Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

Contents lists available at ScienceDirect

Journal of Wind Engineering & Industrial Aerodynamics


journal homepage: www.elsevier.com/locate/jweia

A review of transmission line systems under downburst wind loads


El-Sayed Abd-Elaal a, b, *, Julie E. Mills a, Xing Ma a
a
School of Natural and Built Environments, University of South Australia, Australia
b
Department of Structural Engineering, Faculty of Engineering, Mansoura University, Egypt

A R T I C L E I N F O A B S T R A C T

Keywords: Outages of power due to transmission tower failures can cause social and economic disasters. Investigations of
Downburst transmission line failures around the world have recorded that they are generally from high intensity winds from
Transmission towers downbursts and tornadoes. Downbursts represent the greatest threat due to the extreme and extended wind events
Structural analysis that they generate. Many studies have investigated the applications of these events on transmission line systems
Failure analysis (TLS). However, the wide ranges for the different downburst parameters and the varying representations of
Retrofitting
downburst wind speeds, which are different from boundary layer wind profiles, have complicated the investi-
gation of transmission line failures under these types of loads. This study reviews the research to date on TLS
under downburst wind loads. It explores downburst wind loads, their simulation models and the structural
behaviour of TLS under downburst wind loads. Modelling of TLS, static and dynamic analysis are all reviewed.
Failure analysis, critical downburst parameters, ideal retrofitting methods to avoid such catastrophic failures, and
optimization criteria of TLS are also discussed. Finally, recommendations for future research are made.

1. Introduction towers. Transmission line conductors have been studied alone and their
reactions have been imparted to towers at the connection points, and
More than 90% of transmission tower failures in Australia are due to towers have been investigated alone using a linear or nonlinear static
severe thunderstorm events that include downburst winds (Li, 2000), and analysis (Shehata et al., 2005; Shehata and El Damatty, 2007; Darwish.
the situation is similar in several regions around world which have 2010). By contrast, Yasui et al. (1999), Battista et al. (2003) and Gani and
comparable climatic conditions. The interruption of electricity due to Legeron (2010) highlighted the importance of the fluid flow–ca-
failure of transmission towers can generate social and economic di- bles–structure interaction when evaluating the towers' behaviour under
sasters. In addition the failure of one or two towers can trigger a long wind forces.
chain of failures, which can destroy several transmission towers in one Several studies investigated the structural response and failure anal-
event (Dempsey and White, 1996). ysis of transmission towers under downburst loads, but they did not
Savory et al. (2001) introduced the first model for isolated trans- consider retrofitting procedures. Some reinforcement methods exist for
mission towers under localised wind loads, but the unbalanced distri- upgrading transmission towers, such as the leg retrofitting method, dia-
bution of these types of loads on entire transmission line systems (TLS) phragm bracing, friction-type reinforcement and x-brace type. The effi-
pushed researchers to study the structural behaviour of entire TLS. ciency of these methods, convenience of the reinforcement, cost and
However, these studies were limited to stationary downbursts, which are optimal distribution of reinforcement through the TLS subjected to
different to the transient downbursts that are more usual for these events. downburst, are questions that need answers.
The translation speed increases the downburst velocity in the front of the Assurance of structural safety with optimal design is a basic objective
storm and reduces the downburst speed in the rear. There are also sug- in structural design and therefore modern TLS must be upgraded to
gestions that the translation speed causes some forward and backward confront this case of loading. The best or optimal arrangements of towers
gaps in distributions of horizontal wind, which could cause different in TLS as well as the best orientation in front of wind loads need
distribution of wind speed on several panels of a TLS. discussion.
Earlier studies focused on guyed transmission towers and treated a Several questions are suggested about the behaviour of TLS under
TLS as two separate parts: transmission line conductors and transmission downbursts, starting from modelling downburst events, modelling TLS,

* Corresponding author. School of Natural and Built Environments, University of South Australia, Australia.
E-mail address: abdee001@mymail.unisa.edu.au (E.-S. Abd-Elaal).

https://doi.org/10.1016/j.jweia.2018.07.004
Received 16 March 2018; Received in revised form 22 May 2018; Accepted 4 July 2018
Available online 11 July 2018
0167-6105/© 2018 Elsevier Ltd. All rights reserved.
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

applying these types of loads, developing the design parameters and that the cooling source models have succeeded in doing (Selvam and
retrofitting old towers. This paper provides a comprehensive review of Holmes, 1992; Vermeire et al., 2011a). Earlier studies have used three
earlier studies, including those of the authors, and then several proposals approaches for simulating downburst wind speeds: experimental, nu-
for further research. merical and analytical/empirical models. These will be considered in
turn.
2. Downburst wind loads
2.2. Experimental and numerical simulation models
Design specifications which assume that the atmospheric synoptic
wind is the basis of wind loads put TLS at risk due to localised wind Experimental and numerical simulation procedures have been con-
events such as tornadoes and downbursts. These two events pose a sig- ducted by many researchers. Bakke (1957) started an early experimental
nificant threat to transmission towers. The probability of occurrence of simulation for a wall jet. Holmes (1992) and Cassar (1992) modelled
localised wind events in a specific area is low, but the threats to an overall downburst wind speeds in a wind tunnel; Wood et al. (2001) employed
system grow due to the extension of TLS over long distances, thus the impinging jet models for different embankment heights; Chay and
increasing the probability of one of these events crossing the transmission Letchford (2002) studied the profiles of downburst winds using a sta-
line (Holmes et al., 2008). tionary wall jet tunnel then a moving downburst in a wind tunnel
Downbursts occur when warm air ascends through a cloud, and then (Letchford et al., 2002). Kim and Hangan (2007) investigated different
rises above the top of the cloud, creating a dome of warm air. The air scales of downbursts in wind tunnels and concluded that the impinging
cools at this level and then begins to drop, collapsing the dome and jet simulations are scale dependent. Mason et al. (2009) utilized the
rushing back to the ground, creating an outburst of damaging air cooling source model to study downburst storms. They used a dry,
(''Downbursts' of air are called danger to aircraft' 1979). The practical non-hydrostatic, sub-cloud and axisymmetric model. The sensitive wind
diameters of downburst are roughly 1 km and the extent of the outburst field parameters relating to variations in downburst size, initiation
flow is 1.0–6.0 km (Wilson et al., 1984). Hjelmfelt (1988) showed that height, and intensity, in addition to forcing duration and downburst
the downdraft diameter ranged from 1.5 to 3.0 km, based on 11 events. shape were determined. Later Mason et al. (2010) developed the previous
Analyses of extreme wind speeds in Australia showed that downbursts model from an axisymmetric model to a three-dimensional model.
are the extreme wind type at the height of 10 m (Holmes, 2002). Boss Table 1 summarises previous experimental and numerical simulation
(2010) reported that for every tornado damage report, there are almost findings. In the course of applying downburst wind loads to TLS,
10 downburst reports. adopting numerical or experimental models present several difficulties.
In addition to the complication of coupling numerical or experimental
simulation with structural analysis, there are other reservations. One is
2.1. Simulation of downburst loads
scale dependency. Kim and Hangan (2007) and Xu and Hangan (2008)
evaluated the scale effects for steady state and unsteady state simulation
The spread of downburst winds have been modelled using three
and confirmed the scale dependency.
different models: ring vortex, impinging jet model and cooling source
model. The ring vortex model forms a vortex ring before touching the
ground. The impinging jet model forms a radial out-flow after touching 2.3. Analytical/empirical models and turbulence component
the ground similar to the wind field, and the results from simulated
microburst events by using a large impinging jet were more consistent The downburst wind speed is described as the sum of the mean wind
with full-scale data (Letchford and Illidge, 1999). The impinging jet component (U), and turbulent wind component (u). Oseguera and
model has been further developed to include the formation of the ring Bowles (1988) and Vicroy (1991) developed an analytical model for the
vortex, but these models are not able to describe the buoyancy effects mean speed of a downburst wind in two components: radial speed and

Table 1
Experimental and numerical simulation findings.
Author Height of maximum horizontal velocity Radial position of maximum Comments/Findings
horizontal velocity

Hjelmfelt (1988) 50–100 m 0.75D to 1.0D Hjelmfelt (1988) investigated 11 field events and estimated rough dimensions
for downburst wind and vertical profiles of horizontal downburst wind speeds
Wood et al. (2001) 0.016D 1.5 D The height of maximum velocity increases with increasing radial distance.
Chay and Letchford (2002) 50–100 m 1.0D The mean pressure distributions on objects immersed in downburst events
differ from those in traditional boundary layer studies.
Hangan et al. (2003) 0.08D (numerical simulation) and from 1.0D (numerical Conducted experimental and numerical simulations and concluded that the
0.02D to 0.03D (experimental simulation) simulation) height of the maximum velocity decreases as Reynolds number increases.
Chay et al. (2006) 0.023D to 0.025D 1.0D to 1.25D Developed an analytical model for describing downburst mean and turbulent
wind components.
Kim and Hangan (2007) less than 0.05D 1.1D Conducted numerical simulations of impinging jet steady state and unsteady
state and concluded that the maximum velocity increases and the height of the
maximum velocity decreases as the Reynolds number increases.
Xu and Hangan (2008) 0.03D 1.1D Conducted numerical simulations and examined the different downburst
parameters, including cloud-base height, boundary conditions, scale and
terrain roughness
Mason et al. (2009) 0.011D 1.25D Conducted numerical simulations using cooling source model and concluded
that the height of the maximum velocity decreases with increasing downburst
diameter.
Vermeire et al. (2011) 0.015D 1.425D Compared between the impinging jet models and the cooling source models.
They found that the impinging jet model is not accurate, particularly for the
near surface out flow, and the magnitude of wind components are over-
predicted above the height of maximum radial velocity
Abd-Elaal et al. (2013a) 0.016D 1.46D Analysed several observed downburst events to estimate the time periods of
downbursts events

504
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

vertical speed. However, their model did not consider the outflow ring Darwish (2010) extracted the turbulence component from the
vortex and it was temporally independent. In addition there was a sig- recorded Lubbock-Reese downburst. He calculated the moving average
nificant difference between their profiles and the field and numerical of the speed over a certain period of time (filtering period) and subtracted
data. it from the total speed. He found that the maximum forces in guyed
Holmes and Oliver (2000) added the translation speed, and allowed transmission tower members were increased by 15–20% due to the
for downburst decay with time. Chay et al. (2006) extended the model downburst turbulence component.
with several modifications. They improved the height of maximum wind Several researchers have developed analytical models for simulating
speed as a function of radius. Li et al. (2012) introduced a further the downburst turbulence component. Chen and Letchford (2004)
improvement by including the nonlinear effects of boundary layer employed the evolutionary power spectral density method to generate
growth. Abd-Elaal et al. (2013b) presented a new pair of shaping func- the Gaussian stochastic process kðx; y; z; tÞ. Then, Chay et al. (2006)
tions which are more accurate for simulating the profiles of horizontal suggested the use of the Auto Regressive Moving Average method and
and vertical downburst wind speeds. Later Abd-Elaal et al. (2014) developed the turbulent wind speed (uÞ to be a function of more pa-
developed decay functions that can be added to the earlier empirical rameters including the horizontal position, as shown in Eq. (2).
expressions to depict the temporal profile of downburst wind speeds.
The turbulent wind component (u) was introduced by an amplitude- uðx; y; z; tÞ ¼ aðx; y; z; tÞkðx; y; z; tÞ (2)
modulating process in height and time parameters (Eq. (1)) (Chen and Su et al. (2015) evaluated these approaches to drive the time-varying
Letchford, 2004), where, aðz; tÞ and kðz; tÞ are the amplitude modulation mean wind component. They developed an empirical formula to deter-
function and the Gaussian stochastic process, respectively. mine the probable window sizes, taking into consideration the
time-varying trend of downbursts and the structural fundamental fre-
uðz; tÞ ¼ aðz; tÞkðz; tÞ (1)
quency. They also recommended a discrete wavelet transform with a
Holmes (2001) stated from the observed downburst at the Andrews higher order of Daubechies wavelet and Ensemble empirical mode
Air force base, Washington, D.C., 1983, that the turbulence amplitude is decomposition approaches.
0.1 times the non-turbulent wind speed, which means aðz; tÞ ¼ 0:1U ðz;
tÞ. Holmes and several researchers computed the turbulence intensity
2.4. Topographic effects
from experimental, numerical and field events as summarised in Table 2.
Their results showed that the turbulence intensity is in the range of
Few researchers have investigated the topographic speed-up factors
0.08–0.36 and it increases with increasing surface roughness.
for downburst wind speeds. Selvam and Holmes (1992) numerically
investigated topographic speed-up factors over a single hill, of slope 0.25,
and concluded that the speed-up factors are generally much lower than
Table 2
for boundary layer flow. Letchford and Illidge (1998) measured speed-up
Summary of downburst mean turbulence intensity.
factors for various topographic features, slopes 0.2–0.6, at a single po-
Author Mean turbulence intensity Comments/Findings sition, and then measured speed-up factors at various radii from the
Holmes (2001) 0.1 From the observed downburst at stagnation point (Letchford and Illidge, 1999). They concluded that the
the Andrews Air force base, crest speed-up factors increased with embankment gradient, and
Washington, D.C., 1983
decreased as the embankment was placed further from the impact point.
Chen and 0.1 Analysed the full-scale Lubbock-
Letchford Reese downburst event and Wood et al. (2001) studied topographic effects at several distances from
(2005) concluded that the turbulence the crest of the embankment with the testing surface at different dis-
variance equal to 0.03 tances from the jet outlet. They concluded that the proposed speed-up
Chen and between 0.08 and 0.11 Carried out multi-scale factors from Letchford and Illidge (1999) were slightly conservative.
Letchford correlation analysis of the
Later researchers worked to involve more topology features. Otsuka
(2006) Lubbock-Reese downburst event
and adjusted the mean (2006) and Mason et al. (2007, 2010) presented numerical simulations
turbulence intensity. for downburst outflow wind over two-dimensional ridges and 2D topo-
Holmes et al. between 0.9 and 0.11 Employed the moving average graphic features such as hills and escarpments. Abd-Elaal et al. (2018)
(2008) filter method to analyse the
investigated the changes in downburst wind speeds as they pass over real
Lubbock-Reese downburst event
Solari et al. 0.1 Analysed more than 90 topography. They concluded that the assumption of averaging the un-
(2015a, b) thunderstorm events and dulating terrain slopes over a horizontal length of 500 m was inadequate
concluded that turbulence and highlighted the importance of including three-dimensional simula-
intensity is independent of the tion for terrain topology rather than using 2D topographic features.
ratio between the height above
ground and the terrain
roughness length. 2.5. Downburst line
Elawady et al. between 0.11 and 0.14 Measured downburst turbulence
(2017) intensity from scaled
experimental testing at the A downburst line is defined as the occurrence of two or more
WindEEE dome. simultaneous downburst events which have outflows that generate a line
Stengel and between 0.1 and 0.3 Analysed a recent presumed of divergence outward from the line axis (Vermeire et al., 2011b).
Thiele downburst wind event and Around one-eighth of all downburst events are classified as a downburst
(2017) concluded that turbulence
intensity is time sensitive.
line (McCarthy et al., 1982; McCarthy and Wilson, 1986). However,
Aboshosha between 0.08 and 0.12 for Conducted a large eddy downburst lines represent a greater threat due to their large surface
et al. (2015) open exposure, and between impinging jet simulation on footprint, which increases the number of structures under threat at one
0.08 and 0.36 for urban different exposure surfaces and time compared with a single downburst event (Oliver et al., 2000).
exposures extracted the downburst
Vermeire et al. (2011b) studied downburst lines and found that their
turbulence component from
numerical simulation models. velocity profiles and the wind speed direction are inconsistent with those
They concluded that the expected for an isolated downburst event. The peak maximum speed in
turbulence intensity increases the downburst line has an amplification factor of up to 1.55 compared
with increasing surface with isolated downburst events and the damaging surface footprint
roughness.
increased by a factor of up to 70% compared with the isolated downburst.

505
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

2.6. Drag and lift coefficients By contrast, ASCE-74 (2010) does not include specific consideration
of downburst events and it refers to either the application of tornado-type
Drag and lift coefficients for transmission tower members, conductors narrow-front loading or a Gust-Front Factor (GFF) for extreme wind
and wires have been computed for horizontal winds in accordance with loadings, but its GFF considers the varyiation of the conductor height and
typical wind load codes, but these coefficients did not account for the the terrain exposure.
downburst cases where the wind loads are inclined. Mara (2007) studied Kwon and Kareem (2009) developed a gust-front factor (GFF) that
drag and lift coefficients for various vertical and horizontal angles of accounts for variations in load effects in gust-front winds. This GFF
wind projections using wind tunnel experiments, and calculated the drag summarises both the kinematic and dynamic features of gust-front
and lift coefficients for inclined wind. Later, Mara and Hong (2013) induced wind effects on structural systems. They recommended that
investigated the effect of wind direction on the response of a single this factor can be used in conjunction with existing design standards, as a
transmission tower. They concluded that the capacity curve of the tower device for treating conventional synoptic wind. However, this study was
is dependent on the wind direction. limited to building structures and they suggested that it could be
extended to more complex structures.
2.7. Standards and gust-front factor methods
2.8. Downburst wind modelling summary and recommendations for future
Most international standards and codes do not describe or design for research
downburst events and the prediction of downburst wind loads is ignored
in most of them. The Australian/New Zealand Standard (2010) “AS/NZS Downburst wind speeds have been simulated and produced through
7000:2010 Overhead line design” is one of the standards to consider the experimental, numerical and analytical approaches. However, there is a
effect of downburst winds through the introduction of two simple design lack of data for full-scale events to validate these results in the spatial and
charts for the vertical and the horizontal distribution of horizontal wind temporal domain. Most studies investigated the downburst as an axi-
speed, while the vertical wind speed is ignored. Aboshosha et al. (2016) symmetrical event while the parent cloud motion and subcloud wind
presented a comprehensive comparison between international code shear have rarely been included. However, the cloud and subcloud mo-
design limitations including design wind speeds, wind pressure, tion change the resultant downburst wind profile on extended structures
conductor forces, tower forces and the gust factors. such as TLS.
Fig. 1 presents the suggested Span Reduction Factor (SRF) for the Scale dependency of downburst simulation models has been
suggested distribution of horizontal wind pressure on extended structural confirmed by researchers, and this difficulty affects the simulation re-
systems from four sources; the Australian/New Zealand Standard “AS/ sults. Large scale numerical simulation has high computation cost. The
NZS 7000:2010 Overhead line design”, the estimation by Holmes et al. changes in downburst flow profiles with different simulation scales
(2008) for the Lubbock-Reese downdraft of 2002, the estimation by should be considered, particularly because most structural systems and
Aboshosha and El Damatty (2013) and the suggested one for the synoptic the main surface parameters such as surface roughness and terrain cat-
wind. This comparison indicates the large difference between the sug- egories are located in a very low height range of < D/50 where D is the
gested profiles for downburst events and the synoptic wind profile. downburst diameter.
However, there is no specific definition for SRF. Holmes et al. (2008) The effects of topography on the downburst wind speed and the
squared the time histories of wind from the Lubbock-Reese downdraft to corresponding speed-up factors still need further research, as most
become proportional to wind pressure, and then averaged them in groups studies have not simulated complex topology and were limited to 2D
to simulate the wind loads over various span lengths. Aboshosha and El topographic features while 3D modelling of different topographic fea-
Damatty (2013) defined the SRF as the ratio between the peak conductor tures is needed. In addition, the changes in the vertical downburst wind
reaction that accounts for the spatial correlations of the wind field, to the speed component above different topographic features have been
corresponding value based on the assumption of full correlation. ignored.

1.00
Peak Load Reduction Factor

0.80

0.60
SRF for synoptic wind
Holmes et al. (2008)
AS/NZS 7000:2010
Haitham andand
Aboshosha El El
Damatty (2013)
Damatty (2013)
0.40
0 200 400 600 800 1000 1200 1400
Width, (m)

Fig. 1. Span reduction factor for peak pressures.

506
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

More research should be conducted to develop a GFF or SRF for perpendicular linear springs. Other researchers suggested that the insu-
representing downburst wind events. Initially, a precise definition for lator strings act as a three-dimensional pendulum but they modelled
span reduction factor should be developed. This definition should reflect them as two node, three-dimensional truss elements, assuming that in-
the critical loading scenarios that lead to maximum forces in different termediate hinges are at the connection points between the insulators
TLS elements. Then, entirely recorded field events should be employed and tower cross-arms (Hamada and El Damatty, 2011). However, others
such as the investigated events by Holmes et al. (2008) and Stengel and modelled the insulators as 6 truss elements (Gani and Legeron, 2010).
Thiele (2017) as these events can reflect the real coherence in downburst These models allow the insulators to rotate freely and there was no limit
winds. The developed GFF or SRF might summarise the kinematic and for the swing angle. Yan et al. (2010) modelled the insulators as
dynamic features of wind effects on transmission line systems and multi-rigid-body systems connected with pins, the pins being modelled
consider the variations of the conductor height and the terrain exposure. by providing connector elements in ABAQUS (ABAQUS Users' Manual,
Lastly, it is widely predicted that stronger and more frequent extreme 2003), and the clamp being modelled as a rigid-body. Yang and Hong
weather events including downbursts will occur due to climate change (2016) modelled the insulator as two-node link element with both ten-
(Brook, 2013). These beliefs seem to have been confirmed by recently sion and compression capability.
recorded strong winds that have created havoc across Europe, America, Duranona and Cataldo (2009) suggested that downbursts can develop
Caribbean and Australia in the last decade. Therefore, there is a necessity intense winds which can cause excessive sway angles that bring the
to investigate the probability of changes to different downburst param- conductors to a closer distance to the towers. Battista et al. (2003)
eters such as event frequency, wind intensity, and downburst size that implied that the insulators are the most important component of the
can be expected due to climate change. system when considering the analysis of wind flow and TLS interactive
dynamic behaviour and response.
3. Structural analysis of TLS under wind loads

3.1. Modelling of TLS under downburst 3.2. Linear, non-linear analysis and joint slippage of transmission towers

Savory et al. (2001) developed different models for transmission Transmission towers are one of the most complicated forms of lattice
towers under tornadoes and downbursts, which were limited to isolated structure systems to analyse. The behaviour of transmission towers under
towers. However, the unequal distribution of wind loads which occurs complex loads cannot be predicted by using simple techniques and there
due to these types of events highlights the importance of studying the are significant differences between forces in members calculated from
entire structural systems instead of isolated towers. Shehata et al. (2005) linear analysis and those measured from full-scale tests (Albermani and
introduced a finite element model for whole guyed TLS under downburst Kitipornchai, 1992). Prasad Rao and Kalyanaraman (2001) studied the
wind loads. Later, Shehata and El Damatty (2007) studied the effect of non-linear behaviour of lattice towers, and concluded that methods of
downburst characteristics (jet velocity, location and diameter) in the design which are based on a linear analysis are not consistent with test
development of forces in transmission tower members. Darwish et al. results. Prasad Rao et al. (2010) investigated several types of premature
(2010) added the effect of turbulent speed to downburst mean speed failures during full-scale testing of transmission towers. They found that
during the investigation of guyed TLS. Then, Darwish (2010) studied a non-linear analysis gave results that were more accurate than the results
self-supported transmission line under stationary downbursts, and later obtained from linear analysis and it became possible to predict the
added the parent storm translation speed and investigated the trans- probable capacity of the transmission tower by using non-linear analysis.
mission line system under downburst loads. However, this study did not Albermani et al. (2009) described a non-linear analytical procedure to
consider the continuous change of the downburst centre with the parent predict the failure of transmission towers by employing beam-column
storm translation speed. and truss elements for modelling tower members. Oliveira et al. (2007)
Aboshosha and El Damatty (2015) developed a procedure to calculate investigated guyed transmission towers using three different procedures
the reactions of conductors subjected to downburst loads in the trans- to find which modelling procedure best represented the towers. The first
verse and the longitudinal directions. Elawady and El Damatty (2016) modelling method used truss elements; the second modelling method
presented a procedure to estimate the critical conductor's pretension used beam elements with rigid connections; and the third modelling
force subjected to downburst loads, while considering the different method used beams for the tower main legs and truss elements for the
geometric and material parameters affecting the behaviour of the con- bracing elements. They evaluated the three modelling methods by using
ductors. Elawady et al. (2017) conducted an aero-elastic scaled experi- static linear and non-linear, and dynamic analysis. They recommended
mental test of a multi-spanned transmission line at the WindEEE dome. using beam elements for main members and truss elements for bracing
Conductors and ground wires have been modelled in different ways. elements.
They have been modelled as a three-node isoparametric finite element Kitipornchai et al. (1994) studied theoretically the effect of bolt
(Desai et al., 1995). Others have modelled them as a two-dimensional slippage on the deflection and ultimate capacity for lattice towers and
consistent curved beam and each cable span was divided into 10 concluded that slippage of bolts has a significant effect on deflection but
consistent elements (Shehata et al., 2005). Bartoli et al. (2006) and Cluni no effect on the ultimate strength. Ungkurapinan et al. (2003) investi-
et al. (2008) modelled the cables as a set of 2-joint truss elements, gated experimentally several variables that influence joint slippage such
geometrically non-linear and mechanically-linear, and verified the model as: load applied, construction clearance and number of bolts. They
experimentally. Gani and Legeron (2010) and Yan et al. (2010) modelled developed a mathematical expression for the relationship between joint
the conductors and ground wires as tension-only truss elements. Later, slippage and the applied load. Jiang et al. (2011) studied the impact of
the conductors and ground wires were modelled as three-dimensional, joint modelling in lattice transmission towers. They conducted their
non-linear cable elements using an elastic cable formulation to simu- research experimentally and validated it numerically and concluded that
late the behaviour of the cable, with each cable span being divided into structural analysis models which ignored joint slippage effects and ec-
30 cable elements (Hamada and El Damatty, 2011). Yang and Hong centricities are approximate in representing the global response of lattice
(2016) modelled the conductors and ground wires as link elements, with towers measured in full-scale tests. They stated that the effect of joint
each span modelled using 30 two-node link elements. Elawady et al. slippage on the ultimate capacity alters according to the case of loading.
(2017) used aircraft cables to simulate the scaled bundle during scaled In the case of torsional loading there is no significant effect on the ulti-
experimental tests. mate capacity whereas in the case of flexural loading the ultimate ca-
Shehata et al. (2005) suggested that the insulator strings act as a pacity was reduced by approximately 15% compared with models
three-dimensional pendulum and hence can be modelled by two without joint slippage being considered.

507
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

3.3. Static and dynamic analysis of transmission line system calculating critical microburst parameters for each member in trans-
mission towers. Darwish (2010) gave priority to effects of varying r/D
For better understanding of the dynamic analysis of transmission line and θ rather than the effects of varying D. Later, El Damatty and Elawady
systems under downburst wind speeds, the distinction between the fre- (2018) appended transmission line parameters such as tower types,
quencies of the downburst mean wind component (non-turbulent) and spans, heights and conductor properties for the earlier study.
the turbulent component should be identified. Downburst wind loads are
not like synoptic wind loads, they produce very strong wind that changes 3.6. Failure analysis of transmission towers
its speed and direction in a very short time. Kim and Hangan (2007)
indicated from analysing a full-scale downburst event while considering Failure analysis of transmission towers is important for identifying
the translation velocity of downburst, that the downburst frequency is the critical zone or critical members in transmission towers and it assists
0.025 Hz (T ¼ 40 s) in the translation side and 0.01 Hz (T ¼ 100 s) in the in determining the required reinforcements if towers need upgrading and
opposite translation. Holmes et al. (2008) investigated the turbulence for their arrangement through the TLS. Prasad Rao et al. (2010) examined
a full-scale downburst; and found that the peaks occurred at a frequency different kinds of tower failures through full-scale testing of transmission
range of 0.005–0.4 Hz. The turbulence component for the same full scale towers, concluding that the failures could occur due to the bracing sys-
downburst event has been investigated by Darwish et al. (2010). They tems (220 kV DE type tower), due to redundant members (220 kV
found that the peak power spectrum occurs at frequencies less than DE-D9DT type tower), or due to main leg members (400 kV tangent
0.01 Hz. tower). Moon et al. (2009) prepared a half-scaled structure test to eval-
Regarding towers and conductors frequencies, Yasui et al. (1999) and uate the failure mode of transmission towers. They concluded that adding
Battista et al. (2003) have indicated that the natural frequency of
self-supporting transmission towers was 1.28 Hz and 1.35 Hz, respec-
Table 3
tively. Shehata et al. (2005) computed the natural frequencies for guyed Summary of static and dynamic analysis findings.
transmission towers and conductors as 1.73 Hz and 0.12 Hz, respectively.
Author System type Comments/Findings
There is a large gap between the mean wind frequency of 0.025 Hz
and the minimum frequency calculated for self-supporting transmission Battista Conducted analytical- The dynamic characteristics of the
et al. numerical modelling for self- transmission towers and the lateral
towers (1.28 Hz), guyed transmission towers (1.73 Hz) and conductors
(2003) supported towers under movement of the cables increase the
(0.12 Hz). Therefore, studying transmission tower systems under down- boundary layer winds importance of fluid
burst mean wind components will not need dynamic analysis. However, flow–cables–structure interaction,
the effect of the downburst turbulent component is still somewhat the height of the chain of insulators
ambiguous. Some scholars suggest that quasi-static analysis is sufficient, described the dynamic properties of
the TLS, and the chain of insulators
reasoning that there is large aerodynamic damping, and others promote
tended to move as a double
the importance of dynamic analysis. Table 3 summarises the static and pendulum
dynamic analysis results. Darwish Evaluated numerically the The boundary conditions have a
et al. dynamic characteristics of significant effect on the dynamic
(2010) conductors under the properties. They considered the
3.4. Influence of conductor bundling downburst turbulence influence of pre-tension forces on
component. natural frequency and calculated
the natural frequency, mode shape,
Long high voltage TLS are commonly constructed with conductor
sagging and stresses. They
bundles. Bundling of conductors has rarely been included in transmission concluded that quasi-static analysis
line studies, where wind actions on conductor bundles cause aero- is sufficient in assessing the
dynamic coefficients and vibrations that differ from those found on single downburst turbulence component
conductors. Cooper (1973) obtained from experimental data the aero- where the resonant component of
the turbulence is damped out by the
dynamic coefficients for a bundle of two conductors. Then, Braun and aerodynamic damping.
Awruch (2005) studied numerically the aerodynamic and the aeroelastic Lin et al. Investigated experimentally They initiated fifty-seven ensembles
characteristics of a bundle of conductors (Fig. 2). (2012) aeroelastic models for a single of downdraft outflow and
Riera and Oliveira (2010) researched the wind-structure interaction transmission line span under concluded that in most cases
simulated downburst wind downdraft outflow wind forces
of conductor bundles. They classified the models into four categories;
loads behaved in a quasi-static manner,
rigid bundle, flexible bundle, flexible interacting bundle, and fully but some trials did not and hence
interacting bundle according to the allowed motion for the bundle. They there is a need for further research.
obtained a dynamic response for a normal span under wind loads Yang and Conducted nonlinear inelastic They found a match between the
although they neglected the boundary conditions at the span end. Hong numerical analysis for capacity curves of a single tower
(2016) transmission tower-line with or without the end-springs and
system under downburst showed that including a dynamic
3.5. Critical downburst parameters loads. tower-line interaction has no effect
on the capacity curve of the tower
under downburst loads.
The downburst wind velocity and the associated loads vary with the Elawady Conducted an aero-elastic The responses of the towers and the
diameters of the downbursts, the coordinates of the centre of the et al. scaled experimental test of a conductors under high downburst
downburst, and the jet velocity. Each member in a transmission tower (2017) multi-spanned transmission wind speeds are mainly
line at the WindEEE dome. background, and therefore the
has different critical downburst parameters that produce maximum in-
quasi-static analysis is appropriate
ternal forces in those members. Shehata et al. (2008) developed a finite for investigating the response of
element optimization technique for calculating critical microburst pa- transmission line systems under
rameters for transmission tower members by using genetic algorithms downburst loads.
and grid mutation. They assumed the optimization parameters (D, r/D, θ) Stengel Investigated the equation of They found a big discrepancy
et al. motion for modelling between numerical and
as independent variables, for jet diameter (D) ranges between 500 and (2017) aerodynamic damping of experimental and theory results
2000 m, r/D ratio ranges between 0.0 and 2.2, and θ ranges between 00 nonlinear movement of after a certain level of deflection,
and 900 . They divided each parameter range into several divisions until conductors then they developed a formula to
the population had approximately 100 random instances. They employed suit the nonlinear movement of a
swaying cable.
genetic algorithms with grid mutation for developing the generations and

508
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

can be 1/3 the speed of the downburst (Holmes, 2001). Kwon and Kar-
eem (2009) concluded from several references that the storm translation
speed is around 10–20 m/s. This speed is a significant component and
must be added to the produced speeds from the downburst event, as well
as considering the movement of the downburst centroid.
Furthermore, the assumption of vector addition of parent storm
translation speed does not have any theoretical or numerical proof. The
authors suggest conducting numerical or experimental simulations for
both transient downbursts and isolated downbursts, and then adding the
translation speeds as a vector summation to the isolated downburst wind
speeds. Finally, this summation should be compared with the transient
downburst model in the four dimensions - at different height, different
radial distance, different angular direction and different time.
The structural behaviour of TLS under tilted downburst and down-
burst lines has not yet been investigated and knowledge of the behaviour
of TLS under different downburst sizes is incomplete. While Shehata et al.
Fig. 2. Wind streamline on two-conductor bundle (Braun and Awruch, 2005). (2005) and Shehata and El Damatty (2007) investigated the TLS for a
wide range of downburst diameters, they scaled this range from a single
additional braces may not improve the strength or stiffness of the downburst event and ignored scale dependency. The assumption of
transmission tower and that it was necessary to enlarge the main mem- scaling a range of events from a single downburst event may be
bers or add stiffeners to the weak joints. inadequate.
Finally, analysis of the wind-structure interaction of TLS under
downburst wind loads has been limited to a few experimental studies and
3.7. TLS modelling summary and recommendations for future research no numerical studies have been conducted, according to the authors'
knowledge.
According to the previous review, it is suggested that an analysis
should be developed that harvests the outcomes from all of these studies. 4. Upgrading of TLS
Finite element models should be developed, the main tower leg members
should be modelled as beam elements, the bracing members should be 4.1. Retrofitting transmission towers and transmission line systems
modelled as truss elements, and joint slippage should be included. The
conductors should be modelled as cable elements and the conductor The incompatibility of many old transmission towers with current
bundle should be modelled as a flexible bundle category. Insulator standards pushed researchers to investigate reinforcement methods.
models should be developed, the effect of the main features of insulators There are several reinforcement methods for upgrading old transmission
in terms of type, length, cross-section and swing angle limits should be towers such as the leg retrofitting method, diaphragm bracing system,
added, and the description of connections between insulator chains friction type reinforcement method and x-brace method.
should be introduced. The insulator model should include elements Albermani et al. (2004) used a diaphragm bracing system at mid
shaped like an insulator chain (Fig. 3a), and the type and nature of the height of diagonal members to upgrade old lattice towers. They tested the
connection between insulator strings should be considered. The main upgraded samples under two load cases - bending and torsional, and
different types of insulators have not yet been studied, so the three found that the buckling load capacity of diagonal compression members
common types of insulators: V-chain, I-chain and Relaxation chain need increased by 156–289%. They concluded that using a diaphragm is very
further research as well (Fig. 3b). effective for upgrading old transmission towers with high leg slenderness
In planning the construction of a TLS, the route of a TLS is divided ratios.
into straight sections. At each end of the sections there are tension towers The technique of leg reinforcement is based on attaching an extra leg
and suspension towers are applied between them. Therefore, the tension member to the existing leg using a number of interconnectors in each
towers should be provided for in the model and the interaction between panel. UniSA tower team (Denton et al., 2005; Zhuge et al., 2012)
these two types of towers should be studied. investigated experimentally the load sharing effectiveness for this rein-
The earlier studies did not include transient downbursts except for a forcement method by using three different types of connections (aligned,
single study conducted by Darwish (2010). Darwish (2010) added the alternate and cruciform connection). They concluded that the load is
translation speed as a vector summation for downburst speeds but transferred from the original leg to the reinforcing leg member for all
ignored the translation of the downburst centroid. The translation speed

a) Insulator chain b) Types of insulators (ESDEP n.d.)

Fig. 3. Transmission line insulators (ESDEP, n.d.).

509
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

kinds of interconnection, but the maximum load transfer occurred in the Using adhesive connections instead of welded connection is another
cruciform starred angle connection type. In addition, a series of mono- possibility, in conjunction with U-bolts. Pasternak et al. (2004) found
tonic tests (Mills et al., 2012) and numerical modelling studies (Lu et al., that adhesive connections showed higher strength and stiffness than
2014; Zhuge et al., 2012) on retrofitted tower structures showed an bolted connections, but adhesive connections have abrupt failures and
obvious structural strength increase in towers retrofitted through the leg their long term behaviour is unknown. However, if the adhesive has been
reinforcement method. Recent experimental studies (Lu et al., 2018) applied in conjunction with a bolted connection, their earlier limitations
indicated that long-term dynamic loading histories reduced the bolt-slip could be improved.
load and structural stiffness of cruciform connectors due to surface The authors would suggest researching new techniques and ap-
smoothing and bolt pretension loss, and thus reduced the load trans- proaches for reinforcing limited towers within TLS instead of all towers
ferring effectiveness between original members and reinforcing in TLS. For example, the wind load on transmission towers is calculated
members. from the wind span which is different from the weight span. The weight
Park et al. (2007) utilized friction type slotted connections to upgrade span is the horizontal distance between the lowest points of the con-
the resistance performance of transmission towers to wind loads and to ductors (Fig. 4). The wind span is the span where the wind acts trans-
lessen the force in the main structure members. They investigated two versely on the conductors and equals half the sum of the two spans. It is
units from the middle of the tower body and stated that friction-type formed in a horizontal plane so that the lowest points are at the middle of
reinforcement can fortify the transmission tower's main legs loaded each span. If wind loads are decreased on selected towers in a TLS, then
axially and improve energy dissipation capacity (Park et al., 2009). They an increase will result on the other towers. In this case, the first group of
introduced some modifications to slotted bolted connections to reduce towers will not require reinforcement and the other group will require
practical installation problems such as corrosion resistance at the sliding further reinforcement.
interface and variation of friction force through cycling loading. They If the rotation angle of the insulators is controlled (Fig. 5a), the dis-
recommended using stainless steel plates for slotted connections with tribution of wind loads between the neighbouring towers will be
grinded surfaces and to develop a new closed type of friction damper. changed. This could be achieved by adding elastic members for the
Limited research has been conducted on the x-brace reinforcement insulator chains that will give some fixation and will reduce their swing
method. Park et al. (2007) conducted a small numerical comparison angle (Fig. 5b). This change should apply for half the towers in TLS, for
between x-brace reinforcement methods and friction-type reinforcement example the odd towers. By preventing or reducing one end of the con-
methods. They found that tower legs bear most of the loads through their ductors from swinging and letting the other end swing freely, the position
cantilever action; and the braced members carry little of the load. of the lowest points in the wind span curve will transmit away from the
Tongkasame et al. (2007) compared this pair of methods and concluded odd towers and approach from the even towers. This technique will in-
that their effectiveness depends on the slenderness ratio of the original crease wind loads on the odd towers, which could then be reinforced, and
member, and that the x-brace method achieves good results for members will reduce wind loads on the even towers. It is proposed that this
with a high slenderness ratio of ℓ=i > 100. technique will be economical as it will reduce the number of reinforced
From a large scale view, cascading failures that trigger a long chain of towers to half the number.
failures of TLS cause expensive social and economic issues and strong
winds are the main reason of cascading failure accidents (Alminhana 4.3. Optimal design of TLS
et al. (2016). The CIGRE  Technical Brochure No. 515. (2012) presented
the different methodologies for mitigating cascading failures such as Optimal design of TLS involves many levels of optimization beginning
adding control sliding clamps, load-limiting cross arms, the Bonneville with the selection of conductor configurations and voltage, and ending
power administration approach, installation of anti-cascading towers at with the design details of tower components (Ghannoum and Yaacoub,
set intervals, and load reduction and control devices. However, most of 1989). Ranyard and Wren (1967) developed a procedure to obtain the
these techniques are applicable at early design stages, not as retrofitting best arrangement of towers in TLS, by studying the arrangement of
options to existing TLS. towers in one of their straight sections. They applied several restrictions
to the procedure; towers had to be spaced under a restricted maximum
4.2. Retrofitting summary and recommendations ratio (single span restriction), the distance between the conductors and
any point on the land must realize the minimum clearance, the summa-
Several scenarios are suggested for reinforcement: tion of the length of the adjacent spans must not be more than a restricted
maximum and each tower must carry at least 35% of the combined load
 Introducing reinforcement for critical members only in towers, of the two spans it carries.
whether these members are the main members, the bracing members White (1993) studied the optimization of conductors, the necessary
or the redundant members or critical joints. amount of steel for the reinforcement of conductors and the option of
 Introducing complete systems of reinforcement such as using dia- alloying. He achieved 30% saving in the total weight of support steel, but
phragm bracing, x-brace type methods, leg reinforcement or a com- did not discuss the cost of the alloying option and steel reinforcement.
bination of two methods. Shea and Smith (2006) improved the design of transmission towers
 Reinforcing using solid round steel members through shape and topology optimization. They reduced the structural
mass of existing towers by improving existing tower member section
These reinforcement procedures should then be evaluated under sizing, tower envelope, joint positions and topology. Guo and Li (2011)
downburst wind loads, as well as the general forms of wind loads. Their further developed the structural topology optimization methods, and
effectiveness, convenience of construction, cost and optimal distribution presented a layer combination optimization method to generate new
of reinforcement through the TLS, should all be considered. configurations.
Kumalasari et al. (2005) investigated reinforcement of solid round However, tower optimizations that are based on minimum weight
steel members by adding one or two split pipe(s) with and without end have severe handicaps, as the tower foundations and conductors repre-
welds in conjunction with using U-bolts for all cases. Using two split sent the major cost in TLS (Ghannoum and Yaacoub, 1989). They rec-
pipes without end welds increased compression strength by 30% and if ommended optimization procedures that combined the cost of
the end is welded the increase became 60%. This means that one split foundations and the cost of towers.
pipe with end welds has similar effectiveness to two split pipes without Darwish (2010) studied two guyed towers with heights of 44 m and
end welds. They recommended using two split pipes without end welds 55 m with different guy configurations under downburst loads and found
as the welding is hazardous in the field and very expensive. that increasing the height or changing the guy configuration had

510
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

It should also be noted that the construction of new transmission line


systems is decreasing in several regions around the world. In the 20th
century there were generally a few central power stations such as hy-
droelectric power stations or fossil fuel power stations and TLS networks
were required to transfer the high voltage generated electricity to the
cities. However, with the increasing use of local generation through solar
panels on building rooves, smaller capacity generation through wind
turbines, pumped hydro etc the generation can occur closer to the con-
sumer, so there is some reduced dependency on high voltage TLS. For
example, in 2017 the South Australian government preferred to invest in
and install the world's largest lithium ion battery instead of constructing
a new transmission line system. This may mean there is less need for
design optimisation research for new TLS, but the need for upgrading of
existing TLS is likely to remain for some time.
Fig. 4. Weight spans and wind spans.

5. Conclusion and recommendations for future work


negligible effect on the overall behaviour of tower members subjected to
downbursts. Downburst wind speeds vary significantly from the uniform and
Dagher and Lu (1993) investigated the reliability analysis of trans- symmetrical large-scale wind events. Downburst wind speeds vary in the
mission lines subjected to severe thunderstorms and local winds and four-dimensions of r; θ; h and t, where r is the radial distance from the
concluded that the probability of failure of the line of the transmission downburst centre, θ is the angular direction (angular direction is effective
line is proportional to their length. They posed questions about the op- for asymmetrical downburst events), h is the height above the ground
timum length for the transmission line and the optimum distance be- and t is the time. In addition to the 4-dimension variations, downburst
tween tension tower types. Later, El Damatty and Elawady (2018) wind flows are characterised by rapid time variations, highly correlated
investigated the implications of upgrading three different types of guyed wind, extreme wind speeds at very low heights, vorticity and negative
tower systems with similar spans. The ratio of required increase in the buoyancy, and the localised nature of events with respect to space and
weight of the towers due to downburst loads relative to the initial towers time. Consequently, downbursts result in unequal and different distri-
was found to be 3%, 23%, and 14%, This might point to the significant butions of wind loads over the TLS which differs to the distribution from
effects of the transmission line properties e.g. tower height, width of traditional boundary layer wind flows, and most design standards do not
conductor cross arm, number of conductors, dimensions of conductors, provide enough information about these events.
insulator length, etc. However, this study did not include the analysis of This paper presents a review of the research to date on downburst
the guys that support the towers, where a large percentage of the forces wind loads and ways of simulating them. The assumptions, advantages
transfer directly to the ground supports through the guys. and disadvantages of each procedure have been considered, and
modelling of TLS has been reviewed. Research to date has examined
4.4. TLS upgrade systems and design improvement summary several points that have significant effects on the analysis results, such as
linear and non-linear analysis, joint slippage effects, bundles of conduc-
It is clear that the optimal design of TLS is highly complicated. tors and critical downburst parameters. Finally, the different techniques
However, downburst wind is the governing type of wind load in several that mitigate tower failures have been discussed, ranging from studying
areas around the world, therefore, it is necessary to improve the design of the failure analysis of TLS, to the available retrofitting methods, with
new TLS to mitigate against these types of loads. several recommendations for further research. From this review, the
In general, the different parameters of TLS: spans, tower topology, following critical points can be concluded:
configuration, heights, tower orientations, etc. have been developed
without any consideration of downburst wind loads. It is suggested that  Downburst wind loads have been simulated by different procedures;
all of these parameters must be investigated under these types of loads. experimentally, numerically and analytically, however, these models
For example, the collected field data about transmission tower failures in have not yet been validated or calibrated sufficiently against full-scale
Uruguay under downburst wind loads proposed that the threat of data, which is very difficult to obtain due to the localised nature of
downburst damage increases when the transmission lines are orientated downburst events.
perpendicular to the storm direction (Duranona and Cataldo, 2009).  Although different studies have investigated TLS under downburst
Therefore, they raised questions about the best orientation of trans- wind loads, they were primarily limited to stationary downburst
mission line systems to mitigate the downburst threats.

a) Insulator swing angle

Fig. 5. Insulator swing angle and fixation.

511
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

events. However, moving downbursts could represent the critical Cluni, F., Gusella, V., Bartoli, G., 2008. Wind tunnel scale model testing of suspended
cables and numerical comparison. J. Wind Eng. Ind. Aerod. 96 (6–7), 1134–1140.
wind distribution on an extended structure such as a transmission line
Cooper, K.R., January, 1973. Wind Tunnel and Theoretical Investigations into the
system, due to the forward or backward gaps with horizontal distri- Aerodynamic Stability of Smooth and Stranded Twin-Bundled Power Conductors.
bution of downburst wind which occur due to the parent storm Technical Report, LTR-LA-115. National Research Council of Canada.
translation speed. Also, the structural behaviour of TLS under Dagher, H.J., Lu, Q., 1993. System reliability analysis of transmission lines. Eng. Struct.
15 (4), 251–258.
downburst lines has not yet been studied. Darwish, M., 2010. Characteristics and Design of Downburst Loaded Transmission Lines.
 The review highlighted several factors that should be considered Ph.D. thesis. The University of Western Ontario, Canada.
during modelling of TLS including: bundles of conductors, detailed Darwish, M., El Damatty, A.A., Hangan, H.M., 2010. Dynamic characteristics of
transmission line conductors and behaviour under turbulent downburst loading.
model for insulator that considers swing angle limit, cable-structure Wind Struct. 13 (4), 327–346.
interaction, and joint slippage effects. Dempsey, D., White, H.B., 1996. Winds wreak havoc on lines. T&D World Mag. 48 (6),
 Several reinforcement techniques have been presented, further de- 32–42.
Denton, J., Windsor, D., Mills, J., Tongkasame, C., Zhuge, Y., 2005. Effectiveness of load
velopments and new applications have been suggested. In addition, sharing connections for the reinforcement of steel lattice tower leg members. In:
several improvements have been proposed for optimum design of Australian Structural Engineering Conference 2005, Newcastle, Australia, September.
transmission line systems under downburst wind loads. Desai, Y.M., Yu, P., Popplewell, N., Shah, A.H., 1995. Finite element modelling of
transmission line galloping. Comput. Struct. 57 (3), 407–420.
Downbursts' of air are called danger to aircraft, 1979. The Globe and Mail 21.
References Duranona, V., Cataldo, J., 2009. Analysis of severe storms in Uruguay and their effect on
high voltage transmission lines. In: The 11th American Conference on the Wind
ABAQUS Users’ Manual, 2003. Ver. 6.4:Hibbitt[M]. Karlsson and Sorensen, Inc. Engineering (ACWE), San Juan, Puerto Rico, USA, June.
Abd-Elaal, E., Mills, J.E., Ma, X., 2013a. A coupled parametric-CFD study for determining Elawady, A., El Damatty, A., 2016. Longitudinal force on transmission towers due to non-
ages of downbursts through investigation of different field parameters. J. Wind Eng. symmetric downburst conductor loads. Eng. Struct. 127, 206–226.
Ind. Aerod. 123, 30–42. Elawady, A., Aboshosha, H., El Damatty, A., Bitsuamlak, G., Hangan, H., Elatar, A., 2017.
Abd-Elaal, E., Mills, J.E., Ma, X., 2013b. An analytical model for simulating steady state Aero-elastic testing of multi-spanned transmission line subjected to downbursts.
flows of downburst. J. Wind Eng. Ind. Aerod. 115, 53–64. J. Wind Eng. Ind. Aerod. 169, 194–216.
Abd-Elaal, E., Mills, J.E., Ma, X., 2014. Empirical models for predicting unsteady-state El Damatty, A., Elawady, A., 2018. Critical load cases for lattice transmission line
downburst wind speeds. J. Wind Eng. Ind. Aerod. 129, 49–63. structures subjected to downbursts: economic implications for design of transmission
Abd-Elaal, E., Mills, J.E., Ma, X., 2018. Numerical simulation of downburst wind flow lines. Eng. Struct. 159, 213–226.
over real topography. J. Wind Eng. Ind. Aerod. 172, 85–95. ESDEP, n.d. Lattice Towers and Masts. viewed 17 August (2011), http://www.fgg.uni-lj.
Aboshosha, H., El Damatty, A., 2013. Span reduction factor of transmission line si/kmk/esdep/master/wg15c/l0300.htm
conductors under downburst winds. In: Proceeding of the Eighth Asia-Pacific Gani, F., Legeron, F., 2010. Dynamic response of transmission lines guyed towers under
Conference on Wind Engineering, Chennai, India, December 10-14, 2013. wind loading. Can. J. Civ. Eng. 37 (3), 450–465.
Aboshosha, H., El Damatty, A., 2015. Engineering method for estimating the reactions of Ghannoum, E., Yaacoub, S.J., 1989. Optimization of transmission towers and foundations
transmission line conductors under downburst winds. Eng. Struct. 99, 272–284. based on their minimum cost. Power Delivery, IEEE Transactions on 4 (1), 614–620.
Aboshosha, H., Bitsuamlak, G., El Damatty, A., 2015. Turbulence characterization of Guo, H.Y., Li, Z.L., 2011. Structural topology optimization of high-voltage transmission
downbursts using LES. J. Wind Eng. Ind. Aerod. 136, 44–61. tower with discrete variables. Struct. Multidiscip. Optim. 43 (6), 851–861.
Aboshosha, H., Elawady, A., El Ansary, A., El Damatty, A., 2016. Review on dynamic and Hamada, A., El Damatty, A.A., 2011. Behaviour of guyed transmission line structures
quasi-static buffeting response of transmission lines under synoptic and non-synoptic under tornado wind loading. Comput. Struct. 89 (11–12), 986–1003.
winds. Eng. Struct. 112, 23–46. Hangan, H., Roberts, D., Xu, Z., Kim, J., 2003. Downburst simulation, Experimental and
Albermani, F., Kitipornchai, S., 1992. Nonlinear analysis of transmission towers. Eng. numerical challenges. In: Proceedings of the 11th International Conference on Wind
Struct. 14 (3), 139–151. Engineering, Lubbock, Texas.
Albermani, F., Kitipornchai, S., Chan, R.W.K., 2009. Failure analysis of transmission Hjelmfelt, M.R., 1988. Structure and life cycle of microburst outflows observed in
towers. Eng. Fail. Anal. 16 (6), 1922–1928. Colorado. J. Appl. Meteorol. 27 (8), 900–927.
Albermani, F., Mahendran, M., Kitipornchai, S., 2004. Upgrading of transmission towers Holmes, J.D., 1992. Physical modelling of thunderstorm downdrafts by wind tunnel jet.
using a diaphragm bracing system. Eng. Struct. 26 (6), 735–744. In: Second Australasian Wind Engineering Society Workshop, Monash University,
Alminhana, F., Albermani, F., Mason, M., 2016. A study of transmission line cascades 21–22 February 1992, 29–32.
triggered by conductor breakage and wind loads. In: The 18th Australasian Wind Holmes, J.D., Oliver, S.E., 2000. An empirical model of a downburst. Eng. Struct. 22 (9),
Engineering Society Workshop, McLaren Vale, Australia, 06–08 July 2016, American 1167–1172.
Society of Civil Engineers (ASCE), 2010. Guidelines for Electrical Transmission Line Holmes, J.D., 2001. Wind Loading of Structures. Spon Press, London.
Structural Loading. ASCE Manuals and Reports on Engineering Practice, No. 74, New Holmes, J.D., 2002. A Re-analysis of record extreme wind speeds in region a. Aust. J.
York, NY, USA. Struct. Eng. 4 (1), 29–40.
Australian New Zealand Standard, 2010. Overhead Line Design, Standards Australia/ Holmes, J.D., Hangan, H.M., Schroeder, J.L., Letchford, C.W., Orwig, K.D., 2008.
Standards New Zealand, Sydney. AS/NZS, vol. 7000, p. 2010. A forensic study of the Lubbock-Reese downdraft of 2002. Wind Struct. 11 (2),
Bakke, P., 1957. An experimental investigation of a wall jet. J. Fluid Mech. 2, 467–472. 137–152.
Bartoli, G., Cluni, F., Gusella, V., Procino, L., 2006. Dynamics of cable under wind action: Jiang, W.Q., Wang, Z.Q., McClure, G., Wang, G.L., Geng, J.D., 2011. Accurate modeling of
wind tunnel experimental analysis. J. Wind Eng. Ind. Aerod. 94 (5), 259–273. joint effects in lattice transmission towers. Eng. Struct. 33 (5), 1817–1827.
Battista, R.C., Rodrigues, R.S., Pfeil, M.S., 2003. Dynamic behaviour and stability of Kim, J., Hangan, H.M., 2007. Numerical simulations of impinging jets with application to
transmission line towers under wind forces. J. Wind Eng. Ind. Aerod. 91 (8), downbursts. J. Wind Eng. Ind. Aerod. 95 (4), 279–298.
1051–1067. Kitipornchai, S., Albermani, F., Peyrot, A.H., 1994. Effect of bolt slippage on ultimate
Boss, 2010. Downbursts, Tornadoes Cause Different Damage. Spokesman, Review. http:// behaviour of lattice structures. J. Struct. Eng-ASCE 120 (8), 2281–22817.
www.spokesman.com/stories/2010/mar/14/downbursts-tornadoes-cause-different- Kumalasari, C., Tickle, Vand, M., 2005. Compressive strength of solid round steel
damage/. members reinforced with split pipe(s). Civil Eng. Dimension 7 (2), 61–67.
Braun, A.L., Awruch, A.M., 2005. Aerodynamic and aeroelastic analysis of bundled cables Kwon, D.K., Kareem, A., 2009. Gust-front factor: new framework for wind load effects on
by numerical simulation. J. Sound Vib. 284 (1–2), 51–73. structures. J. Struct. Eng.-ASCE 135 (6), 717–732.
Brooks, H.E., 2013. Severe thunderstorms and climate change. Atmos. Res. 123 (0), Letchford, Illidge, G., 1998. Topographic effects in simulated thunderstorm downdrafts
129–138. by wind tunnel jet. In: Proceedings of the Seventh AWES Workshop, Auckland.
Cassar, R., 1992. Simulation of a thunderstorm downdraft by a wind tunnel jet. In: Letchford, Illidge, G., 1999. Turbulence and topographic effects in simulated downrafts
Summer Vacation Report, DBCE 92/22(M) CSIRO, Australia, February 1992. by wind tunnel jet. In: Proceedings of the Tenth International Conference on Wind
Chay, M.T., Albermani, F., Wilson, R., 2006. Numerical and analytical simulation of Engineering, Copenhagen.
downburst wind loads. Eng. Struct. 28 (2), 240–254. Letchford, C.W., Mans, C., Chay, M.T., 2002. Thunderstorms—their importancein wind
Chay, M.T., Letchford, C.W., 2002. Pressure distributions on a cube in a simulated engineering (a case for the next generation wind tunnel). J. Wind Eng. Ind. Aerod. 90
thunderstorm downburst-Part A: stationary downburst observations. J. Wind Eng. (12–15), 1415–1433.
Ind. Aerod. 90 (7), 711. Li, C., Li, Q.S., Xiao, Y.Q., Ou, J.P., 2012. A revised empirical model and CFD simulations
Chen, L., Letchford, C.W., 2004. A deterministic-stochastic hybrid model of downbursts for 3D axisymmetric steady-state flows of downbursts and impinging jets. J. Wind
and its impact on a cantilevered structure. Eng. Struct. 26 (5), 619–629. Eng. Ind. Aerod. 102, 48–60.
Chen, L., Letchford, C.W., 2005. Proper orthogonal decomposition of two vertical profiles Li, C.Q., 2000. A stochastic model of severe thunderstorms for transmission line design.
of full-scale nonstationary downburst wind speeds. J. Wind Eng. Ind. Aerod. 93 (3), Probabilist. Eng. Mech. 15 (4), 359–364.
187–216. Lin, W.E., Savory, E., McIntyre, R.P., Vandelaar, C.S., King, J.P.C., 2012. The response of
Chen, L., Letchford, C.W., 2006. Multi-scale correlation analyses of two lateral profiles of an overhead electrical power transmission line to two types of wind forcing. J. Wind
full-scale downburst wind speeds. J. Wind Eng. Ind. Aerod. 94 (9), 675–696. Eng. Ind. Aerod. 100 (1), 58–69.
CIGRE  Technical Brochure No 515, 2012. Mechanical security of overhead lines - Lu, C., Ma, X., Mills, J.E., 2014. The structural effect of bolted splices on retrofitted
containing cascading failures and mitigating their effects. In: International Council on transmission tower angle members. J. Constr. Steel Res. 95, 263–278.
Large Electrical Systems, Working Group B2.22.

512
E.-S. Abd-Elaal et al. Journal of Wind Engineering & Industrial Aerodynamics 179 (2018) 503–513

Lu, C., Ma, X., Mills, J.E., 2018. Cyclic performance of bolted cruciform and splice Shea, K., Smith, I.F.C., 2006. Improving full-scale transmission tower design through
connectors in retrofitted transmission tower legs. Thin-Walled Struct. 122, 264–285. topology and shape optimization. J. Struct. Eng. 132 (5), 781–790.
McCarthy, J., Wilson, J., Wand Fujita, T.T., 1982. The joint airport weather studies Shehata, A.Y., El Damatty, A.A., Savory, E., 2005. Finite element modeling of
project. Bull. Am. Meteorol. Soc. 63, 15–22. transmission line under downburst wind loading. Finite Elem. Anal. Des. 42 (1),
McCarthy, J., Wilson, J., 1986. Classify, locate and avoid wind shear (CLAWS) project at 71–89.
Denver”s international airport: operational testing of terminal weather hazard Shehata, A.Y., El Damatty, A.A., 2007. Behaviour of guyed transmission line structures
warnings with emphasis on microburst windshear. American Institute of Aeronautics under downburst wind loading. Wind Struct. 10 (3), 249–268.
and Astronautics Monographs 32, 17–26. Shehata, A.Y., Nassef, A.O., El Damatty, A.A., 2008. A coupled finite element-
Mara, T.G., 2007. The Effects of Multi-directional Winds on Lattice Sections, M.Eng. optimization technique to determine critical microburst parameters for transmission
Dissertation. University of Western Ontario, London, Canada. towers. Finite Elem. Anal. Des. 45 (1), 1–12.
Mara, T.G., Hong, H.P., 2013. Effect of wind direction on the response and capacity Solari, G., De Gaetano, P., Repetto, M., 2015a. Thunderstorm response spectrum:
surface of a transmission tower. Eng. Struct. 57, 493–501. fundamentals and case study. J. Wind Eng. Ind. Aerod. 143, 62–77.
Mason, M.S., Wood, G.S., Fletcher, D.F., 2007. Impinging jet simulation of stationary Solari, G., Burlando, M., De Gaetano, P., Repetto, M.P., 2015b. Characteristics of
downburst flow over topography. Wind Struct. 10 (5), 437–462. thunderstorms relevant to the wind loading of structures. Wind Struct. 20 (6),
Mason, M.S., Wood, G.S., Fletcher, D.F., 2009. Numerical simulation of downburst winds. 763–791.
J. Wind Eng. Ind. Aerod. 97 (11–12), 523–539. Stengel, D., Thiele, K., 2017. Measurements of downburst wind loading acting on an
Mason, M.S., Fletcher, D.F., Wood, G.S., 2010. Numerical simulation of idealised three- overhead transmission line in Northern Germany. Procedia Eng. 199, 3152–3157.
dimensional downburst wind fields. Eng. Struct. 32 (11), 3558–3570. Stengel, D., Thiele, K., Clobes, M., Mehdianpour, M., 2017. Aerodynamic damping of
Mills, J.E., Ma, X., Zhuge, Y., 2012. Experimental study on multi-panel retrofitted steel nonlinear movement of conductor cables in wind tunnel tests, numerical simulations
transmission towers. J. Constr. Steel Res. 78, 58–67. and full scale measurements. J. Wind Eng. Ind. Aerod. 169, 47–53.
Moon, B.W., Park, J.H., Lee, S.K., Kim, J., Kim, T., Min, K.W., 2009. Performance Su, Y., Huang, G., Xu, Y.L., 2015. Derivation of time-varying mean for non-stationary
evaluation of a transmission tower by substructure test. J. Constr. Steel Res. 65 (1), downburst winds. J. Wind Eng. Ind. Aerod. 141, 39–48.
1–11. Tongkasame, C., Mills, J., Zhuge, Y., 2007. Finite element modelling of steel lattice tower
Oliveira, M., Silva, J., Vellasco, P., Andrade, S., Lima, L., 2007. Structural analysis of legs reinforced for increased loads. In: The 4th International Structural Engineering
guyed steel telecommunication towers for radio antennas. J. Braz. Soc. Mech. Sci. and Construction Conference (ISEC-4), Melbourne, Australia, September.
Eng. 29, 185–195. Ungkurapinan, N., Chandrakeerthy, S.R.D.S., Rajapakse, R.K.N.D., Yue, S.B., 2003. Joint
Oliver, S.E., Moriarty, W.W., Holmes, J.D., 2000. A risk model for design of transmission slip in steel electric transmission towers. Eng. Struct. 25 (6), 779–788.
line systems against thunderstorm downburst winds. Eng. Struct. 22 (9), 1173–1179. Vermeire, B.C., Orf, L.G., Savory, E., 2011a. Improved modelling of downburst outflows
Oseguera, R.M., Bowles, R.L., 1988. A simple, analytic 3-dimensional downburst model for wind engineering applications using a cooling source approach. J. Wind Eng. Ind.
based on boundary layer stagnation flow. NASA Tech. Memo. 100632. Aerod. 99 (8), 801–814.
Otsuka, K., 2006. Effects of topography on the surface wind of an isolated wet microburst. Vermeire, B.C., Orf, L.G., Savory, E., 2011b. A parametric study of downburst line near-
In: 4th International Symposium on Computational Wind Engineering, Yokohama. surface outflows. J. Wind Eng. Ind. Aerod. 99 (4), 226–238.
Park, J.H., Moon, B.W., Min, K.W., Lee, S.K., Kyeong Kim, C., 2007. Cyclic loading test of Vicroy, D.D., 1991. A simple, analytical, axisymmetric microburst model for downdraft
friction-type reinforcing members upgrading wind-resistant performance of estimation. NASA Tech. Memo. 104053.
transmission towers. Eng. Struct. 29 (11), 3185–3196. White, H.B., 1993. Guyed structures for transmission lines. Eng. Struct. 15 (4), 289–302.
Park, J.H., Moon, B.W., Min, K.W., Lee, S.K., Kyeong Kim, C., 2009. Experimental study Wilson, J.W., Roberts, R.D., Kessinger, C., McCarthy, J., 1984. Microburst wind structure
on the energy dissipation capacity of a plane transmission tower substructure with and evaluation of Doppler radar for airport wind shear detection. J. Clim. Appl.
friction-type reinforcing members. Power Deli. IEEE Trans. 24 (4), 2062–2070. Meteorol. 23 (6), 898–915.
Pasternak, H., Schwarzlos, A., Schimmack, N., 2004. The application of adhesives to Wood, G.S., Kwok, K.C.S., Motteram, N.A., Fletcher, D.F., 2001. Physical and numerical
connect steel members. J. Constr. Steel Res. 60 (3–5), 649–658. modelling of thunderstorm downbursts. J. Wind Eng. Ind. Aerod. 89 (6), 535–552.
Prasad Rao, N., Kalyanaraman, V., 2001. Non-linear behaviour of lattice panel of angle Xu, Z., Hangan, H.M., 2008. Scale, boundary and inlet condition effects on impinging jets.
towers. J. Constr. Steel Res. 57 (12), 1337–1357. J. Wind Eng. Ind. Aerod. 96 (12), 2383–2402.
Prasad Rao, N., Knight, G.M.S., Lakshmanan, N., Iyer, N.R., 2010. Investigation of Yan, B., Lin, X., Luo, W., Chen, Z., Liu, Z., 2010. Numerical study on dynamic swing of
transmission line tower failures. Eng. Fail. Anal. 17 (5), 1127–1141. suspension insulator string in overhead transmission line under wind load. Power
Ranyard, J.C., Wren, A., 1967. The optimum arrangement of towers in an electric power Deli. IEEE Trans. 25 (1), 248–259.
transmission line. Comput. J. 10 (2), 157–161. Yang, S.C., Hong, H.P., 2016. Nonlinear inelastic responses of transmission tower-line
Riera, J.D., Oliveira, T.T., 2010. Wind-structure interaction in conductor bundles in system under downburst wind. Eng. Struct. 123, 490–500.
transmission lines. Struct. Infrastruct. E. 6 (4), 435–446. Yasui, H., Marukawa, H., Momomura, Y., Ohkuma, T., 1999. Analytical study on wind-
Savory, E., Parke, G.A.R., Zeinoddini, M., Toy, N., Disney, P., 2001. Modelling of tornado induced vibration of power transmission towers. J. Wind Eng. Ind. Aerod. 83 (1–3),
and microburst-induced wind loading and failure of a lattice transmission tower. Eng. 431–441.
Struct. 23 (4), 365–375. Zhuge, Y., Mills, J.E., Ma, X., 2012. Modelling of steel lattice tower angle legs reinforced
Selvam, R., Holmes, J.D., 1992. Numerical simulation of thunderstorm downdrafts. for increased load capacity. Eng. Struct. 43, 160–168.
J. Wind Eng. Ind. Aerod. 44 (1–3), 2817–2825.

513

You might also like