You are on page 1of 19

Wind Effects on Tall Steel Towers

S. Selvi Rajan, P.K. Umesha and G. Ramesh Babu

1.0 INTRODUCTION

Tall steel towers are widely used as radio and communication towers, power transmission
line towers, flood lighting towers in sports stadia, etc. Being tall and flexible, these
structures are highly susceptible to dynamic wind actions. Generally evaluation of wind
loads on tall lattice towers is complex, as there are uncertainties in defining some of the
parameters related to wind characteristics and dynamic properties of the structure.
Computation of the along-wind dynamic response of wind sensitive structures subjected
to random wind loads involves complicated stochastic analysis. Hence for the
convenience of designers, an equivalent static load approach has been used in practice in
which the mean wind loads are multiplied by a ‘Gust Response Factor’ (GRF), G. The
GRF is the ratio of the peak response to mean response of the structure due to the wind
induced random loads on the structure. Several authors [1, 2, 3] have derived and used
semi-analytical statistical approaches with a simple model of the relation between the
upwind turbulent velocity fluctuations and the fluctuating forces on the structures. These
semi-analytically derived expressions have been used in most of the International
Standards [4, 5, 6, 7]. In these approaches, the GRF for the horizontal sway deflection of
the structure is computed by considering only the first mode of vibration of the structure.
It is also assumed that the first mode shape of the structure varies linearly with height.
The present Indian Standard on Wind Loads IS: 875 (Part 3) – 1987 [4], was developed
for tall buildings with uniform mass distribution and with a linear fundamental mode
shape. The method also yields same values of gust response factor for various load
effects such as bending moment and deflections. In case of lattice towers, these
assumptions are less appropriate and hence suitable adjustments have been proposed in
the above method and a semi-analytical procedure is suggested in recent years, for
evaluating wind induced oscillations on lattice towers.

Wind is essentially a random phenomenon both in time and space, and hence is dynamic
in nature. When the natural vibration frequencies of structures are low enough to be
excited by the turbulence in the natural wind, the structures are considered to be
dynamically wind sensitive. The dynamic response analysis of these structures is
generally recommended when the fundamental frequency of the structure is less than
1 Hz.

Basically there are three sources of aerodynamic excitation causing the dynamic
response:
i) Forces induced by turbulent fluctuations in the on coming flow: These forces
cause both background and resonant responses in the along-wind and cross-wind
directions.
ii) Forces induced by vortices shed in the wake of the structure: These forces affect
primarily the resonant responses and occur primarily in the cross-wind direction.
iii) Forces induced by motion of the structure: These forces can add to or subtract
from the available structural damping. Negative aerodynamic damping is
primarily associated with the cross-wind motion and can cause large amplitude
oscillations and in the extreme case lead to aerodynamic instability.

This lecture note briefly describes the basic features of nature of wind and GRF for the
wind sensitive structures. Further, a typical case study to experimentally determine GRF
using wind tunnel studies on group of lattice towers against dynamic wind loads is
described in detail.

2.0 WIND CHARACTERISTICS

Before examining the dynamic effects of wind on structures, it is useful to consider/study


the structure of the wind itself. The natural wind consists of a steady mean flow averaged
over a suitable period upon which are superimposed fluctuations which are normally
called as gusts or turbulence, as given below.

U(t) = U  u(t ) (1)

2.1 Hourly Mean Wind Speed

For choosing the suitable averaging period, consider the results of the analysis by Van
der Hoven of a number of wind records of various lengths as shown in Fig. 1 [8]. The
figure shows the distribution of the turbulence energy over a range of frequencies. This
spectrum consists of two peaks, (i) ‘macro meteorological peak’ at a frequency of 0.01
cycles/hour, corresponding to 4-day transit period and (ii) ‘micro meteorological peak’
between the time periods of 10 minutes and 3 seconds. In terms of the physical
mechanism causing high winds, the first peak can be identified with the passage of a fully
developed weather system (complete isobar system) over the observation point, while the
second peak arises from the fluctuations generated by the stirring of the mean flow by
surface friction (i.e., turbulence in the boundary layer). In between these two peaks and
over a time scale range of 15 minutes and 2 hours, there exists a ‘spectral gap’ in which
the turbulence/wind fluctuations are observed to be insignificant. Hence for an averaging
period of 1 hour, a suitable mean wind speed called hourly mean wind speed can be
obtained.

2.2 Turbulence Intensity

The ratio of r.m.s gust speed to the mean wind speed is called turbulence intensity and is
given as
u
Iu = (2)
U
T
1
u   u 2 (t ) dt
2
where (3)
T 0
Fig. 1 Spectrum of horizontal wind speed

2.3 Variation of Mean Wind Speed and Turbulence Intensity with Height

At large heights above the earth’s surface, where the effects of surface friction can be
ignored, the resultant of (i) forces produced by the atmospheric pressure differences
which act on a given mass of air and (ii) forces due to the curvature and rotation of the
earth, produce a steady motion of wind known as gradient wind, VG. The height at which
the wind velocity is equal to the gradient velocity is known as the gradient height, ZG.
Below the gradient height, the wind is retarded by surface friction from a value of V G at
the gradient height to a value of zero at the earth’s surface. The gradient height normally
lies between 250 m and 600 m above the ground level depending upon the surface
roughness. The variation of mean wind speed with height can be represented using
power-law model or log-law model as given below


(i) Power-law model: U z   z  (4)
U 10  10 

 z 
(ii) Log-law model:U z  2.5 u* ln   (5)
 zo 
where u* = shear friction velocity
zo = terrain roughness length

The results of the natural wind measurements show that the r.m.s gust speed decreases
very slowly with height and as per the physical model of high winds, the total turbulence
or gust speed must tend to zero at the gradient height. For structural engineering
purposes, the r.m.s. gust speed is assumed to be invariant and equal to the value measured
at 10 m above ground level [9]. Hence the variation of turbulence intensity, Iu with height
is inverse to the variation of mean wind speed with height. Fig. 2 shows the gradient
height, variation of mean wind speed with height and also the turbulence intensity values
at 10 m level for different terrain categories [10].

Fig. 2 Variation of mean wind speed with terrain roughness and height

2.4 Turbulence Length Scales

Turbulence length scale is a measure of the average size of the gust in a particular
direction (x, y and z). The turbulence length scales increase with height at the same rate
as the mean wind speed [9]. The turbulence length scale in the along-wind direction can
be calculated from the measurements made at multiple points in the along-wind direction,
x, as given below [11]:
  u ( x, t )u ( x  x, t ) 
Lx = 0   ( x) ( x  x)  dx (6)
 u u 

T
1
T 0
where u ( x, t )u ( x  x, t )  u ( x, t ) u ( x  x, t ) dt (7)

= cross correlation function


2.5 Turbulence Spectrum

The turbulence spectrum describes the distribution of the turbulence with frequency. A
low-frequency fluctuation implies that a large eddy was convected past, and a high-
frequency fluctuation implies a small eddy. A spectral density function, S(n), is defined
as the differential of the variance by frequency, n, as given below
d 2
S(n) = (8)
dn

The typical shape of atmospheric turbulence spectra is shown in Fig. 3 [8]. The area
under the curve represents the variance of the turbulence component, and can be divided
into three ranges by frequency. In the ‘production range’ at the low-frequency end,
turbulence is generated as large eddies from instabilities of the mean flow. The large
eddies have the highest velocities. In the ‘inertial range’ over the middle range of
frequency, these large eddies break up, transferring their momentum to smaller ones, then
yet smaller ones, etc, forming an ‘energy cascade’. Finally in the ‘dissipation range’ at
the high-frequency end, the eddies become so small that viscosity becomes significant
and their energy is dissipated as heat. When the natural frequencies of structures fall in
the ‘inertial range’, the structures are considered dynamically wind sensitive.

Fig. 3 Typical wind turbulence spectrum

2.6 Turbulence Cross Spectrum (Coherence Function)

The coherence function represents the cross correlation of turbulence between points
divided in space and it can be obtained as [3]:
S uu ' ( x, z; x' , z ' ; n)
Coh(x,z;x’,z’;n) = (9)
S uu ( x, z; n) S u 'u ' ( x' , z ' ; n)

where S uu' ( x, z; x' , z' ; n) = cross spectral density function of u(x,z,t) and u(x’,z’,t).
The coherence function tends to diminish as the distance between the points of examine
(x,z and x’,z’) and the wave number (n/U) becomes greater. Vickery [3] proposed the
following relationship:

 2n C 2 ( x  x ' ) 2  C 2 ( z  z ' ) 2 
Coh(x,z;x’,z’;n) = exp  
x z
(10)


U ( z )  U ( z ' ) 

where Cx and Cz are the exponential decay coefficients in x-direction and z-direction

3.0 GUST RESPONSE FACTOR

The peak response of the structure along the height can be given as:

xˆ( z)  x ( z)  g x ( z) (11)

The area under the response spectrum (Fig. 4) comprises two parts, (i) the quasi-static or
steady state or broad banded response spectrum area and (ii) the resonant or narrow
banded response spectrum. Accordingly, the r.m.s. dynamic response has been divided
into background response (in the low-frequency region) and resonant response (around
the natural frequency) and the corresponding peak response is given as:

xˆ ( z )  x ( z )  [ g B xB ( z )] 2  [ g R xR ( z )] 2 (12)

Gust response factor for the structure can be obtained using the mean, background and
resonant responses as given below:

xˆ ( z ) {g B xB ( z )}2  {g R xR ( z )}2
Gx(z) = x ( z) = 1 (13)
x ( z)

Fig. 4 Elements of the statistical approach to gust loads


3.1 Indian Standard Code of Practice for Design Loads for Buildings and
Structures, Part 3, Wind Loads IS: 875 (Part 3) [4]

The IS Code on wind loads, Clause 7.1 [4], states that for buildings and closed structures
having aspect ratio (height / least lateral dimension) greater than 5 or whose natural
frequency in the first mode is less than 1 Hz, the structures have to be designed for
dynamic effects also. IS code [4] gives the Gust Effectiveness Factor for estimating the
peak wind load and response for a dynamically sensitive structure by considering the
natural frequency of 1st sway mode alone. The gust factor is given by:

 SE 
G = 1  g f r B(1   ) 2    (14)
  

The parameters gf r, B, S and E can be obtained from the graphs given in the IS Code [4].
 gfr B 
The parameter  can be calculated as   and used restrictively as per IS Code [4].
 4 
 
For a zero-mean random Gaussian process with a period of T, the peak value is given by
[8]:

<xmax> = x g (15)

where x = r.m.s of the fluctuating component

0.5772
g = statistical peak factor = 2 ln(T )  (16)
2 ln(T )

n
2
S xx ( n) dn
 = zero-crossing frequency = 0

(17)
S
0
xx ( n) dn

4.0 WIND TUNNEL TESTING ON LIGHTNING PROTECTION TOWERS –


TYPICAL CASE STUDY

The lightning protection tower is of lattice type, with a height of 120 m in full-scale. Four
such towers are located at the four corners of a rectangle with sides 90 m x 108.25 m.
The size of the tower at the base is 16 m x 16 m and that at the top is 3 m x 3m in full-
scale. Following are the scope of the study:

(i) To conduct wind tunnel investigations of the model of a system comprising of


four numbers of self-supported lattice steel towers of height 120m, with 10m
high lightning masts and interconnected steel ropes.
(ii) To examine the possibility of galloping oscillations in wire ropes and
(iii) To study the effect of damping on the system.

4.1 Simulation of Wind Characteristics

Considering the dimensions of the test section of boundary layer wind tunnel facility at
CSIR-SERC, Chennai, a model scale of 1:100 was selected. The wind characteristics
corresponding to an open terrain category have been simulated, using a trip board and
floor roughness blocks, as vortex generators. The power law exponent obtained based on
measurements is equal to 0.14, corresponding to open terrain category.

4.2 Preparation of Model Tower

Each of the four lattice tower models was in-house fabricated to a scale of 1:100. The
geometric shape of the model was maintained same as that of the full-scale tower at all
levels. Using mild steel material, sharp edged angle members of sizes with 2mm x 2mm x
0.5 mm of thick were used for the fabrication. The individual members (4mm x 4mm x
0.5 mm) were connected using suitable bracing technique. The solidity ratio of the full-
scale lattice tower varied from 0.14 to 0.27, with most of the panels, being around 0.15.
In the model also these values were modelled as closely as possible and an average value
of  = 0.15 is assumed for calculation purposes. The tower model was firmly connected
to a base plate, which in turn was rigidly connected to the turntable.

4.3 Instrumentation and Calibration

Each of the four towers was instrumented at the base with strain gauges fixed on to the
main leg members, to measure the base bending moments. Full-bridge circuitry was used
for the strain gauge instrumentation. The conventional dead weight loading system was
used along the diagonals and the strain gauges were calibrated. Cross talk between the
strain gauges was found to be negligible. With S1 and S2 as the sensitivity of the strain
gauges 1 and 2 respectively, for the zero angle of wind incidence, the along wind bending
moment was obtained as:

Mx = 0. 707 (S1 V1 + S2. V2) (18)

where V1 and V2 were the output voltages corresponding to strain gauges 1 and 2
respectively. The across wind bending moment was obtained as:

My = 0.707 (S1 V1 - S2 V2 ) (19)

The values of S1 and S2 were experimentally obtained in terms of kg-cm per one Volt of
output.
Further, light weight accelerometers were used at the tip of the tower, both in along and
across wind faces of the model tower, to measure the tip accelerometers. Using the
“Global Lab” application software, which is further custom-tailored at CSIR-SERC, the
voluminous data acquired were processed to obtain traces of tip deflections and spectra of
deflections, both in along wind and across wind directions. A schematic diagram of the
arrangement of tower models in the wind tunnel is shown in Fig. 5. For the tower ‘T1’,
the values of natural frequency in two orthogonal directions were obtained as 60 Hz and
62.5 Hz. The damping ratio, calculated based on tail end region of the free vibration
trace was found to be equal to 0.7% of critical value. Similar values of natural frequency
and damping ratio were obtained for the other towers.

Unlike tall buildings, for lattice towers, the fundamental mode shape significantly
deviates from a linear variation with the height. Good agreement was obtained between
the experimentally obtained first mode shape and the corresponding full-scale tower
mode shape. The dead weight system was also used to determine the influence line for
deflection of the model tower experimentally.

CaseI
108.25 cm
T4
T1

Case II
90
cm

T2
T3

Case III
Case IV

Fig. 5 Arrangement of Lightning Protection Tower Models in the Wind Tunnel


4.4 Wind Tunnel Testing on the Group of Towers

Wind tunnel investigation of all the four lightning protection towers, located at the four
corners of a rectangle was carried out. The arrangement of the towers is shown in Fig. 5.
The experimental set-up is shown in Fig. 6. It may be seen that the sides of the rectangle
measuring 90 cm x 108.25 cm in the model, formed by the tower system are coinciding
with the lines joining the diagonals of the individual towers. The performance of all the
four individual towers were studied for the following four cases of tower- system:

(i) Case – I : Incident wind flow parallel to one of the diagonals of the tower
system, passing through towers, T1 and T3.

(ii) Case –II : Incident wind flow normal to the side of the tower system
connecting the individual towers, T1 and T2.

(iii) Case –III : Incident wind flow normal to the side of the tower system,
connecting the individual towers, T2 and T3.

(iv) Case – IV : Incident wind flow parallel to the other diagonal of the tower
system, passing through towers, T2 and T4.

The tower-group system was investigated for a range of wind speeds varying from 7.0
m/s to 21.0 m/s. The test data were acquired using a sampling rate of 1000 samples/s and
a sampling period of 10 seconds. As stated earlier the acquired test data were
subsequently analysed using ‘Global Lab’ application software to determine statistical
values and to study spectral distribution of base bending moment and tip deflection
response values, including background (contribution due to broad band frequencies) and
resonant (contribution due to a narrow based frequencies centered around natural
frequency) responses. The along and across wind responses for each tower were
evaluated by suitably resolving the measured strain gauge values. The experimental data
were analysed for all four towers in all the four cases, and the results of fitted empirical
equations provided values for mean, background and resonant components of base BM,
as well as values of GRF for BM and deflection too.

For the given spacing of the four tower system, the results indicate no interference effects
and the towers tend to behave independently. The average mean drag coefficient for the
isolated tower was found to be equal to 3.1 and 3.47 respectively for the cases of wind
normal to the face, and for a diagonal wind. The corresponding values recommended as
per IS: 875 (Part 3) 1987 [4] are equal to 3.55 and 4.26 respectively. Similarly, the above
corresponding values as per the Australian Code of Practice, AS: 3995 – 1994 [12] are
equal to 3.15 and 3.55 respectively. The wind tunnel tests can thus be seen to give a
reasonable agreement with Codal values for mean drag coefficient. Further, it was
observed that due to the sharp edged angle members, the lattice tower was not subjected
to significant vortex shedding oscillations. The value of Strouhal Number was
experimentally determined.
Fig. 6 Models of Lightning Protection Towers inside the
Boundary Layer Wind Tunnel

5.0 Tower Testing & Research Station (TTRS)


5.1. Introduction
Transmission line towers are highly repetitive and, therefore, their designs are to be
commercially competitive, even though they are very complex. The main dimensions of a
tower depend on its location on the line, number, sizes and arrangements of conductors
and earthwires, and arrangement of insulator strings. The structural design of the tower is
mainly governed by the wind load, the other loads being those due to line deviation, self-
weight of tower /conductor, broken wire conditions, cascading, erection and maintenance.
These design loads are categorized under reliability, security and safety conditions.
Under these conditions, several load cases are worked out to design a tower. The capacity
of the tower to withstand the design loads depends on the strength of the tower members
in tension /compression, their arrangement, and the joints and their detailing. Design and
detailing of the tower is verified through prototype tower tests in many countries. The
reasons for this are:
• It is a mass produced structure.
• Reduction in the chances of line failures, interruptions to supply, and the
consequential costs of repairs and out of merit generation.
• An acceptable and reliable way of checking towers designed and supplied
competitively when it is in the interests of the contractor to produce a minimum
weight design.
• The design checking of individual members as well as the scrutiny of them and
checking of joints is an involved job due to the complex nature of the structure
and the loading.
• The economic use of steel. Experience has shown that a high proportion of
failures is due to detailing and that the overall cost of design modifications is
small compared with the improvement in tower performance.
• Development of design techniques and long term economies in the use of steel in
towers.
• Other aspects such as material quality, fabrication, assembly, and erection of the
structure need to be checked.
• Training of designers who learn from the behaviour of full scale towers under test.
5.2 Prototype tower tests
When advanced techniques using modern computers are available for structural analysis,
is full-scale test necessary? This is the question frequently raised by some structural
engineers. But, an engineer working in this field knows that the test is mainly to check
the detailing of the tower in addition to the adequacy of member sizes. Mathematical
modeling of bolted joints is cumbersome. The tower is modeled as a pin jointed space
frame for the analysis and for the member design, end restraint (either as pinned or as
partially restrained) and end eccentricities within reasonable limits are considered. In
most of the cases, this design procedure works. But in some cases where the member
forces are very high (though stresses are within limits) or deformations are high or when
the gravity lines of tower members do not meet at joints or when large number of
redundants are introduced, the resistance of the tower to carry the specified ultimate loads
may not be as expected. There can be some unexpected stresses due to lack-of-fit of some
members also. Prototype tests are conducted to study these aspects as well as to check the
quality of materials used.
The main issues in testing are the number of load cases, sequence of load cases, and the
waiting time at 100% loading. The general recommendation is to take up first those load
cases which have the least influence on the results of subsequent tests and the sequence
should help in simplifying the operations of testing. These two criteria may contradict
each other for a particular sequence. Generally accepted sequence is to start from ‘not-so-
critical’ load members and joints after each load case for any possible failures or
deformations which cannot be so easily noted from a distance (like cracks across bolt
holes, bolt shearing, etc.) as these can lead to unexpected failures in a subsequent test.
The waiting period at 100% or above.
General requirements of a testing station are
1. A perfectly horizontal test pad with base connecting devices, which will not allow
any lateral or vertical movement of tower, legs.
2. Loading system for smooth loading of the test tower without any jerk with
provision for simultaneous loading of all load points.
3. Devices for measurement, monitoring, control and recording of loads and
direction of application of loads.
4. Mechanism to prevent overloading and immediate unloading when required for
when power supply fails.
5. Equipment for stress measurement on tower members
6. Facility for calibration of load cells
7. Facility to reach any part of the tower for inspection as and when needed.
5.3 Loading and application
Towers are designed for specific load combinations and during tests these should all be
applied proportionally at the same time. Generally there are two kinds of loads imposed
on a tower.
• Fixed loads such as self weight of the tower, weight of conductors, hardware etc.
• Fluctuating loads such as conductor tension, wind loads and live loads
The effects of the two types of loads on the design of tower are not quite identical or all
the loads are not resisted by the different components of the tower in equal proportion.
Vertical loads (fixed loads) are carried mainly by the leg members, whereas transverse
and longitudinal loads are resisted jointly by the legs and diagonal bracing members.
Therefore separate application of, for example, all the vertical loads first and following
this with horizontal loads will give rise to overstressing of some of the members and
failures of some parts before the full 100% is reached for all loads. A system combining
horizontal and vertical test loads and applying this to towers as single combined loads at
the appropriate angles is followed at the Tower Testing and Research Station (TTRS),
Chennai. Rigging is simplified and it is possible to automatically allow for the weight of
ropes and for backstay forces in the tests. It also reduces the number of loads/ channels to
be controlled by about 50%.
5.4 Salient features of TTRS, CSIR-SERC, Chennai
The Tower Testing & Research Station was designed with Captronic Systems and the
control equipment is real time servo controlled system fabricated and supplied under the
guidance of Captronic Systems personals /authorities. It is located inside a horse-shoe
shaped enclosure provided by an abandoned granite quarry. This gives complete
protection against the actual prevailing wind conditions upto about 45 meters height of
tower. Testing accuracy depends upon the quality of the load cells, angle transducers and
the associated instrumentation. The load cells were manufactured incorporating the latest
developments in strain gauge materials and arrangement of strain gauge arrays to
minimise thermal excursions. The load cells were individually tested over the full
designed temperature range before acceptance. To maintain accuracy of load cells along
with their associated cable (fixed permanently) and the control channel are calibrated
prior to each test series and, if undamaged, recalibrated after the tests. A horizontal type
calibration rig compares the load cell output against a N.P.L calibrated proving ring.
All the loads are applied through servo-controlled hydraulic actuators having long strokes
that take care of system flexibility. The load cell and the angle transducer combination
attached to each loading point is connected to the actuators through high strength wire
ropes, pulleys, D-shackles etc. The actuators that apply the transverse loading on the test
tower are anchored firmly to a cellular box type reinforced concrete foundation structure
located on the hill at approximately 46m above the test pad level. The actuators are
positioned approximately 20o to the horizontal. The rigging wire ropes pass over vertical
pulleys positioned at different levels of a 23m high steel structure erected in front of the
transverse ram structure. Since the actuators are located at a vertical plane, changes in
rigging are possible during testing, in case of an actuator goes out of stroke due to large
structural deformations. The capacity of the transverse ram station is about 10000kN. For
applying broken wire loads, the longitudinal ram station is located on the hill at 65m
above the test pad level. For applying the loads on cross arms, a 37m long reinforced
concrete crest structure is located in front with specially developed horizontal and vertical
pulley system in steel, fixed on to it by special steel inserts located at 1m interval. The
steel fixtures can be so arranged to suit the width of the cross arm in actual test tower
within 10cm accuracy. The capacity of the longitudinal ram station is 4500kN.
The vertical ram station, located in front of test pad at a distance of about 50m and
slightly at a lower level, consists of 17 anchor blocks fixed with hydraulic actuators, for
applying the vertical load or used to control the angle of application of load. The capacity
of this station is 5000kN.
The test pad can accommodate towers having base dimension of upto 22.5 x 22.5m and
cross arms widths upto 36m. The test pad is anchored firmly to the granite rock base by a
series of double corrosion protected high strength anchors. Each anchor is stressed to
650kN. There are 3 rows of anchors along the central line in the transverse direction. This
enables the application of the vertical loads very accurately by choosing anchors at the
back rows to achieve effective vertical rigging scheme. A system comprising of couplers,
secondary anchor bars, holding down beams, and sturdy base plates with the stub angles
welded would be used for holding down the test tower on the test pad. A four storeyed
building near the vertical ram station houses the control room at its top floor, so that the
test tower can be viewed clearly and comfortably. The 40 channel control system with
continuous recording facility for applied loads and angles of their application is housed in
this floor. Fig.7 schematically shows the wire rope arrangement for the application of the
resultant load at the required angle for a typical load point on the tower. In the rope
applying the resultant load, a load cell to measure the load and an angle sensor to monitor
the angle of load application are attached.
The resultant load is controlled by the hydraulic ram at the top and the angle of
application is controlled by the vertical ram at the lower level. The double acting
hydraulic rams are remotely controlled from the control room loading channels. This
control is automatic by means of a closed loop servo system explained in the (Fig.7). The
closed loop system adjusts the resultant transverse and longitudinal components
simultaneously at all load points and ensures that the load are proportional from zero to
maximum. Limit switches help to see that the load applied never crosses the ultimate
value. The magnitude of load and angle of application are automatically recorded on
charts.
Fig. 7 Control system
5.5 Test procedure
Depending on the minimum and maximum values of the ultimate test loads required to be
applied at any particular load point during a series of test, a load cell for that load point is
chosen and the load cell-recorder combination is calibrated for the required ranges of the
load cell, allotting one control channel for each load cell. The details are tabulated in the
form of control room tables. The next stage is the rigging of the tower. A special software
is developed for this purpose at TTRS, which gives the rope diameter, the shackle size,
anchors at the pad level to be chosen, the rams to be connected etc. for a particular load
point. A typical graphic output is given in Fig.8.

Fig. 8 Longitudinal section of construction details


After completing the rigging of the tower (Fig.9), a trial test is conducted by applying
10% to 20% of the loading at all points, to check the rigging of the tower, the functioning
of the rams, the response of the recorder and the entire system behaviour. During the
actual test, the loads are first increased to 50% of the ultimate test loads and a pause of 1
min. is allowed for taking the deflection readings and strain gauge measurements. The
loads are then increased slowly in stages to 75%, 90%, and 95% and held for 1 min. After
reaching 100% of the ultimate test load a pause of 5 minutes depending on the clients /
codal requirement is allowed. The tower should take the load without any distress. If the
tower withstands the loadings without any damage then the test is considered to be a
successful one. The loadings are then slowly brought down to zero and the residual
deformations /stresses are recorded. The next loading is then applied.
Fig.9 : 80m high transmission line tower
6.0 SUMMARY

Evaluation of dynamic response of wind sensitive structures is quite a complicated and


complex procedure requiring the understanding of the behaviour of natural wind and it
aerodynamic effects on the structure and the structure characteristics. For the
convenience of designers, an equivalent quasi-static load is used to design a wind
sensitive structure by multiplying the mean wind loads with gust response factor. Here
the GRF is the ratio of peak response to mean response of the structure. Various authors
have derived semi-empirical expressions for the evaluation of the GRF and these
expressions have been used in most of the International Standards. However, these
expressions were derived by considering the contribution of the first mode of the
structure alone and also by assuming the first mode shape of the structure as linear.
However, when lattice towers are situated in a group as illustrated in this case study,
computation of wind loads on a given tower becomes extremely complex and currently
no analytical method is available for addressing this problem. Under such situations,
testing the models in a boundary layer wind tunnel is perhaps the only method presently
available. Prototype test is conducted to assess the strength of tower and tower like
structures. It is mainly to check the detailing of the tower in addition to the adequacy of
member sizes.

6.0 REFERENCES

1. Davenport, A.G., “Gust loading factors”, Journal of Structures Division, ASCE, Vol.
93, 1967, pp 11-34.
2. Vickery, B.J., “Along-wind loads and response”, Chapter 5, Wind Engineering
Course, Department of Mechanical Engineering, Monash University, 1992.
3. Vickery, B.J., “On the reliability of gust loading factors”, Proceedings of Technical
Meeting Concerning Wind Loads on Buildings and Structures, BSS 30, National
Bureau of Standards, Washington, D.C., pp 93-104.
4. IS:875(Part 3)-1987, “Indian Standard Code of Practice for Design Loads (other than
Earthquake) for Buildings and Structures, Part 3: Wind Loads”, Bureau of Indian
Standards, New Delhi, 1989.
5. AS 1170-2-1989, “Australian Standard: Minimum Design Loads on Structures, Part
2: Wind Loads”, Standards Australia, North Sydney, NSW, Australia, 1989.
6. BS 8100: Part 1: 1986, “British Standard Lattice Towers and Masts, Part 1, Code of
Practice for Loading”, British Standards Institution, London, 1986.
7. ASCE 7-88, “Minimum Design Loads for Buildings and Other Structures”, American
Society of Civil Engineers, New York, 1990.
8. Cook, N.J., “The designer’s guide to wind loading of buildings and structures – part
1”, Butterworths, London, 1985.
9. Harris, R.I., “The nature of wind”, Proceedings of the Seminar on The Modern
Design of Wind-Sensitive Structures, The Institution of Civil Engineers, London,
June 1970, pp 29-55.
10. Venkateswarlu, B., Arunachalam, S., Shanmugasundaram, J. and Annamalai, G.,
“Variation of wind speed with terrain roughness and height”, Journal of Institution of
Engineers (I), Vol.69, January 1989, pp 228-234.
11. Simiu, E. and Scanlan, R.H., “Wind effects on structures: an introduction to wind
engineering”, John Wiley & Sons, New York, 1978.
12. AS 3995-1994, “Design of steel lattice towers and masts”, Published by Standards
Australia.

You might also like