most frequently used in Britain (see Fig. 19). It
172

6.4.3 Grillage foundation: This was the first foundation type described in CEB Specification PL1 and it was used quite extensively on earlier transmission lines in the UK. It is not specified now in the UK, but it has been designed and constructed elsewhere in mountainous areas, and in those parts of the world where materials are not readily available for concrete manufacture. In essence, it consists of steel members bolted or slotted together to form a pad. This is generally open, that is between each steel angle or plate there is left a space, it being assumed that arching of the soil occurs between the steel members. The pad is attached to the tower leg either by a tetrapod or even by a single stub member. This latter form was used for the smaller constructions; often a shear plate or concrete block would be placed near ground level to aid lateral resistance. The CEB requirement was for the top 2 feet (600 mm) of the stub below ground, and the 6 inches (150 mm) above, to be protected by concrete from corrosive attacks, with the rest of the steel galvanised or painted. Nowadays, most grillage steelwork is galvanised and painted; concrete muffs to prevent corrosion are rarely used, but, generally, where the grillages are used the ground will be drier, and thus less aggressive than the ground in the UK. Grillages are not suitable if the soil can flow between the steel members, and so they cannot be used with high ground-water tables. 6.4.4 Concrete ball (Malone anchor): This was the second foundation described in CEB Specification PL1, and was formed by the use of explosives dropped down a small hole to form a ball of minimum diameter 24 in (600 mm). A considerable number of the Malone anchor foundations were installed, but they became viewed with disfavour, mainly because there was a lack of certainty over the true formation of the belled cavity. The explosive enlargement technique has been used more recently in North America to provide expanded bases both for guy anchor foundations and also for some augered shafts. There are still problems over the control of the formation of the bulb, connected both with the correct formulation and design of the explosive charge and, more importantly, with the engineering supervision of the construction work. 6.4.5 Rock foundations (pad and chimney): Many overhead-line specifications, including ESI Standard 43-6 [53], indicate that foundations can be constructed by making a hole in rock and fixing the stub into the hole with concrete. Some specifications quote values of rock shear strength or enhanced frustum angles if rock is encountered. It is generally accepted that a foundation cast within and against rock is going to be more resistant to uplift forces than is a foundation constructed in soil. Although, in theory, the rock foundation could be achieved by drilling a large diameter hole into the rock and setting the stub and cleats into it, this is rarely done in practice; normally a hole is excavated in the rock, and into this is cast a pad and chimney. 6.4.6 Piled foundations: The pile is the classical expedient for transferring a foundation compression load from the surface, through soft surface soil layers to some firm stratum at a greater depth (Fig. 20). Piles can obviously also provide uplift resistance, which effectively comes from the skin friction on the circumferential face of the pile. If the pile is relatively short, because, for example, rock occurs at say a depth of 6 m and above that is only peat, then the design of the piles to take the compressive load is
1EE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

simple; but, because such piles will only provide a very small uplift resistance, this becomes the controlling parameter in the foundation design. Then many more piles

Fig. 20 Completed foundation pile cap for one leg with stub setting template in position

will be required in the array than would normally have been expected. Piled foundations can be designed in two different ways to resist the loads applied by transmission-line towers. The method favoured by the British electricity authorities [54] is to provide a fully stabilised (tripodal) array of piles set in a cap in which the stub terminates. The triangulation ensures that the loads from the tower can be resolved into mainly axial loads in the piles, which is obviously a great advantage, as any bending in piles decreases their efficiency in tension and compression, and also increases the difficulties in designing the piles which for overhead lines are relatively small in diameter. Pile arrays are elsewhere frequently designed with piles mainly raked parallel with the leg, and with horizontal forces resisted by soil passive pressure on the piles and on the faces of the pile cap. There are many types of pile which have been used for overhead-line foundations. Probably that used mostly in the UK was the concrete-shell pile which was driven and then filled with reinforced concrete. The driving equipment was relatively heavy and there were at times problems with access, but the use of shells which could be extended or reduced if the set was achieved later or earlier than was expected allowed flexibility at site. If precast square-section reinforced-concrete piles had been used, then it would have been necessary to have precise knowledge of the effective depth of the piles to allow their accurate casting. Precast piles are more suited to a single major site than to an overhead line, where there would be problems over the transportation of big reinforced concrete columns. Bored cast in situ piles have been used quite extensively. The construction equipment is light, but the rate of progress is relatively slow. The system suffers from a disadvantage that temporary casings are used and that, when these are withdrawn in soft soil below the water table, there is a possibility that necking of the pile can occur. Another relatively light construction system was the base-driven thin-walled steel tube, which was filled with reinforced concrete after drilling. These piles were normally 300-350 mm dia., as opposed to the 450-500 mm dia., shell piles or bored cast in situ piles. In addition, both small diameter ( ^ 1 5 0 mm) root piles
173

and large diameter (^1000 mm) caissons have been used for transmission tower foundations. As with any other form of foundation, the method chosen for its construction will be dependent on a combination of technical and commercial assessments. 6.4.7 Anchor array: In all types of ground it is possible to provide a foundation consisting of an array of anchors (see Figs. 21 and 22). Root piles have been mentioned for poor

Fig. 21

Tops of root piles prior to encasement in concrete blocks

anchors, which are quite short, are installed vertically, and terminate within a cap which is set into reasonably firm material (to resist lateral forces) and into which the stub is set. (b) Ground anchors: Similarly anchors can be constructed in soil. In the last decade, ground anchors have become accepted for civil engineering work [55] such as the tie-back of walls of deep excavations and the anchorages for bridge abutments. Ground anchors have also been used extensively in North America for guy anchors, and to a lesser extent for foundations for self-supporting towers. It appears from the work published by Ontario Hydro [56] that the ground anchor will prove of most general use for overhead lines constructed in compact dry ground conditions. Then the results obtained from the anchors have been in excess of the expectations of the designer. In soft ground, anchors have been constructed, but they are often more difficult to form, and they have to be longer, to develop strength, than the anchors in compact material. In soft ground, arrays would normally be of a fully stabilised design; in dry compact ground the anchors and the cap will satisfactorily resist lateral forces and the anchors can be constructed mainly raked parallel to the leg. Passive anchorages have been used for most overhead-line contracts, although general civil engineering work often considers the use of stressed tendons. This obviously can be of advantage if it is required to increase the consolidation and strength of a soil section, as the prestress in the tendon helps to improve the ground. UK electricity authorities have used tendons grouted into rock and stressed to working uplift load; the intention is that no movement of the anchor cap and tower leg can take place until the uplift load exceeds the prestress load (or working load). It is not believed that ground anchors have been used for transmission-line construction in the UK. 6.4.8 Augered foundations: Although described as augered foundations, this class can perhaps be most conventionally split into two subsections, the small-diameter shaft and the caisson, or large-diameter shaft. It is possible to construct a caisson by auger, by excavator or even by hand: (a) Small-diameter augered shafts: These shafts, which are drilled to a diameter of 300 to 400 mm, most frequently, and to a depth of 3 to 6 m, will normally be set in an array finishing in a cap at each tower leg. In the simple case, for a light transmission construction, one shaft to a leg can suffice; heavier constructions require say four or even eight shafts for each suspension tower foundation. The shafts can be installed vertically, alternatively they can be splayed about the rake of the tower leg. The alternative chosen depends on the assessment of the relative cost and convenience of construction vertically or at a slope, compared with the extra cost of designing and reinforcing a shaft which has to resist a larger shear force. The shafts are normally reinforced with steel bars in the form of a cage, similar to the reinforcement provided for piles constructed for overhead lines. The vast majority of the augered shafts installed have been parallel sided, although consideration has been given to the use of belled, or underreamed shafts. The latter type of construction it was believed would provide a positive physical key into undisturbed ground and, thereby, provide a sure uplift resistance. Investigations carried out by the CEGB [54] (on augered shafts 760 mm dia. and 7.5 to 9 m deep) indicated that underreaming was not always a rewarding technique. There were difficulties if the shaft was not vertical and if the ground conditions were less
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

Fig. 22 Round stub held in position prior to pouring concrete in foundation cap above root piles

ground. They consist of a single high-yield steel (HYS) bar grouted into a hole approximately 150 mm diameter. Similar types of construction can be used in rock or in soil: (a) Rock anchors: The old fashioned foundation in rock (stub anchored in pocket in rock) has been mentioned: to form this, quite a lot of compact material is removed and replaced by concrete. The anchor, on the other hand, is produced by removing a relatively small volume of rock and grouting a bar in the hole. Normally, the tendon will consist of a nonprestressed, deformed HYS reinforcing bar set in a hole perhaps 25 to 50 mm larger in diameter. The anchors have been used quite extensively in many countries and have been successfully tested. They have proved most suitable for use in all sorts of rocky ground conditions; not only for the solid rock, if it can be found, but also for weathered, layered and faulted rock. The anchors appear to stitch the rock layers together, and it is unusual for an anchor to fail before the design load at which it is expected the tendon will start to yield. Normally, the
174

than ideal. It took a finite time to change the drilling tools from auger to underreamer and probably back again, and the time taken was often more than had been assumed. With sticky clays, problems were found with clearing the spoil from within the belling tool. It was decided that it was more efficient to construct a slightly longer straight shaft rather than attempt to bell. Similar conclusions were reached by Ontario Hydro who initially had belled the auger shafts they were constructing for 230 kV lines. Fortuitously, the construction was through terrain most suited to allow the construction of the shafts and the underream. When more heavily loaded foundations were required and the ground conditions were not so ideal for construction, there was a shift away from the belled shaft concept and the designers allowed the use of straight shafts; albeit possibly of greater depth. Research investigations were carried out and the results increased the Ontario Hydro designers' confidence in the straight shaft. (b) Caisson (large-diameter shaft): It is probably expected that the caisson will be a single, generally, vertical shaft constructed below a tower leg, or even beneath a complete tower if the foundation is made that large. The caisson thus may have a diameter from say 900 mm up to almost any conceivable value. Augers have been used to construct caissons; it could be said that the CEGB augered foundation for the 400 kV L6 construction was a caisson foundation. The caisson foundation has been used quite successfully in certain parts of the Middle East, when difficulties were found when attempting to auger smaller diameter shafts. Hand dug caissons (lj to 3 m dia. and 10 to 20 m deep) have been constructed in Hong Kong for some of the 132 kV Tsing Yi Crossing tower foundations and for foundations for the recent 400 kV line. Caissons 2.5 m dia. and 30 m deep were used on the Oosterschelde crossing and caissons 12 m dia. and 100 m deep have been constructed in Bangladesh for a river crossing; in both cases, each caisson supported a complete transmission tower. The shaft is normally constructed of concrete suitably reinforced to resist all the applied forces. 6.4.9 Side bearing foundations (for single member supports): The side bearing foundation, a single monolithic block into which the structure was set, was used on some early UK narrow-based transmission lines (33 kV to 66 kV). As transmission voltages and construction sizes increased, so the side bearing foundation was relegated to use for distribution pole lines. There has, in recent years, been an increase in interest over the use of single-member structures to support high-voltage transmission lines and there have been many new investigations into the possible behaviour of the foundations. Study of CIGRE conferences in the early 1950s confirms that Berio (Italy), Lazard (France), Simpson (UK) and a Belgium group were investigating, in detail, the side bearing capacity of foundations. Although North American utilities and consultants had constructed caisson foundations to support the first of the modern generations of single member transmission structures, the new surge of interest occurred some five years ago with investigations in France, Poland [57] and by EPRI in the USA [58]. The North American techniques are computerised, and it is claimed they allow the construction of foundations with a much smaller inbuilt factor of ignorance than was possible previously. They are basically concerned with the design of laterally loaded piles or caissons. The French, to some extent, have been treading the more traditional Western European path which appreciates the increase in lateral resistance which can be provided by extending the
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

size of a base pad below the side bearing foundation block. There are still many problems related to the correct design of these foundations, including the correct assessment of soil type and the acceptance of allowable movements. These problems are relevant to all transmission-tower foundations and will be discussed in following Sections of this paper.
6.5 Site investigation

As stated previously, site investigations were not frequently carried out before constructing foundations, but, with increasing foundation loads, an increasing proportion of poor ground in which to construct the foundations and with an increasingly large proportion of overhead lines being built in countries with little record of the ground conditions, it has been necessary to revise this practice. In the UK now, the CEGB carries out, for a proposed route, a study of the topography and of available geological information, and on that basis decides where it will carry out ground investigations. The main investigations will be concentrated on angle tower positions, partly because the angle towers exert larger and more constant (deviation) loads than do intermediate towers, and partly because the angle positions are critical from the point of view of the sections of line on either side of them. The sites are fully investigated using boreholes, in situ and laboratory testing, and the results of the investigations are compared with the expected or presupposed variation of the strata forming the subsoil. If there are inconsistencies, then further bores will be sunk to detect strata changes. Between bore holes or at certain locations where more information is needed on the variation of soil properties across a site penetrometer tests are performed. These may be of two basic types; the more sophisticated is the use of the cone penetration test with electrical printout of toe pressures and skin friction, the simpler device is the use of a percussion hammer driving a probe into the ground, the consistency of which is assessed from the time taken to drive the point a predetermined distance. In the UK, engineers benefit from knowing the location of the proposed tower positions, but even so it is considered impractical to propose putting down a bore hole at each position. Care is taken over choosing where investigations shall take place; it is not sensible to carry out an investigation every five spans or every 2 km: you can, with that sort of decision, end up by doing investigations only where the ground is good and completely ignoring areas of doubt. In some parts of the world, clients, prior to issuing a tender, carry out preliminary investigations of the ground likely to be encountered and quite often have a considerable number of full site investigations carried out to allow assessment of different ground conditions. When the transmission line is being constructed, it is often a requirement of these clients that full site investigations are carried out at each approved tower position. Many clients have a less rigorous requirement than this and require a certain number of investigations to determine the soil types and, possibly, penetrometer testing in between the bore-hole positions to give information on the changes in strata. It cannot be envisaged that, for a normal transmission line, foundations will be separately designed to match the ground condition at each and every location; there must be standardisation into discrete foundation types. The purpose of the site investigation is to allow an engineer to decide which type of foundation should be installed. It also should allow him to assess any problems there might be in achieving its construction: is there ground water, is the
175

ground loose so that excavations will collapse, are there boulders, what is the rock like? Normal site investigation techniques give information which allows (a) the identification of soil type and (b) the assessment of soil consistency and possibly of soil shear strength. The second fact allows calculation of the bearing capacity of the soil, the first fact allows the investigator to appreciate the sort of behaviour to be expected from the ground. Neither directly allows analysis of uplift resistance. There is thus difficulty with the specification quoted in Table 11, where the 'heavy' and the 'extra heavy' foundations have identical soil densities and bearing capacities, but different uplift capabilities. How can these be identified? If site investigations have been carried out, then it is possible to consider compression and uplift behaviour on the basis of simple soil mechanics' analysis. The study of the soil type will also allow the investigator to consider more accurately changes in foundation behaviour due to seasonal variations in the ground conditions.
6.6 Soil science related to foundations

The more recent investigations into foundation uplift behaviour have been based on analyses using the principles of soil mechanics [59-61]. Although different actual formulas have been provided by each of the groups, in fact the analyses all attempt to integrate the soil shear strength over some predetermined failure surface. The use of any one of these formulas gives results which more frequently will be nearer the true strength of the foundation than those obtained using empirical methods. But the soil properties have to be known and they often vary quite considerably over the extent of a foundation. This variation gives rise to problems which are difficult to solve: what 'in between' value of strength properties is assumed, because it must be remembered that failure of the soil will be progressive? These investigations have had one main benefit, which the industry has gained even if it has not used the formulas. The results have proved that the frustum theory, while useful, has to be taken with a considerable pinch of salt. The strength of foundations is more closely related to (depth)2 not to (depth)3, and so deep foundations are not as strong as might have been expected. Plan-area/depth ratio is an important consideration, as is uplift bearing pressure. Similarly, as it is shear strength over a surface which is important, there can be changes in uplift resistance if some of the soil layers become weaker [62]. This happens if they get wetter, and this can take place without total submergence. Once water starts to soak into the ground, that ground is less strong. Soil strength is also related to its relative density. If a soil is loose, it is less strong and, in addition, there are more pore spaces, which allow a larger volume of water to enter the soil structure. Compaction of the soil round and above a foundation must be of benefit to the subsequent behaviour of that foundation.
6.7 Foundation movement

equated to tower ultimate loads. The phrase most frequently used to define bearing pressure for the foundation strictly does define this (Maximum allowable earth pressure under specified maximum ultimate loadings including factor of safety. CP 2004 [63] defines allowable pressure as that at the base of a foundation taking into account ultimate bearing capacity of the ground, the amount and kind of settlement expected and the capacity of the structure to take up that settlement.) Uplift design tends to be equated to the ultimate value and little or no attention originally was paid to possible movement at that load. Test results indicated that movements of 100 to 200 mm could be expected in some conditions before a maximum load was reached. Some towers obviously can withstand more differential movement between foundations than can other structures. It depends on base width and on the tower design. Acceptable foundation movement also will be related to whether one leg only is distorted or, less critically, if say both uplift legs move in unison. It seems unlikely that a new working load design method will be introduced for dimensioning foundations to resist uplift, it seems more likely that some factor to reduce movement will be introduced into the present design practice. It may also be that certain types of foundation will be preferred because they move less (a cast direct foundation moves less than one surrounded by fill; anchors and piles move less than pyramids), but when these foundations fail they do so in a brittle manner. Although they may satisfactorily resist a load less than the ultimate with no movement, if the ultimate load is applied continuously the foundation will fail. A block foundation will move more initially, but its movement will compact the soil above and to the side of it, and this can increase its resistance. If the ultimate load is maintained the foundation will fail, but probably not as rapidly as one experiencing brittle failure.
6.8 Foundation testing

It is a natural development, in view of the more recent aspects of foundation design analysis, that testing of foundations is becoming a more frequent requirement in modern specifications (see Fig. 23). It is realised that although foundations can be designed using accepted methods, neither the specifier nor the designer is really sure how the foundations will behave in practice in the local soil conditions found on site. Even though foundation tests are carried out on new contracts, there will still be doubt over the repeatability of the results, and what happens when the ground conditions vary, which they will. It will only be after considerably more information is obtained from test results and made available internationally that

The empirical designs and the soil mechanics' designs have all been based on failure of the foundation: the ultimate load. Although originally the overhead-line engineer's concept appeared to be that at ultimate load each item of the line, conductor, tower, foundation, would fail instantaneously, now there is a belief that there should be an order to failure, and generally that the foundation shall remain intact even if a line and the structure have failed. This could mean that foundation working loads should be
176

3 ^
Fig. 23 Rig for uplift testing rock anchors IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

the transmission engineer will be more certain about what foundation he should install in what ground conditions. 7 Construction

7.1 Introduction and history In the recent past, and more particularly in the last ten years, there has been an unprecedented shift in the relative economic growth rate from the traditional industrialised nations to those of the Third World. The reasons underlying this fundamental change in world economic balance are well documented elsewhere, but suffice to say its effects on the power-line construction industry has been considerable, in so far as, electricity being one of the basic vital components of industrial development, the work opportunities for firms in the industry have moved almost entirely from their own relatively secure home markets to those of the developing world, which are highly competitive. For those companies, many with their origins in Europe, who have taken up the challenge, the transition has by no means been an easy one and, in some instances, both complex and costly. Although some had, in previous years, successfully carried out major overseas contracts, the revenues thus derived had not provided the majority of their turnover and this, combined with a sufficient work load at home, meant that they could be more selective in undertaking only the more potentially attractive projects. However, a virtual relearning of the art was required to accommodate the shift from their well known commercial, fiscal, technical, topographical and climatic environment to often little known situations, where all or any of these factors may well change, not only from territory to territory, but from contract to contract. On the physical side it has been necessary rapidly to acquire the skills and equipment to successfully cope with virtually the full range of world climates and topography, while, simultaneously, the design function has been required to meet many and various specifications under highly competitive conditions. An additional major area of change has been the need, primarily as a result of competitive pressures, to acquire site labour from sources other than the contractor's home territory and, more recently, from territories outside those in which the works are to be executed. While this activity initially produces some problems for the contractor, both fiscal and physical, a knock-on effect has resulted in that such workers, armed with skills thus gained, have formed themselves into groups, subsequently to appear, on the one hand, as useful subcontractors, or, on the other, direct competitors. It has thus been the lot, frequently not a happy one, of the construction engineer, to meet the challenge of continuous changes, some covering fields of activity of which he would have little or no previous experience. 7.2 Logistics, transportation and access These terms, when used in the context of overhead-line construction, cover those activities involved in the positioning of the prescribed materials, together with resources in the form of labour, supervision, machines and associated consumables, at the appropriate site on time. The activities coming within the scope of this function can be all or a combination of individual operations, the mix and difficulty factors being principally determined by the terrain and level of infrastructure existing within the particular territory. Resulting from a review of a number of recent projects,
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

the activity list is somewhat impressive, requiring a wide range of skills on the part of those on whom the responsibility falls: (a) Provision of a base office, together with accommodation for the necessary administrative staff. (b) The acquisition of forward stores areas, the positions being carefully established in the light of existing conventional transport routes from the source of materials, often a port of entry, and the forms of transport that will be required to move materials onwards to the individual tower positions. (c) The recruitment or acquisition, by subcontract or otherwise, of the necessary labour force from either the contractor's home territory, local sources or by the use of this country nationals. When labour is imported into a territory, it is essential that steps are taken to complete all the requisite formalities in good time, particularly with respect to immigration control, residence and work permits. When employing third country nationals it may be necessary to provide for certain basic needs peculiar to their country of origin, particularly, with regard to food supplies and domestics. (d) The provision and support of accommodation, messing facilities and services together with domestic personnel, so located as to minimise the combined costs, on the one hand, of transportion of the labour to site and, on the other, the accommodation support. Such accommodation may, in some cases, be rented existing property and, in others, purpose-built, either by the contractor or local organisation. This is one activity which, in areas of extremely difficult terrain such as remote mountain routes, can give rise to high costs. It has, on occasions, been necessary to establish camps so remote as to render the only practical means of support to be by air, most commonly using a helicopter. It has, for example, been necessary to fully support under conditions of inclement weather a camp of 50 men almost entirely by helicopter at a distance of some 50 km from the nearest practical vehicular access. In such circumstances, it will be seen that the most stringent controls must be maintained to avoid massive cost increases for which no allowance has been made. (e) Provision of bulk transportation either by owned or hired vehicles or subcontract, for the movement of line materials from point of origin to main stores. This applies, not only to the tower and conductor materials from a designated port of entry, but also locally obtained materials such as cement, sand and aggregate, which may be derived from more than one source during the course of the project. (/) Transportation of materials from forward stores and accommodation areas to individual tower sites. This activity not only involves the provision of items of transport which may range from two- or four-legged, through wheeled and tracked to winged, but also the preparation of routes to permit the use of whichever of these means is most cost-effective. In general, it may be said that the greater the expenditure on access preparation, the less the expenditure on vehicles of whatever type, and it is the fine balance between these two factors which is a problem which continually besets the construction engineer. In arriving at the overall cost of given access routes, it is essential that allowance is made for maintenance, particularly where such routes may be subject to being wholly or partly destroyed due to seasonal weather conditions. (g) The provision of an adequate level of support for both vehicles and machines to maximise availability, and thus minimise construction-crew idle time due to failures.
177

To this end it is essential that support in the form of skilled men, spare parts and adequate facilities be positioned as close as possible to the location of the works as this progresses through the route and that full back-up is provided through the administrative chain to ensure that needs are promptly met. (h) When the base office for a given project is located remotely from the principal port of entry for the contract materials, in some instances it is found necessary to position an individual, with experience of customs and related importation activities, to ensure that on-the-spot representation is available to expedite clearance. The need for such representation is much dependent on the prevailing regulations and attitude of the authorities in question, and, thus, precommencement investigation into this aspect is recommended. (/) In meeting the logistical needs of distribution schemes, i.e. those of 132 kV and below, it will be found that although some or, in the case of the larger schemes in the Middle East, all of these activities will need to be carried out, the difficulty factors usually are much reduced. By definition, distribution involves the direct supply of power to the end user, and where there are users there will at least be some level of existing infrastructure by way of roads, local suppliers etc., as opposed to the transmission line which carries power from a source often remote from the final load centre, the route of which may well be over difficult terrain. With the increasing degree of competitiveness prevailing on the majority of line-construction projects, be they transmission or distribution, the level of attention paid to the logistical aspects of the work can often make the difference between commercial success or failure of the venture. 7.3 Application of helicopters to overhead-line construction From the widely differing levels of use of helicopters by the major European and North American based contractors, the conclusion could be drawn that there is little agreement as to their suitability and/or cost-effectiveness. However, the underlying reason behind this apparent divergence is far more closely related to the type of projects carried out by particular contractors, in so far as relatively few have undertaken contracts where the conditions, either physical or fiscal, obviously lent themselves to, or prescribed, their use. In considering the relative viability of aerial, as opposed to surface, transportation, the net figure sought is cost per ton or per passenger per kilometre. In the case of the aerial option, the principal components of this cost are the sum of the all-in operating cost of helicopter complete with fuel, spares, air and ground crew, together with their accommodation, messing and surface transport where needed. In addition, there will be, dependent on the level of operations, administrative and organisation costs, together with materials handling labour. Set against this will be the actual productivity achievable in terms of tonne-km per hour, involving the practically obtainable aircraft performance under the climatic and altitude conditions existing, load factors relating to the types and quantities of cargoes to be delivered to each location and flight difficulty factor resulting from the topography and other hazards in the vicinity of pick-up and set-down points. The climatic conditions are of particular importance as, while one can plan for construction activities to fit within a predictable weather pattern where periods of unacceptable conditions are reasonably well established, in those regions
178

where bad weather is both intermittent and unpredictable, disruption of construction progress can rapidly reach an unacceptable level. This factor becomes particularly significant in relation to the daily transportation of construction crews on and off site, movement of crews and equipment between sites and the direct placing of mixed concrete in foundations. In the case of the surface option, the cost make-up primarily involves initial access route location, access making, the all-in operating costs of the equipment, consumables and support; all of which become significant in the more extreme cases of mountainous or wet terrain. To these costs must be added those relating to the maintenance of the route throughout the construction period, note being taken of the possibility of the loss of a section or sections of the route due to seasonal inclement weather. The second major factor of the surface alternative is the cost of provision, operation, maintenance and support of the transportation vehicles required, often at locations remote from the main operational base. Dependent on the terrain, but particularly in the case of wet conditions, the vehicles are both more costly in terms of initial purchase and also in respect of both spares and maintenance man-hours than is customary with the commonly used range of civil engineering plant. It is this last factor that is frequently overlooked in terms of lower availability, with resulting need for increased vehicle numbers to maintain work progress and also increased risk of work disruption due to breakdowns. A typical example of a situation where the physical conditions gave rise to the need for helicopters was that which arose during the construction of the 400 kV network in the Zagross Mountains of Iran. Owing to a combination of circumstances, the final part of the installation programme had to be carried out at such an accelerated rate, in conditions of remote and precipitous terrain, that there was no practical alternative to the use of helicopters. The quantities of materials and levels of logistic support necessary required the use of machines with sea-level capabilities, varying from f to 9 tonnes. During these operations some 210 drums of conductor, each weighing 4.3 tonnes and 1200 tonnes of tower steelwork were flown over distances up to 85 km and up to altitudes of up to 2100 m above sea level. In addition to this activity, full logistic support had to be provided, latterly under most difficult weather conditions, for 190 men in three remote forward camps. An example of the situation where the use of helicopters was at the request of the client is the current 400 kV project in the New Territories region of Hong Kong. The reasons for such action on the part of the client were the prevention of erosion of the ground surface. This becomes rapidly unstable when the vegetation has been removed as a result of activities such as access making, with consequent pollution of the all-important water-supply facilities. To this was added considerable pressure from conservationists to minimise the destruction of what little uninhabited amenity area remains within this crowded colony. In this instance, the principal use of the helicopter has been the movement of line materials, including reinforcing bars, mixed concrete, prebundled tower steelwork, prepackaged insulators and fittings, from a number of strategically positioned pickup points, each with good road access and, in specific instances, sufficient space for the operation of concrete batching plants. To a lesser extent, but no less vital, the helicopters move the construction crews together with their associated equipment from site to site, giving the opportunity for reduced gang travelling time and hence improved productivity. The choice of heliIEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

copter type is all important to the economic success of the project. In the case of Hong Kong, where the aerial movement of conductor drums and heavy stringing machines was rendered unnecessary by careful siting of the pulling locations, the Aerospatiale Lama 315B was the first choice, based on considerable previous experience of the type. This machine, although at face value not the cheapest per tonne-km, has a number of attributes which render it very suitable for close support and precise work, under conditions of erratic winds, hilly terrain and high temperatures. In essence, the principal reasons governing the use of the helicopter are as follows: {a) direction to do so by the client (b) that it is the most cost-effective method of execution of the works, either in part or in whole (c) given programme period insufficient to permit work to be carried out by alternative economically or practically available means (d) terrain over which the line is to be built unsuited to surface transportation (e) Where, although initial surface access cost may be acceptable, such accesses are liable to loss owing to the results of inclement weather (/) reprogramming due to unforeseen circumstances, giving rise to need for accelerated construction rates. In viewing the potential uses to which the helicopter may be put, it is necessary to clearly understand the two basic functions which it is able to carry out and also the effectiveness of the range of types available in each of these two roles. The primary and most immediately obvious role is that of a transporter of men and materials. In performing these functions its principal limiting factors are altitude, temperature, weather and near-ground obstructions such as trees, steep side slopes, overhead lines, buildings etc. The second, and less obvious, use is that of the small mobile crane with, although limited hook capacity, virtually limitless hook height and operating radius. In this role, its principal limiting factors are similar to those when used as a transport, but those relating to near-ground obstructions can be minimised by the use of special underslung lifting equipment. In reaching a decision as to the number and type of machines to be used in any given set of circumstances to achieve the most economic result for the load mix and work rate involved, careful consideration should be given to the likely effect of the machine limitations. In the context of line construction, the helicopter has been successfully applied to the following activities: (i) Survey (a) Initial route location and marking (b) Potential access and camp site location (c) Positioning and support of ground survey crews (d) Positioning and support of geological and soil investigation crews, together with their equipment (e) Aerial photography (/) Inertial survey of potential route profile (ii) Logistic support (a) Movement of personnel from nearest practical surface access to and from work site (b) Movement of bulk line materials from forward pick-up areas to individual tower sites (c) In the case of concrete, transportation of mixed concrete from either batching plant or ready-mixed truck directly into foundation (Fig. 24) (d) Where appropriate, transportation of prepacked concreting materials and mixing equipment to each tower site
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

'

Fig. 24

SA 315B delivering concrete from 0.3 m 3 lightweight skip

(e) Movement between sites of general construction equipment including foundation templates and formers, pumps, air compressors and drilling equipment, erecting derricks and winches, running-out blocks and pilot rope drums, sagging and jointing equipment. In this particular context it is of importance to carefully select and, if necessary, specially design the equipment used to ensure that it may be safely lifted by the helicopter type chosen; preferably in one piece, or where this is impracticable, that it may be readily dismantled and reassembled on site (Fig. 25)

Fo. 25 CH 47 C lifting partly stripped 4-bundle conductor tensioner to a mountain site 1820 m above sea level

(/) Transportation of consumables such as fuel and spare parts, together with associated specialist personnel (g) Support of forward construction camps located in remote and inaccessible areas (h) In the case of accidents, casualty evacuation (iii) Construction (a) Positioning of foundation templates and form work, reinforcing bar cages and tower leg stubs (b) Erection of derricks prior to commencement of tower building (Fig. 26) (c) Where appropriate, tower erection, either in part or in whole
179

tower attachment positioning, concrete placement and, finally, backfilling where appropriate. To permit the installation of all types of foundations with the exception of the driven displacement pile, removal of some quantity of the existing natural material is necessary and, although the quantities and configurations vary considerably with type, two main problems may arise, these being lack of ground stability and presence of water. Although, following the award of a contract, the contractor may well choose or be required to carry out more detailed soil investigation, in the time normally available, the number of sites where this is practicable will be limited, and thus the information gained will be of a random sample nature. This produces a situation where the final choice of foundation type and size can frequently only be made when the excavation has been commenced.
Fig. 26 SA 315B raising 33.5 m erection derrick

(d) Hanging of insulator sets, complete with runningout blocks (e) Running out of pilot ropes, either by laying or towing, although laying is preferred (Fig. 27) (/) Direct stringing of the lighter distribution conductors

Fig. 27

SA 315B equipped with constant-tension pilot line layer

Capacity of drum = 4 km of 12 mm diameter fibre rope

(g) Cranage for onsite repair of machinery, although the use of the helicopter for this purpose requires careful consideration, with particular regard to the safety of the personnel involved, and also the possibility of the damaging effect of dust arising from the helicopter downwash (h) Erection of poles and light distribution supports. 7.4 Foundation installation It is widely held that of the three principal activities comprising the construction of an overhead transmission line, foundation installation provides the greatest scope for variation between estimated and actual cost. The principal and somewhat obvious reason behind this is that, while tower erection and conductor installation have become well practised skills involving the use of materials of predeterminable quality, foundation installation still retains a fair amount of the unknown, owing to the limited scope for site investigation during the prebid period. Added to this is the allocation by the consulting engineer or the client of the various design types in particular conditions as they arise, which still tends to be somewhat subjective. In the overhead line context the foundation installation function contains four basic activities, namely excavation,
180

7.4.1 Excavation: (i) Pyramid and chimney, pad and chimney and grillage foundations: These designs involve the largest volumes of excavation of the various types discussed, generally taking the form of a pit approximately square in plan and of a depth greater than the length of a horizontal side. Such pits can be excavated either manually or mechanically, or by a mixture of the two methods; local economic factors combined with particular terrain conditions determining' which is the most cost-effective. Ground of poor stability will require shoring or battering and, in cases of water inundation, arrangements for pumping or dewatering will be required. Where rock is encountered, explosives still provide the best method of fragmentation, but in some territories their use is either prohibited or made unacceptably difficult, thus requiring the use of alternative methods varying from hydraulic breakers fitted to mechanical excavators to 'feathers and wedges'. (ii) Explosively belled (Malone anchor) foundations: The principle and background to this form of foundation is noted in Section 6.4.4. From the construction point of view, recent trials have been carried out in the UK in an attempt to improve charge shape design to achieve the production of acceptably consistent bulb shapes and degree of enlargement in a variety of soils and with both cased and uncased bores. However, it is considered that further study and experimentation would be required to establish a system which would gain a reasonable degree of general acceptance. (iii) Piled foundations: Overhead-line contractors have tended when piling, particularly when the displacement type is required, to call on the services of the specialist subcontractor. This situation has grown up, mainly due to the infrequent call for such foundations, tending to discourage the line contractor from investing in the equipment required and the holding of labour with the specialised skills and know-how so important to this particular operation. However, in recent years, contractors have shown an interest in particular forms of piling, namely those which are excavated by means of augers or drills, as the foundation types thus produced have become viable alternatives to the conventional foundation, rather than just being a method of meeting exceptional circumstances. The current situation relating to augered and drilled shafts is discussed in (v) and (vi). (iv) Rock anchors: The installation of the form of rock anchor discussed in Section 6.4.7(a) is a relatively simple operation and with the range of equipment presently available there is little scope for significant improvement in productivity. The majority of contractors commonly use a
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

drifter-type pneumatic drill mounted on a manoeuvrable vehicle frequently running on tracks in the form of a wagon drill. It is customary, first, to excavate the pile cap pit and, subsequently, drill the required numbers of holes using drill rod guides, if the strata is such as to cause lateral deflection of the bit, or if an array of inclined anchor holes is called for. As the quantity of grout per site is small, hand-operated mixers have been found adequate, the resulting grout being delivered to the bottom of the hole by means of a simple hand-operated pump; although, as there is little chance of disturbance to the side of the holes, the grout may be simply poured in from the top. (v) Ground anchors: Current trends in the design of ground anchors call for bores with diameters in the region of 150 mm and depths of 4 m or greater. This is an extension of the rock anchor and the principle can be applied to a wide range of soil types and conditions, more than one of which may well be encountered in a single shaft. However, as this type of anchorage lends itself to dry compacted granular soils which may exhibit some degree of local instability when exposed and subjected to the forces required for drilling, careful consideration as to the type of equipment used is necessary. In the case of ground anchors, the problems relating to straightness of bore can become more difficult as not only are the depths involved greater but also the soils may well contain, not only the smaller sand and gravel materials, but also, particularly in the case of wadies, large boulders occurring at random, and frequently so placed as to give rise to conditions likely to induce considerable drift of the drill bit. In recent years, considerable advances have been made in the design of both drilling machines and perhaps, more importantly, drill bits, so that it is now economically practical to drill both rapidly and accurately through a wide variety of soil and rock types with a single bit type; although under certain conditions such accessories as casings and facilities for the use of foams or muds may be appropriate. In the placement of ground anchors, of equal importance to the drilling activity is that of the grouting in of the anchor tendons. With this design, the quantities are such as to make appropriate the use of grout mixing machines and the depth of the shaft such that grout placement from the bottom upwards is certainly highly desirable, if not in all cases essential. This is considered of particular importance to minimise inclusion of foreign matter disturbed from the hole sides and unfilled voids resulting from air entrapment, where loose material may have been removed by the force of the air flush blast. The use of a colloidal grout mixer combined with a lowpressure delivery pump has the additional advantage of providing a facility for accurate batching of all constituents and ensuring rapid acceptable mixing to give complete hydration. (vi) Augered shafts: This type of foundation normally covers a range of shaft diameters, in the main falling between 300 and 1200 mm with depths up to 10 m depending on ground conditions, although, on occasions, particularly in North America, bores of greater diameter and depth are used. The machines available fall into two groups, the former being more suited to softer materials, namely the top driven continuous flight type and the rotary table Kelly bar machine, used in situations where higher crowd forces and torques are called for. Taking into account the wide range of soils which the overhead-line contractor is likely to encounter, the latter type is conIEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

sidered to give the best all round performance, and particularly comes into its own in the harder and rock bearing soils. As the use of augered shafts has evolved there have appeared two problems in that, first, designers and constructors have worked insufficiently closely to arrive at the most economic result, with regard to depth and diameter without resort to either special machines or machines of such weight and size as to be impractical in all but a very limited number of situations. Secondly, while there have been major efforts on the part of a number of equipment manufacturers to provide a range of reliable and effective machines, there appears to have been little real research and development into the design of auger bits, particularly in relation to those intended for use in rock bearing soils. As with all forms of excavation, the augered shaft is sensitive to conditions of wet or loose ground and it is necessary to ensure that, in addition to the basic machine, casings and other temporary stabilisation equipment are available. When, as noted in Section 6.4.8(a), the foundation design calls for belling or underreaming of the previously augered shaft, a new set of problems face the construction engineer. While there is a fairly wide range of soil types in which a vertical shaft of reasonably consistent diameter can be bored, the sides remaining intact sufficiently long to permit the placing of reinforcing stubs and concrete, the range of soils which can be successfully underreamed without immediate collapse is far more restricted. The two most common factors which prevent the belling of augered shafts are, first, on the one hand,, the presence in the soil of excessive water and, on the other, the presence of gravels in excess of 150 mm. In addition to these more obvious situations, it will be found that, in the cases of granular materials such as wady sands and gravels, results tend to be somewhat unpredictable, and, hence, trials are strongly recommended where possible. 7.4.2 Attachment positioning: Whatever type of foundation is to be installed, there remains the problem of accurately holding in position during concrete placement the attachment device required to connect it to the tower steelwork, usually taking the form of stubs or bolts. There have been few advances in what is a fairly basic function, and opinion is still divided as to whether as a general rule it is preferable to use four independent support devices positioned by measurement or similar devices interconnected by a frame in the form of a nominal template. Some contractors, on the rare occasions when the structural steelwork is available at the time of foundation installation, opt for erecting the lower part of the tower complete with the stubs prior to concreting. In particular instances where wide-based towers have been located on hillsides, two of the foundations at the lower level have been installed independently, the tower base subsequently being erected prior to the concreting of the remaining two upper legs. More recently, with the advent of towers of the size of those being built in Hong Kong on steep side slopes, the use of the interconnecting template has become impractical and thus faster, more accurate and reliable methods of measurement have had to be evolved to render independent leg setting economical. To this end, the facilities provided by a combination of the high-accuracy theodolite and electronic distance-measuring equipment with built-in computer have proved successful, and there are indications that this equipment, with the added facility of laser projection, may speed up the stub setting process.
181

7.4.3 Concrete placement: In the period under discussion, although there have been changes in the methods by which concrete is delivered to the tower position involving greater use of concrete pumping and helicopter transportation, there have been few changes in site practice, with the possible exception of greater stress being laid on the use of suitable chutes and tremie pipes. This is of particular importance in the case of augered shafts, where the effect of free-falling concrete would produce unacceptable levels of segregation and possible inclusion of soil disturbed from the sides of the bore. 7.4.4 Backfilling: The importance placed by consulting engineers and clients on quality control has increased considerably in recent years. Their attention has been directed particularly towards the particle-size range and quality of materials used, and also the degree of compaction achieved throughout the full depth of the backfill, requiring full mechanical compaction at specified minimum intervals and also, on occasions, penetrometer tests as evidence. 7.5 Tower erection In general, the tower building methods have changed little for many years, the majority still being erected by the derrick and winch system; although in North America this is the exception rather than the rule. As towers have gained in both size and weight, so have the derricks employed; although increasing use of light alloys and improved design has tended to allow greater lengths and lifting capacities to be achieved, without too great an increase in weight. As far as the UK origin contractors are concerned, the most significant change in the application of the derrick principle has been that from building almost exclusively conventional double-circuit towers to the need in overseas territories to build the single-circuit gantry-type involving the lifting of a heavy single load to the highest point on the tower, albeit with the derrick in the near-vertical position. Although the derrick and winch system is still the most commonly used, two developments, one in tower configuration and the other, more recently, in crane design, have generated, on the one hand, the need and, on the other, the facility for the introduction of new techniques. 7.5.1 Guyed 'V and portal structures: With the advent some 13 years ago of the use of guyed V-structures on 500 kV lines in territories other than North America, it was necessary for the prospective contractors to evolve economic and effective erection methods to meet this rather special case. By their very nature these towers are best erected in a single lift, but their design normally dictates that such a lift is only taken from points on the gantry. This, in turn, increases the height requirement above that which is the case with a single lift on a selfsupporting tower which can normally be taken a little above the center of gravity. As an example, in the case of the guyed V-tower for a 500 kV quad bundle line, the maximum, i.e. 12 m extension tower weighing some 7.5 tonnes, was successfully erected under test conditions by the old faithful falling derrick method, but, as expected, the operation proved to be unacceptable in terms of cost and potential progress rate. In this case, with the prospect of having to erect some 4000 V-towers, it was decided that, as no existing cranage was available to meet these requirements, a purpose-built machine would be produced. From the outset it was realised that the design problems would be centred on the aspects of rapid travel between tower sites and subsequent deployment of lifting
182

equipment on arrival, rather than those relating to the actual lift; it having been established that from the time the hook was over the tower to that when the lifting slings could be removed was only of the order of some 20 min. The design finally evolved took the form of what was, in effect, a trailer-mounted derrick capable of travelling between sites with a boom of sufficient length to erect 80% of the tower types involved and also with the facility to increase the boom length to cover the remainder once the unit had arrived at the particular site. The derrick, which was a standard lightweight crane boom, was mounted on a levellable base carried on a crawler-tracked trailer closely coupled to and drawn by a D6 Class crawler tractor (Fig. 28). In service, two of these units were employed, each

Fig. 28

Crawler mounted mobile derrick erecting 500 kV V-structure

producing a cycle time totalling some 45 min when working in areas of unobstracted route. A similar principle may be applied to the guyed portal structure, although those so far encountered have been on lower voltage lines and of such a design as to lend themselves to falling derrick erection. However, with the advent of the highly mobile telescopic cranes, the falling derrick system may finally be laid to rest for other than the most inaccessible tower positions. 7.5.2 Crane erection: While for many years the line construction engineer has frequently wished to use cranes as a general method of erection, their long rigging and derigging times combined with either poor or slow crosscountry travel has, in all but exceptional cases, ruled them out. Now, however, this wish is becoming a reality with the arrival on the construction scene of the rough-terrain telescopic hydraulic crane. Machines of this type and of everincreasing height and lift capacity are becoming more generally available and, with the exception of soft ground conditions, a wide range of tower building applications are now possible. However, in planning their use, careful study is required to ensure the most economic results, and it has been found that, in certain circumstances, for instance 400 kV double-circuit towers, a combination of a smaller machine erecting the lower portions of the tower together with a large crane topping out possibly in a single lift, has proved successful. An example of the larger of this type of
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

crane currently in use in transmission-line construction is the Grove RT 980 with a basic rated capacity of 73 tonnes, but with a lift capability of 4.8 tonnes to a hook height of 54 m at 11.6 m radius. Despite their high capital cost when compared with that of traditional erecting equipment, their potential in the form of high productivity has been shown to be such that machines of this type are now becoming a regular feature of line construction, be it transmission or distribution, in those regions where the terrain and access conditions are suitable. 7.5.3 Erection by helicopter: Self-supporting and guyed structures of various configurations have been successfully erected by helicopter; in the former case, both in whole and in part, the majority of such work having been carried out in the USA and Canada. In considering the possible use of this method the following are some of the principal factors requiring study: (a) That it is the most economic method available to the contractor under the particular circumstances involved. (b) Ready local availability of suitable helicopter type as long-distance positioning will involve increased nonproductive costs. (c) If long-distance positioning is unavoidable, minimisation of the number of visits to the project by preassembly of the maximum number of towers. (d) Tower material purchase and delivery scheduled to render (b) a viable proposition. (e) A plentiful number of large, reasonably fiat areas, along a vehicular route as close as possible to the line for the purpose of tower assembly. If the number of such sites is limited it may be necessary to consider aerial haulage of bulk steelwork to inaccessible assembly sites close to the line route using a helicopter type possibly more costeffective than that which is needed to lift the tower as a whole. (/) Possibility of increased costs due to difficulties encountered in the assembly of towers, either in the prone position or in box sections. (g) On the assumption that the line route is difficult and hence tower to tower access is poor, additional cost may result from the need to mobilise, for a short period, a relatively large number of men to form the erecting crews, as the potential erection rate may well call for 4 or 5 ground crews if the flight distances are short. (h) That, particularly in the case of self-supporting towers of such a weight as to necessitate erection in sections, steps are taken at the design stage to ensure that main joints are such as to facilitate rapid and safe landing, either with or without special temporary attachments. It is essential that, at the earliest possible stage, there is cooperation between the design and erection functions to ensure that these details are satisfactorily resolved. As an example of the results which have been achieved in helicopter erection witnessed by the author, Fig. 29 shows a Bell 204B lifting a complete aluminium tower together with insulators and guys on a 138 kV line project in Canada. A high degree of careful forward planning, combined with specially designed lifting aids, together with experienced supervision, ground and air crews, produced the result that, in two consecutive days involving a total of some 15 h and 25 min of flying, some 115 structures were erected. It is unfortunate that the successful planning and execution of major works of this type requires a degree of specialised knowledge and experience which, owing to the limited use of such methods, is held by only a few individuals and organisations.
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

t-

Fig. 29

Bell 204B erecting 138 kV aluminium Y-structure

7.6.1 Wiring installation: In the past 15 years, conductor erection by the tension stringing method has become standard practice worldwide, and its use has extended down the voltage range to include 33 kV lines and, in some instances, lower. As is usually the case, as new problems occur a number of independently derived solutions arise, but with the passage of time a single solution comes to the fore and is adopted by the majority. Conductor erection, as practised by the major international contractors, has been no exception to this rule. An example of this process was the development of the pulling winch, vital to all systems, in which one school of thought chose the path of the drum type with the pulling length of necessity being limited by the amount of rope that could be carried on an acceptable sized machine, whereas the other chose to adopt the capstan type having an unlimited length facility, the rope, in disconnectable lengths, being reeled on to storage drums. The drum-type winch with its multibarrels, usually three in number, has gained limited acceptance, other than in North America, owing to its considerable weight (upwards of 25 tonnes), size and inflexibility with regard to pulling length. On the other hand, the twin bullwheel winch, with its much lower weight (10 tonnes for a 13-tonne capacity machine), combined with its compactness and pull length flexibility, is now used almost exlusively by contractors with worldwide operations. The design of the necessary conductor tensioners now varies little in basic concept in that each conductor passes several times around a pair of multigrooved large-diameter bull wheels which, in turn, drive oil pumps; the resulting energy being dissipated as heat by means of relief valves and oil coolers. There exist some detail variations but, in essence, the majority of the machines are similar.
183

A significant proportion of the capital investment involved in setting up tension stringing operations covers the large number of pulley blocks or travellers and twist resistant ropes through, and by which, the conductors are pulled. As opposed to machines which have a long life in terms of years due to relatively low running rates per year, the pulley block and rope life is more limited due to a combination of wear and casual damage, frequently more in transit than in service, and it is thus in the direction of these items to which contractors must look for new and more cost-effective products. Towards this end many contractors, when erecting lines of heavy bundle-conductor configuration, use only sufficient heavy pulling rope to erect one phase at a time, lighter pilot ropes having previously been installed in all phases. Certain detailed improvements in techniques have been made in recent years, such as the introduction of temporary joint protecting sleeves permitting the making of midspan joints in front of the tensioner, for subsequent pulling through the pulley blocks to their final position in the section without damage. The adoption of this method can save considerable time and also avoid costly delay to the main stringing activity, particularly in difficult terrain where the final joint positions may have poor accessibility. Improvements to sagging techniques have been evolved, mainly concerning the temporary anchoring of conductors after stringing at near sag tension and the making off of all terminal joints at tension tower positions at crossarm level. One of the more significant method changes has been imposed on the contractor by the environment in which he has been required to work, this being the length of single pulls. For a number of years, particularly in the UK, it was customary to pull from tension tower to tension tower, these being deliberately located adjacent to good accesses. However, in the developing world and more particularly in mountainous regions, access for conductor drums and heavy machines must be made, and thus it is necessary to balance this cost against the increased investment in additional stringing equipment, particularly in terms of pulley blocks and ropes. The results of such comparisons frequently give rise to pulling lengths of 6 km and, in some areas of particular difficulty, distances of 9 or 10 km may be necessary. 7.6.2 River crossings: Owing to a combination of conductor sizes and prescribed clearances, the stringing tensions required are frequently such as to be well in excess of the capabilities of machinery designed to meet the needs of the current range of EHV transmission lines. In the past, it was not uncommon for complete systems to be designed on a one-off basis involving, in some instances, special heavy duty winches and tensioners and, in others, such arrangements as catenary support systems; the resulting equipment frequently being purchased by the client for subsequent maintenance purposes. However, more recently, clients have tended to require the prospective contractor to provide all the necessary installation equipment within the contract price, and hence competitive pressures have produced methods which can be executed with, as far as possible, equipment normally found in the major contractors inventory. One method, well known to many construction engineers as the 'endless loop', but long since discarded owing to poor cost-effectiveness in the context of normal line construction, has recently found favour with a number of contractors within the context of river crossings. Although there are a number of possible detailed variations, in essence this method involves the following basic
184

stages: (a) The installation of anchorages on each bank outside the crossing section, each anchorage having attached to it a turn-sheave and tension adjustment tackle. (b) With the river closed to navigation, the establishment of a closed loop of light pilot rope through a pair of horizontally opposed phases and incorporating the turn-sheaves. (c) With the river reopened, the replacement of the light pilot rope with the heavier haulage rope, using only standard tension stringing machines. (d) At each end of the crossing section, but in opposite circuits, the insertion of a phase conductor into the loop. (e) The circulating of the thus completed loop until the conductors reach their required location within the section. This and the previous operation will require to be repeated for as many lengths of conductor as make up each phase. (/) The conductors are sagged and landed in the normal manner. This procedure is repeated at each crossarm level and also for the earthwires as appropriate. A recent and noteworthy example of the use of this system to good effect was in the crossing of the River Jamuna in Bangladesh. In this instance, the crossing section involved was particularly long, the distance between the anchor towers being 13 635 m and involving eleven intermediate suspension towers. It is believed that this is the first use of this method over such a succession of long spans and the whole operation, involving both preparation and the stringing of two earth and six-phase conductors, took only some six weeks.

7.7 Conductors con tain ing fibre - optic elemen ts


As the introduction of conductors of this type has only recently occurred, experience in their handling is of necessity limited and hence that relating only to the BICC product is included. Within these limitations it has been found that, with one major exception, i.e. jointing, the techniques involved in installation are very similar to those for conventional conductors, although some precautions as mentioned later are necessary with single-layer designs. Considerable efforts are being made in both research and development directed towards the production of simple and effective joints for use both at support structures and between conductor lengths in span. In arriving at suitable designs it is not only necessary to maintain both the mechanical and electrical integrity of the conductor but also, and most importantly, to do so without prejudice to the optical properties of the core. Additionally, it is necessary to provide within the joint design a suitably protected and accessible cavity within which connections of optical cores may be effectively and speedily made. In this context, one of the authors recently took part in the installation of the first full-tension midspan joints to be made in an operational overhead-line conductor containing fibre-optic cores. In this instance, the conductor took the form of a small-diameter high-strength shield wire incorporating a four-core optic element running between two existing substations some 6 km apart. At that time, the maximum conductor production length was 2.2 km so that a minimum of two through joints were necessary, in addition to those at the terminations. In this instance, the system owners, Rochester Gas and Electric, opted to install the newly developed midspan joint, as opposed to the pole-mounted type which hitherto was the only form available.
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

In the field the jointing operations, both optical and mechanical, were carried out by the owner's staff with greater facility and speed than had been expected, and the subsequent optical test results were well within the forecast limits (Fig. 30).

In general, it is considered that, with increasing experience, such apprehension that may have existed as to the damage-resistant capability of such an apparently delicate member in the rigorous environment of the field construction activity will rapidly evaporate, and its installation will become just another function to be performed by the competent contractor.
8

Conclusions

Fig. 30 Compression of tension component of shield wire containing four fibre-optic elements

The erection of conductors containing the fibre cores within an aluminium-alloy tube and with more than one layer of outer strands has produced no problems under full-tension stringing conditions to the extent that no degradation of the optical performance has been noted. However, those designs produced as high-strength shield wire replacements have, to date, had only one layer of outer wires and have demonstrated a not entirely unexpected tendency to twist with increasing longitudinal tension, the resulting level of longitudinal extension being undesirable in relation to the optical core. It was therefore necessary to evolve a simple installation procedure to eliminate rather than minimise the twisting effect, this being achieved by the use of a combination of a form of headboard not unlike those commonly used during bundle-conductor stringing and also come-along clamps with inbuilt counter weights for catch-off and termination purposes. Although, since the time of the original trials on the early production runs, the rotational properties of the single-layer designs have been much improved by advances in production techniques, the use of the twist prevention procedure is advised when tension stringing is to be employed. When such single-layer conductors are installed by the slack stringing method, that is with the maximum tension not greater than 2.5% of UTS, the use of the special headboard is not required, but the antitwist clamps should be used during catch-off and termination activities. Previously, certain criteria had been established relating to such constraints as relationship between conductor diameter on such items as running-out blocks, tensioner bullwheels, storage drums and clamping-in shoe profiles. However, it is expected that these will become less onerous as a result of increased field experience and ongoing test programmes. As far as special skills are concerned, unlike conventional overhead-line installation where a man of a single skill range can accomplish all the functions including electrical tests, in the case of the fibre-optic cable the processes are much akin to the conventional telecommunications cable, where the skill and aptitude range is presently considered somewhat wide to cover both pulling in and jointing; thus separate personnel are required. However, this situation may well change rapidly as more experience is gained.
IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

The careful reader will have noticed that the title of the paper is fully justified. The few electrical aspects which have been described are far from exhaustive. They have been selected to show trends of thoughts and recent advances in some technologies. The concepts of load and strength assessments by statistical means to define the reliability of a system are not new. They have been applied for many years in other civil engineering activities, and form the basis of several building codes. What is surprising is that they have not been implemented, at least consciously, in the design and construction of overhead lines, or is it that the awareness of these approaches has grown as a result of recent failures? When the size of a population (here used in the statistical concept of kilometre years of lines) grows, the probabilities of a meteorological event 'finding' a weak spot increases; it is, therefore, necessary to resist the temptation of increasing specific loads for a subjective increase in reliability! Overhead-line engineers should become conscious of the difference between macrometeorology and micrometeorology. As stated many times by a colleague of one of the authors, a line 1000 km long is a good meteorological station! Although many engineers might think that all has been said on the subject, there are still developments in progress and more are possible, even with such a component as the overhead-line conductor. This is a normal process in engineering attitude: when the most pressing problems have been solved to the satisfaction of most people, the time is ripe for investigating the smaller problems. Is it right to assume that a conductor specification, perfectly adequate for medium-voltage lines erected by pulling on the ground, will be satisfactory for erection under tension for EHV lines? Why should conductor properties have discrete values independent of manufacturing processes or of intercrystalline configurations? A recent exercise has shown that the so-called minimum guaranteed strength was approximately 4 to 5 standard deviations from the mean value. Is this an efficient way for using conductors? There is no doubt that composite insulators will be more widely used in a few years. There are still many problems to solve, especially regarding acceptance tests. The problem of brittle failures is still the most difficult to analyse by a short-time test to provide confidence for many years of utilisation. The technical press refers quite frequently to compact lines, in as much as phase spacings have been reduced, but there appears to be no general agreement as to the techniques to be used for compacting the lines. Here again, the solution seems to lie in probability analysis of wind speed distribution, probability analyses of overvoltages, leading to probabilities of flashovers of air gaps. All indications are that tremendous advances in materials for towers cannot be expected. This being so, work can proceed in solving problems of detail such as bearing stresses, end distances etc. What is however possible is the designer's ingenuity in investigating new shapes and improving the appearance of their outlines to make
185

them environmentally more acceptable. Computerisation of designs will help designers in investigating several solutions to arrive at the most economic one. Although statistical techniques may be used to define the design loads, proof of adequacy of design by testing will still be done on a deterministic basis. The variability of soil conditions along a transmissionline route and the various types of foundations utilised are contributing to the complexity of foundation design by differing techniques. Much frustration could be avoided if early agreement was reached between client and contractor as to the mathematical modelling of a particular foundation. Different techniques of calculations would provide widely differing answers. Hence the increased tendency to carry out soil reconnaissance and testing, the latter being only a random check at a particular location of the accuracy of the uplift resistance. There is still wide disagreement between a tower designer and a foundation designer as to what constitutes a damaged state. There is clearly a need to develop techniques for calculating foundation movements, as this could well become a design parameter. With the increased cost of labour, with the need to instal lines in inaccessible areas to satisfy environmental considerations, there has been a shift in opinion regarding the economics of line construction using helicopters. Only a few years ago, the opinion was held that helicopters should only be used in exceptional circumstances, preferably with aluminium-alloy towers (thus saving on weight). Nowadays, helicopter installation is part of the overall economic analysis of construction. In some cases, they have been shown to be competitive, and, in fact, in a particular case, the construction proved possible only if this modern technique was employed. In this paper, the logistics connected with such operations have been highlighted.

Acknowledgments

The authors wish to acknowledge the help of their colleagues, particularly the Chief Engineer of Balfour Beatty Power Construction Limited in the preparation of this work, and they thank the General Manager and the Chief Engineer of the Power Transmission Division, Balfour Beatty Power Construction Limited for permission to publish this paper.

10

References

1 COX, E.H.: 'Overhead-line practice', Proc. IEE, 1975, 122, (10R), pp. 1009-1017 2 'Review of technical standards for overhead lines following demage in December 1981 and January 1982'. Department of Energy, Electricity Supply Industry, September 1982 3 WOOD, A.B., and KENNEDY, M.W.: 'The role of the consulting engineer in the design of high-voltage power transmission projects overseas', IEE Proc. A, 1983, 130, (4), pp. 202-206 4 'Statutory instruments'. 1355: The Electricity (Overhead Lines) Regulations (HMSO, 1970) 5 HAIGH, F.R.: 'New overhead line regulations come into operation', Electrical Times, 1971, 160, pp. 35-36 6 IEC Technical Committee 11: 'Overhead line tower loadings' (to be published) 7 'Part 2: Wind loads'. CP3 (BSI, 1972), Chap. 5 8 'Wind, snow and temperature effects on structures based on probability' (Abacus Press, 1975) 9 ARMITT, J, CO JAN, M., MANUZIO, C, and NICOLINI, P.: 'Calculation of wind loadings on components of overhead lines', Proc. IEE, 1975, 122, (11), pp. 1247-1252 10 WOOD, A.B.: 'Transmission line designthe ultimate load concept', CIGRE, Paper 22-01, 1982
186

11 DI GOIA, A.M., POHLMAN, J.G, and RALSTON, P.: 'A new method for determining the structural reliability of transmission lines', CIGRE, Paper 22-08, 1982 12 GHANNOUM, E.: 'A rational approach to structural design of transmission lines', IEEE, PES Winter Meeting, Atlanta, Georgia, lst-6th February, 1981 13 ASH, DO., DEY, P., GAYLARD, B., and GIBBON, R.R.: 'Conductor systems for overhead lines: some considerations in their selection', Proc. IEE, 1979, 126, (4), pp. 333-341 14 'Interference produced by corona effect of electric systems (description of phenomena, practical guide for calculation)' (CIGRE Publication, Paris, 1974) 15 'Survey of extra high voltage transmission line radio noise', Electra, January 1972,(29) 16 'Specification for CISPR radio interference apparatus for the frequency range 0.15 MHz to 30 MHz'. CISPR Publication 1 (IEC) 17 DEY, P., GAYLARD, B., HOLDEN, G., SMITH, P., CARTER, C.N., MADDOCK, B.J., and KENT, A.H.: 'Official communication using overhead power transmission lines'. CIGRE, Paper 35-01, 1982 18 BOYSE, CO., and SIMPSON, N.G.: 'The problem of conductor sagging on overhead transmission lines', J. IEE, 1944, 91, Pt. II, (21), pp. 219-231 (Discussion, pp. 231-238) 19 REIGER, H.: 'Der Freileitungsbau' (Springer Verlag, Germany, 1960) 20 WINKELMAN, P.F.: 'Sag-tension computations and field measurements of Bonneville Power Administration', Trans. Amer. Inst. Elect. Engrs, I960, 78, Pt. Ill, pp. 1532-1548 21 BRADBURY, J., KUSKA, G.F., and TARR, D.J.: 'Sag and tension calculations for mountainous terrain', IEE Proc. C, Gen., Trans. & Distrib., 1982,129, (5), pp. 213-220 22 DEY, P., GAYLARD, B, MOLT, C.W., and NICHOLSON, J.A.: 'Influence of conductor designs and operating temperature on the economics of overhead lines', Proc. IEE, 1971,118, (3/4), pp. 573-590 23 BRADBURY, J., DEY, P., ORAWSKI, G., and PICKUP, K.H.: 'Long-term creep assessment for overhead-line conductors', ibid., 1975,122, (10), pp. 1146-1151 24 'Permanent elongation of conductorPredictor equations and evaluation methods', Electra, 1981, (75), pp. 63-98 25 ERA Report OT26 26 EDWARDS, A.T., and BOYD, J.M.: 'Influence of terrain and other factors on mechanical oscillation of overhead conductors'. Ontario Hydro Research Quarterly, 2nd quarter, 1968 27 'Standardisation of conductor vibration measurements'. IEEE Committee Report, IEEE Trans., 1966, PAS-85, pp. 10-22 28 'Aeolian vibration on overhead lines'. Working Group 01 of CIGRE SC22, CIGRE, Paper 22-11, 1970 29 HEARNSHAW, D.: 'Spacer damper performanceA function of in-span positioning'. IEEE Paper T74 061-8, Winter Meeting, 1974 30 DAVIS, D.A., RICHARDS, D.J.W., and SCRIVEN, R.A.: 'Investigation of conductor oscillation on the 275 kV crossing of the Rivers Severn and Wye', Proc. IEE, 1963,110, (1), pp. 205-219 31 ROWBOTTOM, M.D., and ALLNUTT, J.G.: 'Mechanical dampers for the control of full span galloping oscillations', IEE Proc. C, Gen., Trans. & Distrib., 1982,129, (3), pp. 123-135 32 HAVARD, D.G., PAULSON, A.S., and POHLMAN, J.C.: The economic benefit of controls for conductor galloping'. CIGRE, Paper 22-02, 1982 33 ROWBOTTOM, M.D.: 'Method of calculating the vulnerability of an overhead transmission line to faults caused by galloping', ibid., 1981, 128, (6), pp. 334-341 34 DAVISON, A.E.: 'Ice coated electrical conductors'. CIGRE, Paper 229,1939 35 BS 137: 'Specification for insulators of ceramic material of glass for overhead lines with a nominal voltage greater than 1000 V. Pt. 1: tests, 1982, Pt. 2: requirements, 1973 36 IEC 305: 1978. 'Characteristics of string insulator units of the cap and pin type' 37 IEC 383: Revised Dec. 1983. 'Tests on insulators of ceramic material or glass for overhead lines with a nominal voltage greater than 1000 V 38 IEC 71: 1976. 'Insulation co-ordination'. Part 1: Terms, definitions, principles and rules, Part 2: Application guide 39 BS 5622: 1979. 'Insulation co-ordination'. Part 1: Terms, definitions, principles and rules, Part 2: Application guide 40 IEC 507: 1975. 'Artificial pollution tests on high voltage insulators to be used on AC systems' 41 IEC TC 36 (Secretarist) 53: 1982. 'Work of T.C. 36: Draft of a guide for the selection of insulators under pollution conditions' 42 THOMAS, A.M., and OAKESHOTT, D.F.: 'Choice of insulation and surge protection of overhead transmission lines of 33 kV and above', Proc. IEE, 1957,104A, pp. 229-239, d. 240-248 43 PARIS, L., and CORTINA, R.: 'Switching and lightning impulse discharge characteristics of large air gaps and long insulator strings', IEEE Trans., 1968, PAS-87, (4) IEE PROCEEDINGS, Vol. 131, Pt. C, No. 5, SEPTEMBER 1984

44 'Lignes aeriennes a haute et tres haute tensionRegies d'isolation'. (Electricite de France, Direction Etudes et Recherches, 1976) 45 MAGIMEY, F.: 'The lightning performance of a 275 kV transmission line in Malaya'. Far East Conference on Electric Power Supply Industry, CEPSI, Paper B.02, December 1978, Hong Kong 46 ACE Report 55: 1977. Classification ' C . British Electricity Board report on lightning protection for networks up to 132 kV 47 CLAYTON, J.M., and YOUNG, F.S.: 'Estimating lightning performance of transmission lines', IEEE Trans., 1964, PAS-83, pp. 1102-1110 48 ANDERSON, J.G.: 'Lightning performance of EHV-UHV lines'. Transmission line reference book on 345 kV and above, Electric Power Research Institute, Palo Alto, California, USA, 1975 49 WHITEHEAD, E.R.: 'Mechanism of lightning flashover on transmission lines'. Final report of Edison Electric Institute Lightning Flashover Research Project RP-50, Paper 72-900, 1972 50 ANDERSON, J.F.: 'Monte-Carlo computer calculations of transmission line lightning performance', AIEE Trans., 1961, PAS-80, pp. 414-419 51 'Design of foundations for steel towers for overhead transmission lines at 132 kV and higher voltages'. ESI Standard 43-4, 1972 52 'Regulations for the construction of overhead power lines above 1 kV\ VDE 0210/5.69, 1969 53 'Construction of overhead transmission lines on steel towers for 132 kV and higher voltages'. ESI Standard 43-6, 1972

54 ASHBEE, R.A., GILLSON, I.P., and CLIFFE, H.L.: 'Overhead line tower foundationsSome recent work on augered footings and the use of conventional piles in raked formation', IEE Conf. Publ. 44, 1968, pp. 53-60 55 LITTLEJOHN, G.A.: 'Design estimation of the ultimate load capacity of ground anchors', Ground Engineering, Nov. 1980, pp. 25-39 56 ADAMS, J.I.: 'Grouted anchor transmission tower footings'. Ontario Hydro Research Quarterly, 3rd quarter, 1969, pp. 1-7 57 DEMBICKI, E., BOLT, A., and ODROBINSKI, W.: 'Stabilite des massifs de foundation soumis a un moment de renversament'. Annales de l'Institut Technique du Batiment et des Travaux Publics, Paper 360, April 1978, pp. 3-27 58 'Laterally loaded drilled pier research'. GAI Consultants, EPRI EL 2197, Vols. 1 and 2, 1982 59 BIAREZ, J., and BARRAUD, Y.: 'The use of soil mechanics methods for adapting tower foundations to soil conditions'. CIGRE, Report 22-06, 1968 60 CAUZILLO, D.A.: 'Metodo di calcolo del carico limite per fondazioni sollecitate a trazione', L'Energia Elettrica, 1973, 50, (6), pp. 3-12 61 VANNER, M.J.: 'Strength tests on overhead line tower foundations. The effect of variation of depth of burial'. ERA Report 5202, 1967 62 VANNER, M.J.: 'Foundation uplift resistance: the effect of foundation type and of seasonal changes in ground conditions'. IEE Proc. C, Gen., Trans. & Distrib., 1982,129, (6), pp. 295-305 63 British Standard Code of Practice CP 2004: 1972, 'Foundations'

IEE PROCEEDINGS,

Vol. 131, Pt. C, No. 5, SEPTEMBER

1984

187

Vous aimerez peut-être aussi