Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Engineering Structures 112 (2016) 23–46

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Review article

Review on dynamic and quasi-static buffeting response of transmission


lines under synoptic and non-synoptic winds
Haitham Aboshosha a, Amal Elawady b, Ayman El Ansary b, Ashraf El Damatty b,⇑
a
Boundary Layer Wind Tunnel Laboratory (BLWTL), Western University, London, ON, Canada
b
Civil and Environmental Engineering Department, Faculty of Engineering, Western University, London, ON, Canada

a r t i c l e i n f o a b s t r a c t

Article history: This study reviews the literature on the dynamic response of a Transmission Line (TL) system under syn-
Received 22 July 2015 optic wind (conventional atmospheric boundary layer) as well as non-synoptic wind loading (down-
Revised 23 December 2015 bursts). Gust-induced response for the conductors and the towers are covered and the limitations in
Accepted 4 January 2016
the current structural design codes for wind loading are identified. Three main sections are considered
Available online 19 January 2016
in this study covering synoptic wind loading, downburst, and main conclusions and recommendations.
For the case of synoptic wind events, four design codes (ASCE 74 2010, AS/NZS 2010, BS 2001, IEC
Keywords:
2003) specialized in TLs are considered for comparison. Using the ASCE 74 as a datum for normalization,
Dynamic response
Transmission line
a code ratio (CR) is evaluated for various parameters to assess the discrepancy between the codes. The
Conductors code ratio for conductor forces CRFc is found to be ranging between 0.81 and 1.44. For tower forces code
Synoptic winds ratio CRFt, a discrepancy range of 0.68 and 1.85 is noticed. The study highlights the main reasons behind
Non-synoptic downburst these discrepancies. For the case of downbursts, the study reveals that the event’s size and its relative
Gust factor location to the tower lead to a number of critical load cases that need to be considered. The study pro-
Buffeting vides important design considerations for both synoptic and non-synoptic winds. At the end of the study,
a list of the main gaps existing in current design codes and recommendations to fill out these gaps is
provided.
Ó 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2. Dynamic and quasi-static responses of TLs under synoptic winds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.1. Literature on the dynamic buffeting response of transmission lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2. Approach used by various codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.1. Design velocity Vd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.2. Design wind pressure qd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.3. Conductor force Fc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.4. Conductor gust response factor GC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.5. Tower force FT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.6. Tower gust response factor GT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3. Quantitative comparison between codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3.1. Design wind velocity and pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3.2. Conductor forces and the gust factor GC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3.3. Tower forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3. Response of TLs under downburst wind loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1. Field measurements and numerical modeling of downburst . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.1.1. Field measurements during downbursts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.1.2. Downburst numerical, analytical and empirical modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

⇑ Corresponding author.
E-mail addresses: haboshos@uwo.ca (H. Aboshosha), damatty@uwo.ca (A. El Damatty).

http://dx.doi.org/10.1016/j.engstruct.2016.01.003
0141-0296/Ó 2016 Elsevier Ltd. All rights reserved.
24 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

3.1.3. Downburst wind field characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36


3.2. Structural modeling of a transmission line under downburst loading. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.2.1. Lattice tower and guy wires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.2.2. Conductor lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3. Behavior of transmission lines subjected to downbursts and corresponding critical load cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Appendix A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

1. Introduction dynamic behavior of TLs is discussed in detail. The procedures used


by four major design codes to obtain the design forces on the con-
Electricity is carried by Transmission Lines (TLs) from the ductors and the towers are compared to identify the sources of dis-
source of power generation to the distribution system. Overhead crepancy between different codes’ approaches.
transmission lines consist mainly of support towers, conductors,
insulators and ground wires. Conductors are responsible for trans- 2.1. Literature on the dynamic buffeting response of transmission lines
mitting the electricity and they are attached to the towers using
insulators. Ground wires protect the line from lightning strike. Dynamic responses of transmission lines can be classified into
Optimal TL design is particularly important when the site of power the following groups: (i) buffeting due to the incoming turbulence,
generation is geographically remote from population centers, such (ii) vortex shedding and (iii) galloping. The current study focuses
as in the case of many hydroelectric dams in North America. The on the dynamic buffeting response of transmission lines which
ubiquity and full integration of electronics into modern life means affects both the lines and the towers. Vortex shedding and gallop-
that power outages due to TL failure are unacceptable from the ing affect mainly the lines and can lead to adverse effects such as
standpoint of both social and economic losses. Past reports of TL clashing of wires, excessive conductor sags, wire fatigue, and gen-
failure due to weather conditions [37,79], including High Intensity erally collapse of line components. For deeper explanation of such
Wind (HIW) events (downbursts and Tornadoes), emphasize the phenomena, authors recommend reading studies by Den Hartog
importance of accurate design wind loads. [40], Davison et al. [36], Havard and Pohlman [59], Rawlins [109]
The current study summarizes the previously conducted work and more recent studies such as Macdonald et al. [90], Ohkuma
related to the buffeting dynamic behavior of TL under synoptic and Marukawa [107], Gurunga et al. [54], Chabart and Lilien [24],
and non-synoptic winds in order to identify the limitations and Dyke and Laneville [44], Boddy and Rice [21], Jones [72], Lilien
gaps in the current codes and to suggest how these gaps can be et al. [89], Yan et al. [142], and Ma et al. [91].
closed. The study is structured in the form of three main sections. Buffeting dynamic response of the conductors is governed by
Dynamic response of transmission line structures under synoptic the fluctuating incoming wind speeds and the properties of the
winds is discussed in Section 2. In this section, four design codes conductors including their damping. Damping of the conductors
[12,13,20,69] specialized in TLs are considered to compare results from a structural contribution and an aerodynamic contri-
between the approaches utilized to estimate the forces acting on bution. Structural contribution of the damping is typically minor
the conductors and the towers under synoptic winds. This compar- in the order of 0.05% as indicated by Bachmann et al. [15] and thus
ison is provided in a tabulated format. For each quantity, a code can be neglected. Aerodynamic contribution to damping, which is
ratio (CR) is obtained by normalizing the values predicted from typically referred to as aerodynamic damping fa , is usually domi-
the three codes [13,20,69] by the value obtained from ASCE 74 nant and can be calculated using the expression proposed by
[12]. Section 3 covers the response of TL structures under down- Davenport [31] and is shown in Eq. (1) [101].
burst winds. This section provides a description of the downburst
wind field and a literature review showing previous field measure- pC d qD2 V
fa ¼ ð1Þ
ments and numerical modeling of downbursts. The last section 4 m f D
(Section 4) identifies gaps in design codes and recommendations
where Cd is the drag coefficient of the conductor and can be taken
to fill out these gaps for both synoptic and non-synoptic wind cases
equal to 1.0 according to the ASCE 74 [12] and AS/NZS [13]; q: air
studied in Sections 2 and 3. This section starts by presenting the
density; m: mass per unit length; D: Conductor diameter; V:  mean
gaps existing in the current design codes related to the gust
response of TLs under synoptic wind and recommendation for fill- wind velocity; f : conductor frequency.
ing these gaps. The second part of Section 4 provides the main con- As indicated from the equation above, the aerodynamic damp-
 and
ing, fa , is directly proportional to the mean wind velocity, V,
clusions and recommendations for the downburst case.
is inversely proportional to the line mass, m. This contribution
can rise up to 60% of the critical damping according to Loredo-
2. Dynamic and quasi-static responses of TLs under synoptic Souza and Davenport [83] for light weight conductors subjected
winds to high velocities (i.e. a critically damped system returns to its
equilibrium position without oscillating).
This section discusses the response of TLs to fluctuating synop- The study conducted by Loredo-Souza and Davenport [83] con-
tic winds and is divided into two main parts. In the first part (Sec- cluded that the background response is, indeed, the main contrib-
tion 2.1), a literature review on the buffeting dynamic behavior of utor to the total fluctuating response. However, the resonant
TLs is presented emphasizing the main parameters affecting the component can be also important if the line characteristics and
response. In addition, a literature review on field tests conducted wind velocities lessen the aerodynamic damping. There were only
during the last few decades to measure loads on conductors and minor resonant effects for the cases of high aerodynamic damping.
towers has been provided. In the second part (Section 2.2), the For low aerodynamic damping, dynamic effects should be
quasi-static method used by design codes to account for the considered but confirmation of a threshold value of aerodynamic
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 25

damping, below which dynamic behavior is of significant concern, For a freestanding tower, Holmes [64] studied the dynamic
requires further research. response of single tower (without conductors) to address the gust
Later on, Loredo-Souza and Davenport [84,85] conducted vari- response factors. These factors were derived assuming a linearly
ous wind tunnel tests on conductors subjected to turbulent synop- tapered tower. The gust factors considered the height effect, mean
tic wind and compared the conductor responses with those velocity profile, mode shape of the tower, and the type of the load.
obtained theoretically using the statistical method employing the Later, Holmes [65] continued what he reported in part ‘‘I” to
influence lines [35]. The study indicated a very good agreement include an expression of the aerodynamic damping of a freestand-
between the experimental and the theoretical results. This statisti- ing tower. In addition, Holmes [65] provided an expression to eval-
cal method is general for any structure. It accounts for both the uate the top deflection of the tower. In addition, Holmes [66]
background and the resonant components and considers the effect extended his comprehensive study of part I and II and reported sta-
of higher modes on the resonant component. tic load distribution of mean, background fluctuating, and resonant
Also several full-scale conductor wind loading experiments components of the wind loads on freestanding lattice tower.
have been conducted during the last few decades to acquire char- The natural frequency of a typical tower, with a 60 m height or
acteristics of wind and corresponding conductor response. Shan less, is typically higher than 1 Hz [69,135,81]. That is significantly
[116] conducted a subsequent investigation to evaluate several higher than the frequencies corresponding to the maximum turbu-
past full-scale conductor wind loading experiments, particularly lence energy, which leads to a negligible resonant response compo-
the experiments conducted by the Bonneville Power Administra- nent for the towers. Harikrishna et al. [58] studied the peak base
tion [133], Ontario Hydro [75], Hydro Quebec [67], and the EPRI moment of a typical transmission tower, which was mainly due
Transmission Line Mechanical Research Center (TLMRC) [115]. to the background component, and found that the peak base
For example, Bonneville Power Administration [133] conducted moment predicted by the Indian code [68] and the Australian code
field measurements on a 500 kV lattice tower subjected to extreme [13] agreed with field measurements, whereas the British code
wind conditions. The tower was attached to conductors with spans [19] overestimated the base moment by approximately 30%. Full-
252 m and 450 m on the sides. The longer span was passing over a scale measurements of axial forces in tower legs by Savory et al.
valley while the shorter span was passing over a flat terrain. A total [110] agreed with predictions from the British code [19], however
number of 23 records of the wind speed, wind direction and struc- a higher sampling rate than available was needed for the predic-
tural response of a transmission line were recorded. The obtained tion of the dynamic structural response.
records were extensively analyzed by Mehta et al. [98] and Mehta On the other hand, taller towers with lower natural frequencies
and Kadaba [99,100] aiming to improve the analytical model are expected to be dynamically excited and the literature on tall
developed by Davenport [34] to predict conductor response in guyed masts for telecommunications applications describes find-
extreme wind. Based on their studies the following points were ings from studies concerning isolated lattice-frame support struc-
found: (1) Winds traversing over the valley showed a wide varia- tures. For example, analytical and full-scale vibration
tion in mean profile and turbulence; (2) Effective drag coefficient measurements up to 341 m height on a guyed mast subjected to
of the conductor was found to vary between 0.48 and 0.75; (3) strong winds showed the importance of including the nonlineari-
Noticeable resonant peaks occurred in the frequency range from ties in the guys [108]. Frequency domain analysis [127] has yielded
0.1 to 0.4 Hz in the conductor response spectra; (4) Most of the reliable estimates of peak responses when the effective stiffness of
fluctuating conductor responses (75%) appeared to be due to the guys, obtained from a static analysis under mean wind loads, was
background component and a minor contribution of the resonant considered. Aeroelastic model experiments by Lou et al. [88] in a
component was found. This was due to the high conductor aerody- wind tunnel found that a tall tower (180 m full-scale height) was
namic damping in the order of 40%. sufficiently flexible to be susceptible to dynamic excitation by
the wind and the resonant component could not be neglected.
However, the effects from the conductors were neglected, which
is a problematic assumption since field measurements indicated
-

that the vibration characteristics of towers were greatly affected


by the conductors [101]. Yasui et al. [143] conducted an analytical
study of wind-induced vibration of conductors supported by either
a suspension or a tension tower. The difference in the way the tow-
Influence Line of F

ers supported the conductors had a noticeable influence on the


response characteristics. Furthermore, from a three-dimensional
FEM model, Battista et al. [16] found that the first mode of vibra-
tion involves both tower and conductor motions; although
increases in conductor tension due to the wind load were not
included in the analysis of the natural frequency and a 2% damping
+

ratio was assumed across all vibration modes, which appears to be


insufficient damping especially for the modes which are mainly
due to conductor motion.
F Lin et al. [81] tested an aeroelastic model of a guyed mast, at
1:100 length-scaling, in a wind tunnel. A half-span of lines on
either side of the mast was modeled, thus full- and multi-span
motions were neglected. Loredo-Souza and Davenport [86] com-
pared the peak forces developing in different tower members
obtained using various codes with those obtained using the statis-
tical methods employing the influence line [35]. Their results indi-
cated a possible 20% underestimation of the peak forces obtained
by the codes than those obtained by the statistical method. This
Fig. 1. Influence line for the member force that has a changing sign (modified from underestimation is a result of neglecting the contribution of higher
Cook [27]). modes that happens in the members having an influence line with
26 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

Table 1
Summary of previous studies’ scope and main finding.

Reference Scope Main finding/conclusion


Davenport [31] Studied the dynamic behavior of suspension Proposed an expression for cable aerodynamic damping
bridge cables
Loredo-Souza and Davenport [83–85] Conducted aeroelastic wind tunnel tests for Background response was the main contributor to the total fluctuation response
transmission line wires
Wang et al. [135] Studied the dynamic behavior of Transmission towers can be considered non-dynamically excited
Lin et al. [81] transmission lines
Loredo-Souza and Davenport [86] Studied the peak responses of transmission Peak response was mainly due to the background component
Harikrishna et al. [58] line and compared the results with different Indian Standard [68], Australian Standard/New Zealand Standard [14] agreed
codes with the field measurements while British Standard [20] overestimated the
base moment by 30%
Peil and Nolle [108] Assessed the importance of conducting Tall towers were dynamically excited
Sparling and Wegner [127] dynamic analysis on transmission towers
Lou et al. [88] Conducted numerical dynamic analyses of Natural frequency of the line was affected by the method of supporting the
Yasui et al. [143] TL subjected to turbulent wind wires. Coupled tower-conductor vibration modes were found
Battista et al. [16]
Behncke and Ho [17] Reviewed code approaches to evaluate the Span Reduction Factor and gust velocity factors presented in the codes were
peak responses suitable for synoptic wind and not for non-synoptic wind
Solari [125,126] Assessed the gust response factor developed Good agreement between the responses obtained from wind tunnel tests and
Simiu and Scanlan [121] by Davenport [32] using the gust factor expression was found
Holmes [61] Studied the responses of the line under Span reduction factor for synoptic wind cannot be used for downburst wind
Aboshosha and El Damatty [6] downbursts

an alternating sign as indicated in Fig. 1. Loredo-Souza and Daven- 2.2. Approach used by various codes
port [86] concluded that the dynamic response of transmission
towers depends strongly on the turbulence intensity and both Design codes provide a simplified approach to obtain the peak
the structural and aerodynamic damping of the towers. So far, forces acting on the elements of TLs to be used for design. In this
there is no direct expression for the aerodynamic damping of the subsection, four design codes [12,13,20,69] specialized in TLs are
towers. considered to compare between the approaches provided by these
Regarding field measurements of wind loads on lattice steel codes to estimate the forces acting on the conductors and the tow-
towers, EPRI study at the TLMRC was conducted by Shan [114] to ers. A summary of this comparison is provided in Table 2. For a bet-
collect data from a full-scale lattice transmission tower. This has ter understanding of the comparison shown in this table, the main
been done by erecting an experimental wind tower at the TLMRC terms and parameters presented in the table are first defined in
site. A number of anemometers were placed at different elevation this subsection. This is followed by Section 2.3 which provides
levels of the wind tower. Strain gages were also mounted on the leg the main steps required to apply design approaches specified in
posts of the tower. The purposes of this experiment were to estab- the four codes and the results of a detailed comparison between
lish the wind characteristics at the TLMRC site and to gain experi- these codes.
ence using different types of instrumentation and data acquisition
techniques in field-wind loading experiments. The main results 2.2.1. Design velocity Vd
that were drawn from this experiment can be summarized in the Wind velocity used in the design, Vd, of a TL represents the aver-
following points: (1) the magnitude of wind velocity and direction aged velocity over a small time interval, i.e. 3 s. This velocity can be
can vary considerably during a short period of time; (2) the mean exceeded annually with a probability of 1/R, and through the n year
vertical wind profile does not hold constant as usually assumed; life span of the TL with a probability of 1  (1  1/R)n. Inverse of
(3) the turbulence intensity and the gust factor increase as the the annual probability of exceedance, R, is commonly referred as
height above the ground decreases; (4) the averaging time can the return period of the storm. Typically, a 50 year return period
greatly influence the results of wind data analysis; (5) although is used for the design of the TLs. Design codes allow for using dif-
wind contains little energy beyond 1 Hz, structural responses ferent return periods, R, by introducing a scaling factor to the
above 1 Hz can be excited; (6) strong relationships exist between 50 year return period such as the ASCE 74 [12] and the AS/NZS
the wind velocity and the responses in the leg posts of the wind [13]. As mentioned above, reference wind velocities can be aver-
tower. The study also provided some comments on how to conduct aged over time intervals of 3 s, 1 min, 10 min or 1 h. Conversion
the field-wind loading experiments as well as how to analyze the between wind velocity averaged using different time intervals
wind and response data. can be obtained using a curve provided by Durst [43].
Table 1 summarizes the scope of previous studies focused on This curve was based on wind speeds not exceeding 40–50 mph
the behavior of TL under synoptic wind and shows their main find- (18–22 m/s) and according to Sissenwine et al. [122] and Behncke
ings. In the next subsection, the quasi-static approach utilized by and Ho [17], the velocity ratio could be different at higher wind
four different design codes to account for the dynamic effect on speeds. The ratio between the 3 s averaged velocity to the 10 min
the TLs components is discussed in detail. The procedures provided or the 1 h averaged velocity is referred as the gust velocity factor,
by these design codes to obtain design forces on the conductors Gv.
and the towers are compared in order to assess the discrepancy
between different codes’ approaches. The main results of this com- 2.2.2. Design wind pressure qd
parison are utilized in the last section (Section 4) to identify the Design wind pressure, or sometimes is referred to as the peak
gaps noticed in the four codes considered in the current study wind pressure, qp, represents the dynamic pressure caused by the
and to provide possible recommendations to fill out these gaps. design wind velocity Vd, and can be calculated using Eq. (2).
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 27

1
qp ðzÞ ¼ qV 2d ðzÞ ð2Þ FC
2 V3s Incoming Velocity
Conductor
where q is the air density.
In general, the air density q depends on the temperature and Conductor Tower
the atmospheric pressure. It is equal to 1.225 kg/m3 at 15 °C at
an atmospheric pressure 101.325 kPa. For other temperature and Fig. 2. Layout of the conductor and the applied force Fc (modified from the IEC
pressures, the expression described by Eq. (3) can be used. [69]).

q 288 1:2104 H
¼ 0 e ð3Þ 2.3. Quantitative comparison between codes
q0 T

where q0 is the density at 0 elevation H at 15 °C (288 K) and equals This comparison focuses on assessing how different codes esti-
to 1.225 kg/m3, T0 is the temperature in K, H: the elevation above mate the main parameters affecting the design of TLs under synop-
the ground. tic wind loads. These parameters include: design wind velocity and
pressure, conductor gust response factor, forces transferred from
conductors to the tower and forces acting directly on towers. For
2.2.3. Conductor force Fc each quantity, a code ratio (CR) is obtained by normalizing the val-
Conductor force Fc represents an overall wind force transmitted ues predicted from the three codes [13,20,69] by the value
from the conductor to a tower. This force is usually in the trans- obtained from ASCE 74 [12]. The details of this comparison are
verse direction of the conductor regardless of the wind direction given below.
as shown in Fig. 2.

2.3.1. Design wind velocity and pressure


2.2.4. Conductor gust response factor GC
In general, the four codes use the wind velocity at 10 m eleva-
Conductor gust response factor, GC, is used to scale the conduc-
tion above the ground for open terrain exposure as the reference
tor forces resulting from the design pressure (based on 3 s gust
velocity. The ASCE 74 [12] and the AS/NZS [13] use a 3 s averaged
speed) to account for the dynamic response of the conductor.
wind velocity for the reference wind speed, V3sOpZ10. The BS [20]
Accounting for the dynamic effect using the gust factor was origi-
and the IEC [69] use a 10 min averaged wind velocity, V10mOpZ10.
nally proposed by Davenport [32,33] and examined by many
The codes allow for converting the velocity of the open terrain
researchers such as Solari [125,126], Simiu and Scanlan [121] and
exposure which is measured at 10 m height to obtain the velocities
introduced in a new format by Zhou and Kareem [146]. The origi-
for other exposures and heights, V(z). A comparison between the
nal version of the gust factor proposed by Davenport [33] was
expressions employed by various codes for the velocity V(z) is pro-
based on a 1 h-averaged wind speed. This means that it combines
vided in Table 2. In order to allow for a direct comparison, velocity
the gust velocity, Gv, and the gust response factor. In general, the
expressions shown in the table are described as functions of the 3 s
gust response factor for a structure accounts for the background
gust wind speed for the open terrain exposure at 10 m height
and the resonant components of the responses. Fortunately, con-
V3sOPZ10. This is straight forward for the ASCE 74 [12] and the AS/
ductors have a high aerodynamic damping at the range of the wind
NZS [13] as their reference velocities are the 3 s averaged. For
speeds used in the design, which suppresses most of the resonant
the BS(2001) and the IEC [69], the conversion from the 10 min
component, and thus allows the design codes to neglect the reso-
averaged to the 3 s averaged is conducted using the formulas
nant component. It shall be mentioned that assuming a minor res-
provided in the two codes, where a gust velocity factor Gv is intro-
onant component might not be true for lower wind speeds causing
duced, as indicated in Table 2.
smaller aerodynamic damping. Such a small aerodynamic damping
Design wind velocity, Vd(z) or V3s(z), is defined in the ASCE 74
may not be enough to suppress the resonant motion and conse-
[12] using the power law and is defined in the BS [20] and the
quently may cause fatigue problems to the cross arms.
IEC [69] using the logarithmic law. The AS/NZS [13] provides a
table to describe the velocity variation, which is fitted in the cur-
2.2.5. Tower force FT rent study using the power law as indicated in Table 2. Design
FT represents the total force acting on a tower panel in the direc- pressure corresponding to the peak wind speed, qp, described by
tion of the wind. This force is a function of the design pressure, qp. the four codes is also compared in the table.
It is also a function of the drag coefficients, Cdx and Cdy, and the Fig. 3 illustrates the design pressure qpn(z) at different eleva-
panel projected areas, Ax and Ay, in the panel x and y directions, tions for the sea side, open and sub urban terrain exposures nor-
respectively. Drag coefficients of the panel, Cdx and Cdy, are typi- malized by the design pressure qpOpZ10 at 10 m for open terrain
cally a function of the panel porosities /x and /y, in the x and y exposure as expressed by Eq. (4).
directions, respectively.
qp ðzÞ
qpn ðzÞ ¼ ð4Þ
qPoPZ10
2.2.6. Tower gust response factor GT
Similar to the conductor gust response factor Gc, the tower gust As shown in the figure and indicated in Table A.1, at a height of
response factor, GT, is used to scale the forces acting on the tower 30 m for instance, design pressure obtained by the four different
panels resulting from the design pressure, qp, to account for the codes are different. It is noticed from the results that the code ratio
dynamic response of the tower. As discussed earlier and indicated CRp ranges between 0.91 and 1.13. The minimum value of CRp is
by Holmes [61], typical towers in the order of 60 m or less that noticed in case of using AS/NZS [13] and IEC [69] approaches for
have a fundamental frequency in the order of 1 Hz or more have suburban and sea side exposures, respectively. The maximum
a minor resonant component. That is because most of the wind value of CRp is reached by using BS [19] approach for Suburban
fluctuating energy is associated to small frequencies less than exposure. These discrepancies are due to the slight differences in
1 Hz. Based on this, design codes normally neglect the resonant the expressions and parameters used to describe the velocity pro-
component of the towers. files and consequently the pressure in each code.
28 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

Sea side Open Sub urban


100 100 100
ASCE
AS/NZS
90 90 90
BS
IEC
80 80 80

70 70 70

60 60 60
Z (m)

Z (m)

Z (m)
50 50 50

40 40 40

30 30 30

20 20 20

10 10 10

0 0 0
0.5 1 1.5 2 0.5 1 1.5 2 0.5 1 1.5 2
P (z) P (z) P (z)
pn pn pn

Fig. 3. Normalized peak wind pressure Ppn(z) as defined by various codes.

2.3.2. Conductor forces and the gust factor GC the expressions used for the Gv or the SRF do not depend on the
Expressions used in the codes for the force FC transmitted from height or the exposure. Only the ASCE 74 [12] allows for changing
the conductor to the towers are summarized in Table 1. As indi- the Gv or the SRF with varying the conductor height and the terrain
cated from the table, the expressions are similar in the considered exposure as indicated in Table 2.
four codes. The force depends on the angle h, which is the angle As mentioned previously, the ASCE 74 [12] code uses a 3-s gust
between the incoming wind and the perpendicular to the line as velocity as the reference wind speed, however, when the gust
indicated in Fig. 2. The term GC is called the conductor gust response of the conductor is calculated, the code uses a formula-
response factor, which is sometimes called the Span Reduction Fac- tion that is based on 10 min-averaged speed. The code introduces
tor (SRF) when used as a multiplier for the peak design pressure qp. a conversion factor Kv based on Durst curve. Other codes such as
The Gc or the SRF accounts for the lack in the correlation due to the BS [20] and the IEC [69] use different conversion approaches.
the increase of the conductor span S. As indicated from Table 2, This step in particular leads to differences in the order of 10–15%
the AS/NZS, the BS and the IEC use a SRF that only depends on the [17].
span length. According to Davenport [35], this SRF can be expressed It should be also mentioned that out of the four studied codes,
by Eq. (5), where the term JLs is called the joint acceptance function AS/NZS [13] is the only code which proposed a different expression
which accounts for the lack of the correlation along the span S and for the conductor gust factor under downburst wind. Following the
is expressed by Eq. (6). work of Holmes [61], which recognized a significant difference
pffiffiffiffiffi between synoptic and downburst wind loading, AS/NZS [13] pro-
1 þ 2IðzÞg f J Ls posed a different expression GCDb for the gust factor as indicated
SRF ¼ Im ð5Þ
1 þ 2IðzÞg f in Table 2.
A comparison between the normalized force Fcn transmitted
Z 1 Z 1 from the conductors to a typical tower, defined by Equation, is
Lx jn1n2j
J Ls ¼ u2v ðn1 Þu2v ðn2 ÞiFc ðn1 ÞiFc ðn2 Þeð Ls
Þ
dn1 dn2 ð6Þ shown in Fig. 5. The comparison is conducted for two conductors
1 1 that are placed at 20 and 40 m elevations for a span S ranging from
200 to 800 m. As indicated from the figure, and summarized in
where I(z) is the turbulent intensity at the conductor elevation z; gf:
Tables A.2 and A.3 the normalized forces FCn(20) and FCn(40) calcu-
velocity peak factor; n1 and n2 are local axis which equals to 1, 0, 1
lated by the four codes generally increases with the conductor
at the left tower, the tower of interest, the right tower as indicated
height and decreases with the increase of the conductor span.
in Fig. 4; iFc: Influence line for the overall conductor force transmit-
ted to the tower iFc = 1 + n, where n = 1:0 and iFc = 1  n, where qp
Fc
n = 0:1; Ls: transverse length scale of the turbulence; Im : is an inte- F cn ðzÞ ¼ 2
¼ SRF ð7Þ
qpOpZ10 SD cos ðhÞ qpOpZ10
gral that depends on the distribution of the mean forces along the
R1
spans and the influence line Im ¼ 1 iFc ðn1 Þu2v ðn1 Þdn1 ; uv is the The tables also show that the discrepancy in the code ratio for
velocity distribution function which equals to V(n)/V0, where V0 is conductor forces CRFc(20) for a conductor located 20 m above the
the velocity at any reference point. For the case of horizontal con- ground has a minimum value of 1.02, 0.99, 0.83 and a maximum
ductors uv (n) = 1. value of 1.4, 1.3, 1.42 for the sea side, open and suburban expo-
As indicated from Eqs. (5) and (6), SRF depends on the turbulent sures, respectively. For a conductor located at 40 m above the
intensity I(z), which vary with height, and on the turbulent length ground, it is noticed that CRFc(40) has a minimum value of 1.01,
scale Ls, which vary with the terrain exposure. Consequently, the 1.04, 0.81 and a maximum value of 1.44, 1.29, 1.31 for the sea side,
SRF is expected to vary with the height and the exposure type. This open and suburban exposures, respectively. In addition, the com-
is not the case for the AS/NZS [13] the BS [20] and the IEC [69], as parison shows that the minimum values of the code ratio CRFc
Table 2
Comparison between various codes.

Term Def. ASCE [12] AS/NZS [13] BS [20] IEC [69]


Vref Reference velocity V3sOpZ10 V3sOpZ10 V10mOpZ10 V10mOpZ10
 a1 pffiffiffiffiffi  z a=2 G K G
Vd(z) or V3s(z) (m/s) Design velocity
V 3s ðzÞ ¼ 1:418V 3sOpZ10 Zh V 3s ðzÞ ¼ b V 3sOpZ10 10m V 3s ðzÞ ¼ kt lnðz=z0 Þ 1:435
v V
3sOpZ10
r v
V 3s ðzÞ ¼ 1:355 V 3sOpZ10
Zg
V 3s ðzÞ ¼ G
v kt lnðz=z0 ÞV 10mOpZ10 V 3s ðzÞ ¼ G logðz=z0 Þ

v logð10=z0 Þ V 10mOpZ10
V 3s ðzÞ ¼ Gv V 10m ðzÞ V 3s ðzÞ ¼ G
v V 10m ðzÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
G
v ¼ 1 þ lnðz=z0 Þ
2:28
G
v ¼ a lnðzÞ þ b
qp(z) (N/m2) Design wind pressure  2a   z a    
qp ðzÞ ¼ 0:60V 23sOpZ10 b V 3sOpz10 2
qp ðzÞ ¼ 0:613 1:435 lnðz=z0 Þkt K 2g qp ðzÞ ¼ 0:613 1:355
V 3sOpZ10 2 2 
qp ðzÞ ¼ 1:232V 23sOpZ10 Zh
Zg
10m K r Gv
FC Conductor force qp ðzÞGc Sd cos ðhÞ2 qp ðzÞGc Sd cos ðhÞ2 qp ðzÞGc Sd cos ðhÞ2 qp ðzÞG
C Sd cos ðhÞ
2

H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46


pffiffiffi a1 pffiffiffiffi GC ¼ 1:3  a lnðSÞ G
c ¼ 1:575  0:106 lnðSÞ
ðS=210mÞ
GC Conductor gust 1þ13:23 jð10=Z h Þ FM BC GCNW ¼ 0:59 þ 0:41e
GC ¼
response factor K2 GCDb ¼ 1:0  0:3125 S200
1000 6 1:0
v
1
BC ¼ ð1 þ 0:8S=Ls Þ
FT Tower force F ¼ qp ðzÞGT ðC Dx Ax cosðhÞ2 þ C Dy Ay sin ðhÞ2 Þ F ¼ qp ðzÞGT ðC Dx Ax cos ðhÞ2 kh Þ þ ðC Dy Ay sin ðhÞ2 kh Þ F ¼ qp ðzÞGT kh ðC Dx Ax cos ðhÞ2 þ C Dy Ay sin ðhÞ2 Þ F ¼ qp ðzÞG
T kh ðC Dx Ax cos ðhÞ
2

þC Dy Ay sin ðhÞ2 Þ
GT Tower gust response pffiffiffi a1 pffiffiffiffi GT ¼ 1:0 GT ¼ 1:05 2

GT ¼
1þ13:23 jð10=Z h Þ FM BT G
T ¼
dz þezþf
G2
factor K 2v v
1
BT ¼ ð1 þ 0:373h=Ls Þ
8 9
CDx,y Drag coefficient for the >
> 4:0 / < 0:025 >
> C Dx;y ¼ 4:1/2  6:0/ þ 4:0 C Dx;y ¼ 4:1/2  6:0/ þ 4:0 C Dx;y ¼ 4:1/2  6:0/ þ 4:0
< =
towers 4:1  5:2/ 0:025 6 / < 0:44
C Dx;y ¼ kh ¼ 1 þ 0:55 sin ð2hÞ2 k2 kh ¼ 1 þ 0:2 sin ð2hÞ2 kh ¼ 1 þ 0:2 sin ð2hÞ2
>
> 1:80 0:045 < / < 0:69 >
> 8 9
: ; > 0:2 / 6 0:2 >
1:3 þ 0:7/ 0:7 < / < 1:0 >
< >
=
/ 0:2 < / 6 0:5
k2 ¼
>
> 1  / 0:5 < / 6 0:8 >
>
: ;
0:2 0:8 < /
Par. Different parameters Zg = 213, 274, 366 m b = 1.115, 1.000, 0.856, 0.669 z0 = 0.01, 0.05, 0.30, 1.0 m Kr = 1.08, 1.00, 0.85, 0.67
used in the code a = 11.5, 9.5, 7.0 a = 0.064, 0.130, 0.090, 0.185 kt = 0.17, 0.19, 0.22, 0.24 a = 0.291, 0.373, 0.494, 0.615
aFM = 10.0, 7.0, 4.5 for sea side, open, sub urban and urban terrains a = 0.073, 0.082, 0.098, 0.110 b = 1.05, 0.976, 0.912, 0.814
j = 103  3, 5, 10 for sea side, open, sub urban and urban d = 104  2.0, 2.0, 2.0, 2.0,
Ls = 51.8, 67.0, 76.2 m terrains, respectively e = 102  2.32, 2.74, 2.98,
for sea side, open and urban terrains, 3.84
respectively f = 1.47, 1.68, 2.27, 2.92
Kv = 1.43, and is used to z0 = 0.01, 0.05, 0.3, 1.0 m
for sea side, open and urban
terrains, respectively

V3sOpZ10: 3 Second gust velocity measured in open terrain at elevation z of 10 m; V10mOpZ10: 10 min averaged velocity measured in open terrain at elevation z of 10 m; subscript Op refers to an open terrain; qr: reference pressure; qp:
peak pressure; qm: mean pressure; FC: conductor force perpendicular to the line; FT: Tower force in the wind direction; GC: Gust factor for the conductor; GC3sEq: equivalent gust factor for the conductor in the IEC [69] code if the
peak pressure is used as the reference pressure; GT: Gust factor for the tower; GT3sEQ: Equivalent gust factor for the tower if the peak 3 s gust pressure is used; The superscript ⁄ indicates that the function is given in the standard in a
graph and is fitted here using the presented expression; The super script ⁄⁄ indicates the name of the these factors are given different name in the codes and the name assigned in this study is used for comparison purpose.
Subscripts NW refers to normal winds and Db refers to downburst wind.

29
30 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

due to the lack of correlation while increasing the tower height, as


shown in Fig. 6. The code increases the gust factor GT with the
increase of the terrain roughness, which is the same trend for the
IEC [69] code. It should be mentioned that the tower gust factor
n1, n2
n1, n2=-1 n1, n2=0 n1, n2=1 GT can be calculated using similar expressions to those in Eqs. (5)
and (6) while using the influence line of the base tower moment,
iBM, or the base tower shear iBS. According to these equations, rough
terrain exposures having large length scales Ls, would result in
higher gust factor GT than the smooth terrains, which agrees with
Fig. 4. Parameters required by the Joint Acceptance Function (JAF).
the trend predicted by the ASCE 74 [12] and the IEC [69] codes.
Similar to the conductor gust factor, the ASCE 74 [12] intro-
duces a conversion factor Kv to convert the gust factor formulation,
are noticed in case of AS/NZS [13] and the maximum values are which is based on 10-min averaged velocity, to be used with a 3-s
obtained from BS [20] and IEC [69]. This is related to the discrep- reference velocity. This conversion leads to an average error of 10–
ancy between the peak pressure qp(z) and between the expressions 15%.
used for the GC (i.e. SRF) recommended by the four codes. It is By comparing the expressions used by various codes to obtain
noticed that the least force prediction is found by the ASCE 74 the tower forces FT, it appears that the ASCE 74 [12] is missing a
[12] code for the sea side and the open terrain exposures and by factor called Kh used to account for the increase of the projected
the AS/NZS [13] for the suburban exposure. It is also noticed that area subjected to the wind with the increase of the pitch angle h
the maximum force for the sea side and open exposures is esti- up to 45°. This agrees with the findings of the wind tunnel testing
mated by the BS [20] and for the suburban exposure by IEC [69]. conducted by Mara et al. [93]. Normalized tower force, FTn, defined
by Equation is calculated for two different angles of attack and
compared as shown in Fig. 7 and summarized in Tables A.4 and
2.3.3. Tower forces A.5. The comparison is conducted considering a square tower with
Expressions for the force due to the design pressure qp(z) a height of 25 m that has equal projected area A and solidity ratio /
applied to a tower panel, Ft, are summarized in Table 2. This force in both x and y directions. Two wind angle of attacks, 0° and 45°,
acts in the direction of the wind load. As indicated from the expres- are considered in the comparison. Tables A.4 and A.5 summarize
sions presented in the table, design codes use the gust response the normalized tower forces FTn(0°) and FTn(45°) estimated using
factor for the tower GT to scale the design pressure qp(z). The AS/ the four codes. For the forces acting on tower panels in case of 0°
NZS [13] uses a tower gust factor, GT, equal to 1.0 and the BS angle of attack, the discrepancy in the code ratio for tower forces
[20] uses a factor of 1.05, which means that the two codes do CRFt(0°) shows a minimum value of 1.08, 1.05, 0.68 and a maxi-
not in general allow for reducing the loads resulting from the lack mum value of 1.53, 1.29, 1.27 for sea side, open and suburban ter-
of correlation while increasing the tower height. On the contrary, rain exposures, respectively. These are compared to CRFt(45°), in
the ASCE 74 [12] allows for reducing the load on the tower panels case of 45° angle of attack, which has a minimum value of 1.30,

Sea side z=20m Sea side z=40m


1.5 1.5

1 1
Cn

Cn
F

0.5 0.5

0 0
200 300 400 500 600 700 800 200 300 400 500 600 700 800
S (m) S (m)

Open z=20m Open z=40m


1.5 1.5

1 1
Cn

Cn
F

0.5 0.5

0 0
200 300 400 500 600 700 800 200 300 400 500 600 700 800
S (m) S (m)

Suburban z=20m Suburban z=40m


1.5 1.5

1 1
Cn

Cn
F

0.5 0.5

0 0
200 300 400 500 600 700 800 200 300 400 500 600 700 800
S (m) S (m)
ASCE AS/NZS BS IEC

Fig. 5. Conductor normalized force FCn.


H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 31

Sea side Open Sub urban


60 60 60

55 55 55

50 50 50

45 45 45

Z (m)
Z (m)

Z (m)
40 40 40

35 35
35

30 30
30

25
25 25

20
20 20 0.8 0.9 1 1.1 1.2 1.3
0.7 0.8 0.9 1 1.1 0.7 0.8 0.9 1 1.1
GT GT
GT

ASCE AS/NZS BS IEC

Fig. 6. Gust tower response factors GT used in various codes.

Sea side θ=0o Sea side θ=45o


6 6

4 4
tn

tn
F

2 2

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
φ φ

Open θ=0o Open θ=45o


6 6

4 4
tn

tn
F

2 2

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
φ φ

Suburban θ=0o Suburban θ=45o


6 6

4 4
tn

tn
F

2 2

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
φ φ
ASCE AS/NZS BS IEC

Fig. 7. Normalized tower force Ftn.

1.16, 0.82 and a maximum value of 1.85, 1.56, 1.53 for the same results is due to the difference in the design pressure in addition
type of exposure. As indicated from the reported values, it appears to the difference between the expressions used for the Kh and for
clearly that angle of attack 45° lead to higher values of CRFt. This is the drag coefficient Cd.
because ASCE 74 [12], used for normalization, does not take into
account the effect of change in the angle of attack (i.e. Kh = 1.0). FT
F Tn ¼ ð8Þ
Fig. 8 shows the variation of the forces with the change of the wind qpOpZ10 A
angle of attack, h. These figures are plotted considering a typical
solidity ratio of 20%. As inferred from the Kh equation, presented
in Table 2, tower forces increase with the angle h until a value of 3. Response of TLs under downburst wind loading
45° then tend to decrease. It is found that the IEC [69] code predicts
higher forces for the open and suburban exposures than those pre- A downburst is an intensive downdraft air that induces very
dicted by other codes. Maximum force for sea side terrain exposure strong wind in all directions when striking the ground. Fujita
is predicted by the BS [20] code. The discrepancy between the [48] defined a downburst as a mass of cold and moist air that drops
32 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

Sea side φ =0.20


6

4
tn
F

0
0 10 20 30 40 50 60 70 80 90
θ

Open φ =0.20
6

4
tn
F

0
0 10 20 30 40 50 60 70 80 90
θ

Suburban φ =0.20
6

4
tn
F

0
0 10 20 30 40 50 60 70 80 90
θ
ASCE AS/NZS BS IEC

Fig. 8. Variation of the FTn with the wind angle of attack h.

radar data from the JAWS project and reported the horizontal
and vertical profiles of the microburst. Choi [28] reported measure-
ments from wind and rain stations at Singapore using 20 m mast.
The study reported that gust speeds mostly take place during thun-
derstorms. The author compared the characteristics of the thun-
derstorm and non-thunderstorm winds and reported that the
turbulent effect and gust factors are higher in the first case. Later
on, Choi [28] carried out field measurements and experimental
thunderstorm simulations to investigate the wind profile with
height during thunderstorms at Singapore. The field measurements
were conducted considering five levels of measurements along a
150 m tower height while the experimental work adopted the
impinging jet methodology. The study emphasized the effect of
the distance from the thunderstorm center, intensity of the storm,
and the ground roughness. Orwig and Schroeder [105] compared
the metrological data of a downburst thunderstorm to those of
synoptic winds in a research project called 2002 Thunderstorm
Outflow Experiment. In their experiment, a linear array seven tow-
Fig. 9. Downburst event. Source: http://lacasanaranja.wordpress.co. ers were utilized to record two high intensity wind events, a rear-
flank downdraft of a supercell, and a derecho. The analysis showed
significant differences with those events of non-thunderstorm
suddenly from the thunderstorm cloud base, impinges on the winds. Gunter and Schroeder [53] utilized two high-resolution
ground surface, and then horizontally diverges from the center of mobile Doppler radars to measure full-scale data of three thunder-
impact as shown in Fig. 9. In basic terms, downbursts are down- storm events in a project named SCOUT. The unique technology of
drafts with sufficient energy to reach ground level. their experiments allowed analyzing the vertical profile of the
flow. The measurements showed different behavior between each
3.1. Field measurements and numerical modeling of downburst of the measured events in addition to the instantaneous dual Dop-
pler wind speed and direction of each of individual event. The anal-
3.1.1. Field measurements during downbursts ysis showed that the location of the maximum speed of each event
As mentioned in Section 1, wind field prediction for downbursts was higher in elevation than those of numerical simulations and
is a major challenge. A limited number of field measurements for other metrological measurements. Lombardo et al. [87] analyzed
downbursts are available in the literature. These include the North- the archived data obtained Automated Surface Observing System
ern Illinois Meteorological Research (NIMROD) and the Joint Air- ASOS to identify a number of downburst thunderstorms and
port Weather Studies (JAWS) reported by Fujita [49], and the compared them to ABL events. The results showed that a shorter
FAA/Lincoln Laboratory Operational Weather Studies (FLOWS) averaging times (15–60 s) can be used for downbursts compared
reported by Wolfson et al. [140]. The initial diameter (DJ) of the to 1–5 min for ABL winds. In addition, the study revealed that gust
downdraft reportedly ranges between 600 and 1700 m as provided factors correspond to those thunderstorm events differ between
by Hjelmfelt [60]. Wilson et al. [138] used the Doppler weather thunderstorm types. Based on the date recorded at different ports
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 33

(a) Vortex ring model (b) Impinging jet model

Ellipsoidal Cooling
Function

(c) Cooling source model


Fig. 10. (a) Vortex ring model [111], (b) impinging jet model [111,73] and (c) cooling source model [129].

in Europe such as Genova, Savona, La Spezia, Livorno, and Bastia, Computational Fluid Dynamics (CFD) model and validated their
De Gaetano et al. [39] analyzed the set of data recorded using CFD assumptions by carrying out pressure and velocity measure-
through a semi- automated procedure to separate different events. ments with an impinging jet. Kim and Hangan [73] compared their
The authors stated that separating the synoptic and non-synoptic simulated mean wind profile with full-scale data [141], experi-
events is a hard task due to the presence of third-class events that mental results [41,42], as well as a generic 1/7 power law bound-
has intermediate properties between the main two classifications ary layer profile. Kim et al. [74] reported that downbursts can
of winds. Solari et al. [123] reported the thunderstorm records that produce different loading for tall structures and, consequently, dif-
have been recorded through ‘‘Wind and Ports” project. Using an in- ferent collapse modes. These CFD models of an impinging jet
situ wind monitoring network, they analyzed the main properties yielded time series for a vertical (axial) component (VVR) of the
of a number of thunderstorm records detected in the Ports of velocity field, as well as a radial (horizontal) component (VRD). At
Genoa, La Spezia and Livorno during the period of 2011 to 2012. a fixed point in space, these two velocity components are functions
Solari et al. [123] reported the mean values and the coefficient of of horizontal location relative to the center of the jet and height
variation of three wind velocity ratios that are believed to have a relative to the ground.
significant effect on structures. Sengupta and Sarkar [113] simulated downbursts using the
The variable nature of downbursts, both spatially and in time, impinging jet method employing K-epsilon, K-omega, Shear Stress
makes full-scale study a difficult task. Transport (SST) and Large Eddy Simulation (LES) turbulence mod-
els and compared the resulting profiles with those from an exper-
3.1.2. Downburst numerical, analytical and empirical modeling iment. Their results showed a reasonable agreement between the
Due to the difficulty of field measurements for HIW events such profiles obtained from the LES and from the experiment. The appli-
as downbursts, most of the research has relied on numerical mod- cability of using LES to simulate downbursts is also indicated from
eling or reduced-scale physical modeling (i.e. experiments) to pre- the results of Hadžiabdić [55], Chay et al. [22], Gant [51], and
dict the downburst wind field as well as the response of structures Aboshosha et al. [8]. Later, Zhang et al. [145] studied the down-
to such events. Three numerical approaches in the literature are as burst simulation using, experimentally, the steady impinging jet
follows: (a) Ring Vortex Model, (b) Impinging Jet (Impulsive Jet) model and, numerically, the cooling source model. The PIV mea-
Model, and (c) Cooling Source (Buoyancy-Driven) Model, Fig. 10. surement of the impinging jet model showed a good match with
The Ring Vortex Model [147,71,132,111] simulates the vortex ring the average wind velocity obtained previously by field measure-
that is formed during the descent of the downdraft air column. As ments. However, the study showed inconsistencies in the transient
reported by Savory et al. [111], the Ring Vortex Model is not accu- features of each model. The study provided a set of advantages and
rate in simulating the downburst field near the ground after the air disadvantages for each model.
column touches the ground. Mason et al. [95] implemented the cooling source method based
The impinging jet model is based on the analogy between an on a dry, non-hydrostatic, sub-cloud and axisymmetric model. One
impulsive jet impinging upon a flat surface and a downburst Fujita year later, Mason et al. [96] extended this work to a three-
[48]. The cooling source method is based on introducing a cooling dimensional model. In both papers, the Scale Adaptive Simulation
source inside the computational domain simulating the cooling (SAS) method was used, which is an improvement for the unsteady
process in the cloud base. This cooling process increases the weight Reynolds-Averaged Navier–Stokes (URANS) method used to pre-
of the cloud base and formulates the downdraft. Selvam and dict unsteady turbulent flow. However, Gant [51] reported that
Holmes [112] developed a two-dimensional, steady numerical the SAS method appears to be over-predicting the turbulent viscos-
model for an impinging jet. Hangan et al. [57] improved this basic ity of jet-type flows. Vermeire et al. [129] simulated downbursts
34 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

Table 3 z0, summarized in Table 3 for different terrain exposures. Accord-


Aerodynamic roughness for different terrain exposures [137]. ing to Blocken et al. [18], the maximum roughness z0 that can be
Terrain exposure z0 (m) accurately modeled by a wall function is governed by Dz > 60 z0,
Smooth 0.005 where Dz is the height of the first grid layer. Mason et al. [95]
Open 0.03 and Vermeire et al. [129] used a first grid layer with a height Dz
Roughly open (suburban) 0.25 equal to 1.0 m, which shades doubts on their results for the terrain
Very rough (forest or urban) 1.00 roughness greater than 0.016 m.
The aforementioned structural studies of downburst loading
provided a good base of knowledge about mean flow characteris-
using the cooling source approach with Large Eddy Simulation tics. However, the characteristics of the turbulence (e.g. turbulence
(LES) and the results showed good agreement with Mason et al. intensities or length scales) associated with downbursts are yet to
[95] and disagreement with impinging jet models. be fully clarified. Recently, Aboshosha et al. [7] characterized the
Most studies have considered a single downburst column turbulent component of downburst wind field for 4 terrain expo-
whereas in a study that modeled two downbursts, representing sure conditions employing LES. Those terrain exposure were simu-
part of a downburst line, which commonly occur in nature. Ver- lated using fractal surfaces employing the modified roughness
meire [130] showed that two adjacent downbursts may result in gradient drag-based model [7], which does not have a limitation
peak near-ground wind speeds that are 55% greater than those on the maximum roughness z0 that can be used.
associated with one of those downbursts on its own, together with Many research attempts were done to simulate the downburst
a 70% greater damage footprint. However, even cooling source wind field analytically. Oseguera and Bowles [106] and Vicroy
models neglect some of the other key physics of real events, nota- [131] developed an analytical model that counted for the mean
bly the forcing associated with precipitation, and the fact that such wind speed of stationary downbursts. The main drawback of their
events are not axi-symmetric. Recently, LES simulation of complete models was that it eliminates the translation velocity. The model
downburst-producing thunderstorms [103] have revealed the also ignored the nonlinear variation of the downburst boundary
complexity of downburst outflows, although circumferential aver- layer thickness. Later, Chay et al. [22] improved the model pro-
aging of the wind field suggests that it may be possible to estimate vided by Vicroy [131] by modifying several factors such as the
the highest wind speeds at any time within the event by using the equation of the intensity-scaling factor so that the model could
averaged wind field (essentially an axi-symmetric metric) multi- predict the maximum out flow speed. They also considered the
plied by a consistent peak factor [104]. effect of the distance from the downburst center on the maximum
Kim and Hangan [73] studied different field and numerical wind speed. The model also took into account the effect of the
measurements of downbursts and were able to conclude a con- translation speed. However, Chay et al. [22] they did not consider
stant estimate of 6 m/s for the translation velocity of the event the continuity equation which controls the relation between the
which was then included in their impinging jet simulation. Lin vertical and the horizontal speed. In addition, their study did not
et al. [82] conducted a physical and numerical comparisons provide an analytical solution for the downburst turbulent compo-
between the translating (slot jet) and non-translating (steady, nent. Abd-Elaal et al. [1] continued working on the previous mod-
round impinging jet) simulations of downbursts. The study els by incorporating varying characteristic lengths into the velocity
revealed that the non-translating downburst underestimates shaping functions and added external shaping functions for inten-
velocity profiles, on the other hand, the slot jet velocity profile sity scalars. In Their model, Abd-Elaal et al. [1] were able to account
tends to exceed the observations. Although both simulations ade- for both the radial and the vertical profiles of the downbursts tak-
quately represents profile shape and height, one of the differences ing into consideration the nonlinear effects of the downburst
between them is that the translating cooling source have shown boundary layer nonlinear development.
peak velocities at noticeably lower heights than the other profiles. Different empirical simulations of the downburst wind filed are
Design codes define the wind speed at any point as being the found in the literature. Holmes and Oliver [62] developed an
sum of a mean wind component (non-turbulent) and a turbulent empirical model of the horizontal wind component of the down-
component. Most studies have considered only the non-turbulent burst. The model was developed using the basic knowledge of
component of downburst wind velocity. The turbulence character- the impinging jet model. The study provided two main equations
istics are essential for quantifying the peak loads and responses for the radial velocity and the corresponding storm translation
experienced by surface structures as indicated by Kwon and Kar- speed. The model was then validated using field measurement of
eem [77], Chen and Letchford [25,26], Chay et al. [22,23] and the downburst recorded at Andrews Air Force Base in 1983. Wood
Holmes et al. [63]. An innovative approach is introduced by Solari et al. [141] proposed an empirical equation to predict the normal-
et al. [124] which treated a SDOF structure subjected to thunder- ized mean velocity profile using wind tunnel experimental data of
storm similar to response the spectrum approach usually used a downburst. Li et al. [80] developed an empirical 3D axisymmetric
for structural seismic design. In their approach, a thunderstorm steady state wind filed of downburst using the impinging jet basis.
response spectrum is defined to allow the calculation of the max- The model considered the nonlinear effects of the downburst pro-
imum structural response. Similar to the previously denoted stud- file by incorporating varying characteristic lengths into the velocity
ies, the study emphasized the importance of considering the peak shaping functions. In their study, a vertical shaping function for the
records of thunderstorms. radial component of the downburst was developed and showed
Recent downburst simulations have considered the effects of good agreement with CFD results. However, Li et al. [80] reported
terrain roughness [95,96,129] and topography [97] on downburst that the nonlinear effects could not be incorporated into the verti-
and synoptic winds. Terrain roughness in these studies was mod- cal velocity model. Both studies accounted only for the steady state
eled using neutral wall functions. Wall functions are typically used flows of the downbursts. Abd-Elaal et al. [2] developed an empiri-
to economically simulate the wall bounded flow using few number cal model that is able to predict the downburst turbulent compo-
of grid layers. Without using a wall function, simulation of wall- nent using several recorded field measurements of this event.
bounded flows, such as downbursts, will be computationally very The frequency features of the downburst is the main factor that
expensive due to the required large number of grid layers near affects the structural response. Wang et al. [134] utilized a hybrid
the ground to model the rapidly varying velocity gradient. Terrain time–frequency domain method using the stationary wavelet and
roughness is typically characterized by the aerodynamic roughness Hilbert transforms to model the instantaneous features of the
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 35

Velocity (m/s)

Time (sec)

Fig. 11. Contribution of the turbulent component to the overall wind velocity.

Tower one is the outflow of a rear-flank downdraft and another one is


Transmission Line Derecho as reported by Gast and Schroeder [50] to evaluate differ-
ent downburst simulation techniques such as the moving average,
discrete wavelet transform and empirical mode decomposition.
R The study recommended an empirical equation to calculate the
appropriate averaging window size. The study recommended a
carful choice for the window size it may differ for different down-
bursts. The analysis also showed that the discrete wavelet trans-
Downburst form as well as the empirical method had a good performance in
estimating the varying mean wind speed.
DJ Fig. 11 shows a sample mean and turbulent wind velocity
record obtained from CFD simulation conducted by Aboshosha
et al. [7]. The figure indicates the contribution of the turbulent
component to the overall downburst wind velocity. An envelope
Fig. 12. Downburst parameters with respect to the tower of interest. of the peak response can be approximated by scaling up the mean
component alone. Most quasi-static analyses of TLs have relied on
this scaled mean approach.
downburst accurately. Orf et al. [102] conducted high-resolution Turbulence intensity quantifies the magnitude of wind velocity
simulation of a downburst using three-dimensional cloud model. fluctuations relative to a mean wind velocity. Turbulence length
The near ground outflow was analyzed using the circumferential scales give indications of the size of the flow structures that affect
spatial averaging of the wind speed components. The analysis the span reduction factor and the gust factor [63]. The turbulence
showed that the general location of the maximum wind speed associated with downbursts plays an important role when quanti-
was preserved; however, the averaging did not maintain the out- fying loading for the design of transmission lines. For the case of
flow structure, specifically the main horizontal vortex. Su et al. synoptic wind loading, design codes [12,13] apply a gust load fac-
[128] utilized the two full-scale downburst records, which tor to account for the reduction of the turbulence correlation with
recorded near Reese Technology Center, Lubbock, Texas, where space, the terrain type and the dynamic response of the structure.

250 250

200 200

R/Dj=1.00
Height (m)

150
Height (m)

150
R/Dj=1.00 R/Dj=1.10
100 R/Dj=1.10 100 R/Dj=1.20
R/Dj=1.20 R/Dj=1.30
50 R/Dj=1.30 50

0 0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
Normalized Vertical Velocity V VRD/Vj Normalized Vertical Velocity V VRD/Vj

Fig. 13. Vertical profile of the horizontal/radial outflow wind speed [120,73]. Fig. 14. Vertical profile of the vertical/axial outflow wind speed [120,73].
36 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

1.2 to the gust load factor for synoptic wind, whereby a factor was pro-
DJ=500
posed for scaling up conventional wind loads so as to match the
DJ=1000
1 loads resulting from gust-front winds. This approach contains var-
DJ=1500
ious components affecting the loading due to gust-front winds,
DJ=2000
such as a kinematic effects factor for the variation in the vertical
0.8
profile of wind speed, a pulse dynamic factor for the dynamic
effects introduced by the sudden rise in wind speed, a structural
J
V /V

0.6 dynamics factor for the non-stationarity of turbulence in gust-


10

front winds and a potential load modification factor to represent


0.4 the transient aerodynamics.

3.1.3. Downburst wind field characteristics


0.2
Downburst is a localized event which makes the wind field very
dependent on the spatial location relative to the downburst center.
0 Fig. 12 illustrates the spatial variables of the downburst relative to
0 0.5 1 1.5 2 2.5 3 3.5 4
R/DJ a transmission tower. These variables represent the jet diameter
(DJ), and the location of the downburst center relative to the tower
Fig. 15. Downburst reference horizontal velocity (V10) at 10 m height [120,73].
(represented by the polar coordinates R and h).
Shehata et al. [120] conducted an extensive parametric study to
investigate the downburst wind profiles. Fig. 13 illustrates the vari-
1.4 ation of the radial velocity, normalized with respect to the jet
velocity, along the height. The maximum velocity profile was
1.2 Maximum peak found to occur at an R/DJ value of 1.2. The absolute maximum
velocity is approximately equal to 1.1 VJ. El Damatty et al. [47]
1 reported that the value 1.1 VJ can be considered as a design refer-
ence velocity at 10 m height (V10). Vertical component is typically
0.8 Constant value less than the horizontal component close to the ground (i.e. 100 m
VRD/VJ

or less) as shown in Fig. 14.


0.6 Mara et al. [93] found that the wind inclination can reach 20°
Minimum peak above horizontal in a critical case for a lattice tower. Fig. 15 shows
0.4 the variation of the reference velocity, V10, normalized with the jet
velocity, VJ, for different values of jet diameters. This figure indi-
0.2 cates that the absolute maximum velocity, which occurs at R/DJ
of 1.2, increases with decreased jet diameter. Furthermore, for
0 the time variation, the time history of the radial velocity compo-
0 100 200 300 400
nent (VRD), as well as the vertical velocity component (VVR), follows
Time (sec)
a trend with a maximum peak followed by a minimum peak as
Fig. 16. Typical time history of the horizontal/radial velocity [120,73]. shown in Figs. 16 and 17. The subsequent stabilization at a con-
stant value is an artifact of a simulation approach in which the
jet momentum source starts impulsively (correct) but then is left
0.6 on (non-physical). After the peak, wind speed converges to the
Maximum peak
steady state jet which is not realistic. It should be mentioned that
CFD simulations employing impinging jet properly simulate the
0.4
magnitude and the time instance of the peak velocities which are
the key parameters of the peak loads induced by downburst.
0.2
In the parametric study conducted by El Damatty et al. [47], the
Constant value
time of occurrence of the maximum radial velocity was found to be
VVR/VJ

proportional to the factor (DJ/VJ).


0 Table 4 summarizes the scope of previous studies focusing on
the downburst wind field and their main findings.

-0.2 3.2. Structural modeling of a transmission line under downburst


Minimum peak
loading

-0.4
0 100 200 300 400 3.2.1. Lattice tower and guy wires
Time (sec) Shehata et al. [120] developed a finite element model that sim-
ulates the tower members and the guys as a two-node, linear,
Fig. 17. Typical time history of the vertical/axial velocity [120,73]. three-dimensional frame element with three translational and
three rotational degrees of freedom per node. Each tower member
was simulated by one element while each guy was modeled by five
In the case of typical transmission lines with support tower heights elements. Rigid connections were assumed between the tower
of 50 m or less, dynamic response is usually neglected due to the members as these are physically connected using multi-bolted
high aerodynamic damping of the conductors and the high natural connections that can transfer moments. The conductors were stud-
frequency of the tower. Treating a downburst outflow as a gust ied separately and then their reactions were reversed and applied
front, Kwon and Kareem [77] presented an alternative analysis at the tower connection points. For the downburst loading, a pro-
framework called the gust-front factor approach, which is similar cedure was developed to scale up a small impinging jet wind field
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 37

Table 4
Previous studies on downburst wind field.

Study Scope of the study Comments


Fujita [48,49] Categorized downbursts events based on field Events are categorized to downburst, microburst, and macroburst
Wolfson et al. [140] observations of actual events Gust factors correspond to downburst events differ between
Wilson et al. [138] thunderstorm types
Choi [28]
Orwig and Schroeder [105]
Gunter and Schroeder [53]
Lombardo et al. [87]
De Gaetano et al. [39]
Solari et al. [123]
Didden and Ho [41] Simulated downburst flow experimentally Simulations were based on the analogy between downburst and
Donaldson and Snedeker [42] impinging jets
Oseguera and Bowles [106] Modeled of downbursts analytically Proposed series of models that account for velocity changes in space
Vicroy [131] and time
Chay et al. [22]
Holmes and Oliver [62]
Wood et al. [141]
Li et al. [80]
Su et al. [128]
Abd-Elaal et al. [1,2]
Kim et al. [74] Conducted impinging jet simulation of downburst events Validated steady and unsteady simulation of downburst using
Kim and Hangan [73] experimental impinging jet test and field measurements
Hangan et al. [57]
Selvam and Holmes [112]
Sengupta and Sarkar [113]
Mason et al. [95–97] Simulated downbursts using the cooling source model and Elevation of the peak velocity increases with the increase of the
Vermeire et al. [129] studied terrain exposure effect and topography roughness. Hill tends to increase the magnitude of the peak velocity
Hadziabdic [55] LES simulations of downburst using impinging jet. Various Turbulence wind field was characterized and it is found that
Chay et al. [22] terrain exposures are modeled using fractal surfaces downburst turbulence is very well correlated in circumferential
Gant [51] direction
Aboshosha et al. [8]
Vermeire et al. [129]
Orf et al. [103,104]
Lin et al. [82] Assessed the effect of translating velocity of downburst Provided an estimate for the translating velocity of downburst and
Kim and Hangan [73] assessed its contribution to the peak profile
Solari et al. [124] Proposed a thunderstorm response spectrum The proposed criteria allows the calculation of the maximum
structural response
Wieringa [137] Terrain roughness Provided an estimate for the aerodynamic roughness
Blocken et al. [18]
Kwon and Kareem [77] Assessed the turbulent component of downbursts Turbulence component of downburst is essential to quantify peak
Chen and Letchford [25,26] responses of structures
Chay et al. [22,23] An expression for downburst span reduction factor
Holmes et al. [63]
Aboshosha et al. [7]
Holmes et al. [61]
Shehata et al. [120] Lattice transmission towers subjected to downbursts Downburst size, location, and angle of attack significantly affect the
Mara et al. [93] structure response

[57,120]. Shehata and El Damatty [117,118], and Shehata et al. minor due to the relatively high natural frequency of the towers
[119] studied a guyed tower while Darwish and El Damatty [29] compared to the event natural frequency.
studied the behavior and the failure modes of a self-supported Lin et al. [81] developed an aeroelastic model for a single span
tower. El Damatty and Aboshosha [45] studied both guyed and of a transmission line. The guyed lattice tower was simplified to
self-supported towers to assess the behavior and the failure modes. an equivalent mast at a length scale of 1:100 while synoptic or
Ladubec et al. [78] improved upon the linear analysis by Shehata downburst wind loading were applied with a time scale of 1:10.
and El Damatty [118] and studied the P–D effect in tower response In either, case of atmospheric boundary layer or downburst wind
to a downburst wind field by using nonlinear space frame elements loadings, the structural response was generally quasi-static. Reso-
to simulate the tower members. The study showed an increase of nant dynamic response was less evident with the downburst wind
20% in the peak axial forces in the chord members of the main legs, than with the synoptic wind. Mara and Hong [94] studied the
as compared to the results from a linear analysis. Mara and Hong inelastic response of TL tower subjected to downburst wind and
[94] studied the inelastic response of TL tower subjected to down- found that tower’s capacity depends on the wind direction.
burst and synoptic wind fields. The study showed a dependence of
the tower capacity on the wind direction for both wind fields. 3.2.2. Conductor lines
The inclusion of the turbulent component in the structural anal- Finite Element Analysis (FEA) with the 2-D non-linear consis-
ysis may magnify the response due to the combined effects of the tent beam element [76] has been modified by Gerges and El
fluctuating (background) component and the resonant component, Damatty [52] to include the geometrical non-linear effects, and
which means a lower failure capacity. Wang et al. [135] studied the was utilized by Shehata et al. [120] to accurately simulate conduc-
dynamic effect of a downburst on tall transmission towers. Wind tor line properties and predict reactions for downburst loading.
tunnel tests were conducted to determine the wind load coeffi- Fig. 18 shows an element and each cable span is modeled as ten
cients of the transmission towers and then the towers were ana- elements. They considered the geometric non-linearity due to large
lyzed under downburst wind loading. Dynamic effects were line span and relatively small line cross-section, which causes large
38 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

(Displacement and Rotation)

(Displacement Only)

(Rotation Only)

Fig. 18. Consistent beam element coordinate systems and degrees of freedom (reproduced from Greges and El Damatty [52]).

Clamped End Clamped End


Y

X
Springs simulating insulator flexibility

Fig. 19. Modeling of the transmission line under study (reproduced from Shehata et al. [120]).

displacements from wind loading, as was modeled by Desai et al. cross-sectional area, projected area, span, insulator length and
[38] with a three-node iso-parametric finite element. The analysis sag. The inclusion of the flexibility of the insulators at tower/con-
was performed in the horizontal and the vertical directions sepa- ductor connections, rather than assuming fully-hinged boundary
rately to obtain the response under the radial and vertical down- conditions, also has a significant effect on the natural frequencies
burst velocities, respectively. The 2-D element was acceptable for and mode shapes. Line sag is inversely proportional to the pre-
downbursts as their associated velocities in the horizontal direc- tensioning force. The level of the pre-tensioning force is found to
tion are much higher than those in the vertical direction and, thus, have a major influence upon the natural periods and mode shapes
decoupling between the two directions can be justified. Shehata of line vibrations due to the effect of sag on the conductor’s stiff-
et al. [120] reported that modeling six conductor spans, as shown ness [30]. In addition, environmental factors such as temperature,
in Fig. 19, is enough to obtain accurate results of the transmitted highly affect the conductor sag where the actual sag at the time of
forces from the conductor to an intermediate tower. downburst occurrence may vary from reported values. High flexi-
Savory et al. [111] studied the failure of transmission towers in bility and the expected nonlinear behavior of the conductors result
the cases of tornado and downburst wind loading. Conductor in a time consuming FEA procedure, since many iterations are
forces were neglected and, as a result, failures were only associated required to investigate different downburst configurations that
with tornado loading and no failure was observed with downburst lead to the critical case of loading [4]. In view of this fact, there
loading. On the contrary, Shehata et al. [120] predicted three differ- is a need for a computationally efficient technique that can analyze
ent failure modes for transmission towers with downbursts. This is multi-span conductors under both vertical and horizontal HIW
mainly due to the strong effect of a downburst upon a relatively loading and can take into consideration the conductor properties.
localized region of a transmission line. The most critical failure Irvine [70] derived a closed-form solution for the reactions of a
mode was found to be due to the significant variation in the longi- single-spanned conductor, where the loading can be fitted with a
tudinal tensile forces from the lines upon the support towers. She- third-degree polynomial. In addition, Yu et al. [144] considered
hata et al. [120] revealed that this is the most critical failure mode highly concentrated loads to derive an exact solution to calculate
as the resultant, large, longitudinal force transmitted to the tower the reactions for a single conductor span. Both solutions neglected
cross arms leads to an out-of-plane bending moment in this region. the flexibility of the insulators, which Darwish et al. [30] concluded
El Damatty and Aboshosha [45] discussed a similar mode of failure. was important in quantifying the forces carried by the towers.
Aboshosha and El Damatty [4] conducted a parametric study to Although Winkelman’s earlier solution [139] accounted for the
check the expected values of the conductor longitudinal forces. insulator flexibility, the differences between tensile forces on con-
With a jet velocity of 40 m/s, the longitudinal force transmitted ductors in adjacent spans was neglected and, as such, the longitu-
to the tower due to the unbalanced tension was up to 60% of the dinal reactions that are transmitted from the conductors to the
transverse force. supporting towers would be ignored. Based on the analytical solu-
El Damatty et al. [47] showed that since the conductor struc- tion for an elastic catenary, Ahmadi-Kashani and Bell [11] and Wei
tural response is highly nonlinear, different conductor types would et al. [136] developed cable elements to simulate a whole span.
experience different responses based on the conductor characteris- Although this solution is more efficient, due to the reduction in
tics. They concluded that the main parameters that affect the lon- degrees of freedom, these elements were only developed for uni-
gitudinal and transverse forces on a conductor are its material, form wind load and are not suitable for HIW. Subsequently,
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 39

Table 5
Previous studies on the structural behavior of TL under downbursts.

Study Scope of the study Main finding/conclusion


Shehata et al. [119,120] Investigated the behavior of TLs (tower and conductor) under Internal forces in the tower members depends on the relative
Shehata and El Damatty downbursts loading location of the downburst to the tower
[117,118]
Darwish and El Damatty [29]
Ladubec et al. [78]
Kwon and Kareem [77] Proposed a new framework for the gust factor called the Gust The study resulted in a new gust factor approach that works for
Front Factor (GFF) approach synoptic and non-synoptic wind
This new GFF converges to the Davenport’s GF when used for
synoptic wind
Darwish et al. [30] Studied the effect of turbulent component of the downburst on A 20% increase in the maximum wire forces is due to the
the TL conductor response turbulence inclusion and negligible dynamic effect is noticed are
noticed
El Damatty et al. [47] Studying the critical load patterns of a downburst acting on a Three load cases were proposed representing the worst case
El Damatty and Elawady [46] TL conductor scenario of a downburst load acting on a TL
Aboshosha and El Damatty [5,9] Developed tools in the form of numerical technique and a The technique is found to be 180 times faster than FEM, while the
closed form solution for the conductor reaction due to closed form solution allows for a direct prediction of the
downburst maximum conductor reactions
Mara and Hong [94] Studied the inelastic response of TL tower subjected to the Yield and maximum tower’s capacities vary with the wind
wind coming from various directions direction
Aboshosha and El Damatty [6] Used statistical approach [35] to evaluate the span reduction Span reduction factor was found higher than the factor for normal
factor for the conductor reactions using downburst parameters wind. The obtained factors well-agrees with the field
obtained from LES [7] measurements Holmes et al. [61]

Aboshosha and El Damatty [5] developed a numerical technique to A summary of previous studies on the behavior of TLs (includ-
analyze multi-spanned conductors under HIW. This solution is the ing either tower, conductor or both) is provided in Table 5. The
first semi-closed form solution for a multi-spanned conductor sys- table shows the scope of the studies and their main findings.
tem under non-uniform loading in both the vertical and horizontal
directions, where insulator rigidity is considered. The technique
was approximately 185 times faster than FEA. Later, Aboshosha 3.3. Behavior of transmission lines subjected to downbursts and
and El Damatty [9] proposed an engineering method suitable for corresponding critical load cases
practitioner engineers to evaluate the maximum conductor reac-
tions due to downburst. A large research has been conducting in the University of Wes-
Darwish et al. [30] modified the two-dimensional nonlinear tern Ontario regarding the behavior of the transmission line struc-
finite element model of the transmission lines developed by She- tures subjected to downburst wind events. Shehata et al. [120],
hata et al. [120] to study the dynamic characteristics of the conduc- Shehata and El Damatty [117], Shehata et al. [119], and Darwish
tors under turbulent downburst loading. The turbulence and El Damatty [29] adopted quasi-static analysis procedures (rel-
component was extracted from full-scale data and then added to evant for a mean component that varies slowly with time) with the
the mean component of the downburst wind field developed by downburst wind field from Hangan and Kim [56] as an input. She-
Kim and Hangan [73]. Large deformations and the pre-tension hata et al. [120] reported that the effective period for downburst
loading were modeled. The aerodynamic damping was determined wind speed variation ranged between 20 and 22 s while the vibra-
with an expression given by Davenport [32], which was derived for tion frequencies for the transmission tower and conductor were
synoptic winds and assumed an average velocity in the calculation 0.58 s and 8.25 s, respectively. As such, no strong dynamic effect
of the aerodynamic damping in order to overcome the localized was evident.
nature of downbursts as the wind velocity varies with time and Using the structural analysis model developed by Shehata et al.
also spatially along the conductor length. The study concluded that [120], Shehata and El Damatty [117] conducted a parametric study
the resonant component due to the turbulence is negligible due to to investigate the critical downburst configuration by varying the
the large aerodynamic damping, the dynamic response is mainly jet diameter (DJ) and the location of the downburst center relative
due to the background component of wind velocity fluctuations to the tower (R). A guyed transmission tower, which collapsed in
and turbulence accounted for almost 20% of the conductor deflec- Manitoba in 1996 due to a downburst event, was used to perform
tion and reactions. Aboshosha et al. [10] studied the dynamic and this parametric study. The critical downburst parameters, in terms
quasi-static response of single span and multi spanned lines under of the size of the event and its location relative to the tower, lead-
both the synoptic and downburst loads. For multi-spanned lines, ing to maximum forces in the tower members, were identified. The
the study showed that the resonant component contribution com- study revealed that the critical downburst parameters vary based
pared to the peak reactions is in order of 6% assuming different on the type and location of the members. Unsurprisingly, the chord
wind intensities. However, for single span conductor, the resonant members, diagonal members and cross arm members had different
component contribution to the peak response was higher (in order critical downburst configurations. Shehata and El Damatty [118]
of 16%) at low speeds. extended their numerical scheme by including a failure model
Holmes et al. [63] isolated the turbulent component of the for the tower members, which was used to study the progressive
downburst velocity and produced a peak load reduction factor collapse of the tower failure. An optimization routine was imple-
for the spatial variation along the longitudinal direction from a mented by Shehata et al. [119] to predict the critical downburst
set of impinging jet CFD data [74]. The span reduction factor was parameters and the corresponding forces on a transmission line
found to range between 1 and 0.8 for a separation of 720 m with by an automated procedure. This finite element-optimization tech-
downburst wind loading. A similar result was found by Aboshosha nique was validated by comparing the maximum forces and critical
and El Damatty [6] using downburst wind field obtained from LES. downburst parameters in a number of tower members to the cor-
The effect of terrain roughness on SRF remains uncertain. responding values obtained from an extensive parametric study.
40 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

line connecting the center of the downburst and the tower of inter-
est is perpendicular on the line. Under this load case, the conduc-
tors will experience significant transverse reaction. This force
will cause a large shear force in the guys area. This load case is sim-
ilar to the normal wind load case required by the codes. Thus, the
main difference between the downburst under this load configura-
tion and the synoptic wind load would be the difference in both
wind speeds. Shehata and El Damatty [118] reported that one of
the major failure mechanisms for this type of structures occurs
when the virtual line connecting the downburst center to the cen-
ter of the transmission line is parallel to the line. This causes a zero
reaction contribution from the conductors. This load case caused
failures initiation in the chord members as well as the guys zone.
This was then confirmed by the late study conducted by Darwish
and El Damatty [29] who studied the behavior of self-supported
tower, Aboshosha and El Damatty [3] who studied one guyed
tower and another self-supported tower, and El Damatty and Ela-
wady [46] who studied the behavior of three guyed and three
Fig. 20. Field evidence of lattice tower failure at cross arms (modified from self-supported towers. Fig. 20 shows a failure of cross arms docu-
Manitoba Hydro [92]). mented by an electric energy provider.
Darwish et al. [30] investigated the importance of the transla-
tion velocity of the downburst, using three critical downburst con-
figurations. It was found that the tower failed at the same radial
Shehata and El Damatty [118] conduct a failure analysis for a
velocity regardless of the contribution of the translational compo-
guyed tower subjected to a microburst where they compared the
nent to this velocity. Therefore, Darwish et al. [30] reported that
internal forces developing in each of the tower member to its
there is no need to consider the translation velocity of the down-
capacity calculated using the ASCE (1992) standards. In their study,
burst and it is sufficient to vary the location of the downburst in
they considered a microburst jet velocity of VJ = 70 m/s. The analy-
space for a large number of separate stationary events. In addition,
sis showed that a failure region was generated at the cross arm and
Darwish et al. [30] found almost no variation in the dynamic char-
the guys zones. This is justified since the downburst longitudinal
acteristics of the conductors under the different loading configura-
reaction affected the cross arm zone; this load configuration is
tions. Darwish and El Damatty [29], El Damatty and Aboshosha
not considered in the load cases recommended by the standards.
The authors also indicated a failure mechanism when the virtual

Fig. 21. (a) VRD distribution along tower height with h = 90° and 0°, R/DJ = 1.3 and Fig. 22. (a) VRD distribution along tower height with h = 30°, R/DJ = 1.6 and
DJ = 500 m and (b) VRD distribution over six line spans with h = 0°, R/DJ = 1.3 and DJ = 500 m and (b) VRD distribution over six line spans with h = 30°, R/DJ = 1.6 and
DJ = 500 m [46]. DJ = 500 m (modified from El Damatty and Elawady [46]).
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 41

Table A.1
Pressures at 30 m elevation qpn(30) normalized by the peak pressure at 10 m for open terrain exposure.

Col. No. 1 2 3 4 5 6 7
Codenexposure qpn(30)ASCE qpn(30) AS/NZS CRp(30) (AS/NZS) qpn(30) BS CRp(30) (BS) qpn(30)IEC CRp(30) (IEC)
Sea side 1.43 1.43 1 1.47 1.03 1.30 0.91
Open 1.26 1.33 1.05 1.31 1.04 1.22 0.97
Suburban 0.98 0.89 0.91 1.11 1.13 1.02 1.04

qpn: normalized pressure calculated using Eq. (4).


Code ratio for pressure CRp using ASCE 74 [12] as a datum for normalization.
CRp(30) (AS/NZS) = Column (2)/Column (1).
CRp(30) (BS) = Column (4)/Column (1).
CRp(30) (IEC) = Column (6)/Column (1).

Table A.2
Normalized forces transferred from a 20 m elevated conductor, Fcn (20).

Col. no. 1 2 3 4 5 6 7
Coden L Fcn (20) ASCE Fcn (20) AS/NZS CRFc(20) (AS/NZS) Fcn (20) BS CRFc(20) (BS) Fcn (20) IEC CRFc(20) (IEC)
Sea side exposure
200 0.90 1.02 1.12 1.24 1.37 1.12 1.23
300 0.87 0.94 1.08 1.20 1.39 1.07 1.23
400 0.84 0.88 1.05 1.17 1.39 1.04 1.23
500 0.82 0.85 1.03 1.15 1.39 1.01 1.22
600 0.81 0.83 1.02 1.13 1.4 0.99 1.21
700 0.80 0.82 1.02 1.12 1.4 0.97 1.21
800 0.79 0.81 1.02 1.11 1.4 0.95 1.20
Open exposure
200 0.84 0.90 1.07 1.03 1.23 1.08 1.29
300 0.80 0.82 1.03 0.99 1.24 1.03 1.29
400 0.77 0.78 1.01 0.96 1.24 1.00 1.29
500 0.75 0.75 0.99 0.94 1.24 0.98 1.29
600 0.74 0.73 0.99 0.92 1.24 0.95 1.29
700 0.73 0.72 0.99 0.91 1.24 0.94 1.28
800 0.72 0.72 0.99 0.90 1.24 0.92 1.28
Suburban exposure
200 0.70 0.62 0.88 0.77 1.08 0.98 1.39
300 0.67 0.57 0.85 0.73 1.09 0.94 1.40
400 0.64 0.54 0.84 0.70 1.09 0.91 1.41
500 0.62 0.52 0.83 0.68 1.09 0.89 1.42
600 0.61 0.51 0.83 0.66 1.08 0.87 1.42
700 0.60 0.50 0.84 0.64 1.08 0.85 1.42
800 0.59 0.50 0.84 0.63 1.07 0.84 1.42

Fcn: normalized conductor force calculated using Eq. (7).


Code ratio for conductor forces CRFc using ASCE 74 [12] as a datum for normalization.
CRFc(20) (AS/NZS) = Column (2)/Column (1).
CRFc(20) (BS) = Column (4)/Column (1).
CRFc(20) (IEC) = Column (6)/Column (1).

[45], Aboshosha and El Damatty [4] and El Damatty et al. [47] per- (2) At h = 90°, R/DJ = 1.30 and DJ = 500 m, the same wind load
formed additional parametric studies to investigate the critical profile is applied on the tower face parallel to the major axis
downburst configurations. The studies agreed that changing the of the line, as in the previous case, while no wind load is
location of the downburst (R/DJ and h) has a stronger effect on applied on the conductors.
the value of the axial force in all tower members when compared (3) At h = 30°, R/DJ = 1.60 and DJ = 500 m, the vertical profiles of
to changing the downburst size (DJ) which has a minor effect. In velocity in the direction perpendicular and parallel to the
addition, the ratio of the span (L) to the jet diameter, DJ, plays a line can be approximated as having uniform values of 0.8
vital role in the existence of these critical cases. El Damatty et al. VJ and 0.47 VJ, respectively, to stipulate the tower wind load.
[47] proposed three critical load cases to calculate static forces act- A difference in downburst radial velocity distribution at the
ing upon towers and conductors under downburst loading, where two adjacent conductors results in a nonlinear longitudinal
the loads should be applied on both the tower of interest as well force acting on the tower, Fig. 22. The transverse reaction
as adjacent conductor spans from each side. The three cases are of the wires can be estimated assuming a uniform velocity
shown in Figs. 13 and 14 and explained below: of 0.65 VJ.
Comparison between load cases described above and load cases
(1) At h = 0°, R/DJ = 1.30 and DJ = 500 m, the downburst outflow
for synoptic wind shows two main differences (1) Downburst has
is modeled as impacting perpendicular to the major axis of
an oblique case of loading (h = 30°) causing differential conductor
the transmission line. The radial velocity applied on the sup-
tension on the sides of the tower. Synoptic wind does not generate
port tower can be considered to be equal to 1.1 VJ. To
differential conductor. (2) Downburst load cases in the transverse
approximate the conductor’s transverse reaction, an equiva-
and longitudinal directions (h = 0° and 90°) are close to synoptic
lent uniform velocity distribution with a magnitude of 1.06
wind load cases, but the load magnitude is mostly higher. Load
VJ can be used, as shown in Fig. 21.
magnitude for downburst requires further research.
42 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

Table A.3
Normalized forces transferred from a 40 m elevated conductor, Fcn (40).

Col. no. 1 2 3 4 5 6 7
Coden L Fcn (40) ASCE Fcn (40) (AS/NZS) CRFc(40) (AS/NZS) Fcn (40) BS CRFc(40) (BS) Fcn (40) IEC CRFc(40) (IEC)
Sea side exposure
200 1.00 1.11 1.11 1.42 1.42 1.17 1.17
300 0.96 1.02 1.06 1.38 1.43 1.12 1.17
400 0.93 0.97 1.03 1.34 1.43 1.09 1.16
500 0.92 0.93 1.01 1.32 1.44 1.06 1.15
600 0.90 0.91 1.01 1.30 1.44 1.04 1.15
700 0.89 0.90 1.01 1.28 1.44 1.02 1.14
800 0.88 0.89 1.01 1.27 1.43 1.00 1.13
Open exposure
200 0.95 1.07 1.13 1.21 1.28 1.14 1.21
300 0.90 0.99 1.09 1.16 1.28 1.10 1.21
400 0.88 0.93 1.06 1.13 1.29 1.06 1.20
500 0.86 0.90 1.04 1.11 1.29 1.03 1.20
600 0.84 0.88 1.04 1.09 1.29 1.01 1.20
700 0.83 0.87 1.04 1.07 1.28 0.99 1.19
800 0.82 0.86 1.04 1.05 1.28 0.98 1.19
Suburban exposure
200 0.81 0.70 0.86 0.94 1.15 1.05 1.29
300 0.77 0.65 0.84 0.89 1.15 1.00 1.30
400 0.74 0.61 0.82 0.86 1.14 0.97 1.30
500 0.73 0.59 0.81 0.83 1.14 0.95 1.30
600 0.71 0.58 0.81 0.81 1.14 0.93 1.30
700 0.70 0.57 0.81 0.79 1.13 0.91 1.30
800 0.69 0.56 0.82 0.77 1.12 0.90 1.30

Fcn: normalized conductor force calculated using Eq. (7).


Code ratio for conductor forces CRFc using ASCE 74 [12] as a datum for normalization.
CRFc(40) (AS/NZS) = Column (2)/Column (1).
CRFc(40) (BS) = Column (4)/Column (1).
CRFc(40) (IEC) = Column (6)/Column (1).

4. Conclusions employed in each code in addition to the various methods used


in the codes to covert between the wind speeds averaged over
This section summarizes codes guidelines related to response of 10 min and 3 s.
TLs under synoptic and non-synoptic winds. Gaps exist in the  The differences between the four codes in terms of design pres-
codes related to the response of TLs under synoptic wind and pos- sure, gust response factors for the conductors, GC, and the tower,
sible ways of covering these gaps are summarized below: GT, and expressions for the tower force FT lead to larger differ-
ences between the forces estimated by those codes. The dis-
 The tower frequency and the conductors’ aerodynamic damp- crepancy of the code ratio for conductor forces CRFc is found
ing, below which the resonant component needs to be to be ranging between 0.81 and 1.44. For tower forces code ratio
accounted for, require further investigations. CRFt, a discrepancy range of 0.68 and 1.85 is noticed. Assessment
 The assumption of the minor dynamic effect of the conductors of the accuracy of the estimated forces based on various codes
due to the high aerodynamic damping is true for high wind can be conducted based on wind tunnel testing or full scale
speeds. This assumption requires further investigations under measurements of real towers.
low wind speeds. Vibration of the conductors under low wind  The ASCE 74 [12] code uses 3-s gust velocity as the reference
speed might lead to fatigue problems in the cross arms. wind speed, however, when the gust response of the structure
 Most of TL codes are valid for towers with an overall height less is calculated, the code uses a formulation that is based on
than 60 m. For taller towers, the dynamic effect of the towers 10 min-averaged wind speed. The code introduces a conversion
may be accounted for using the gust factor approach while con- factor Kv based on Durst curve, which is different than the
sidering the resonant component. Accounting for the resonant approaches adopted by the BS [20] and the IEC [69]. This con-
component by the gust factor approach can be done using a fre- version leads to differences in the order of 10–15%.
quency domain analysis, which is typical applicable for linear  The ASCE 74 [12] does not account for the increase of the tower
structures. Research needs to be done to modify the frequency force for skewed angle of attack up to 45°. This appeared to
domain analysis in order to account for the nonlinear behavior underestimate the tower forces as found from a previous wind
expected for tall towers. tunnel study.
 Design pressure obtained by the AS/NZS [13], BS [20] and IEC [69]  Design codes separate between the gust factor for the conductor
codes and normalized by values obtained from ASCE 74 [12] and that for the tower. For the cases of no dynamic effect, this
shows that the discrepancy in the code ratio CRp is ranging may overestimate the forces transmitted at the tower base as
between 0.91 and 1.13. The minimum value of CRp is noticed in such an assumption means full correlation between the conduc-
case of using AS/NZS [13] and IEC [69] approaches for suburban tor and the tower forces. A reduction factor can be applied to the
and sea side exposures, respectively. The maximum value of conductor forces in this case. For the cases of potential dynamic
CRp is reached by using BS [20] approach for Suburban exposure. effects, the dynamic properties of the system, represented in
This discrepancy is due to the different velocity profiles the modal frequencies, mode shapes and modal aerodynamic
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 43

Table A.4
Normalized forces Ftn acting on a tower panel for angle of attack 0°.

Col. no. 1 2 3 4 5 6 7
Code n u Ftn (0°) ASCE Ftn (0°) AS/NZS CRFt(0°) (AS/NZS) Ftn (0°) BS CRFt(0°) (BS) Ftn (0°) IEC CRFt(0°) (IEC)
Sea side exposure
0.0 3.46 4.39 1.27 4.64 1.34 3.84 1.11
0.1 3.10 3.84 1.24 4.05 1.31 3.36 1.08
0.2 2.65 3.30 1.25 3.49 1.32 2.89 1.09
0.3 2.20 2.86 1.30 3.02 1.37 2.51 1.14
0.4 1.75 2.52 1.44 2.65 1.52 2.20 1.26
0.5 1.56 2.26 1.45 2.38 1.53 1.98 1.27
0.6 1.56 2.09 1.34 2.21 1.42 1.83 1.17
0.7 1.55 2.02 1.35 2.13 1.37 1.76 1.14
0.8 1.61 2.03 1.26 2.15 1.33 1.78 1.11
0.9 1.67 2.14 1.28 2.26 1.35 1.87 1.12
1.0 1.73 2.34 1.35 2.47 1.43 2.05 1.18
Open exposure
0.0 3.67 3.94 1.07 3.94 1.07 4.16 1.13
0.1 3.29 3.44 1.05 3.44 1.05 3.63 1.11
0.2 2.81 2.96 1.06 2.96 1.06 3.13 1.11
0.3 2.33 2.57 1.10 2.57 1.11 2.71 1.16
0.4 1.85 2.26 1.22 2.26 1.22 2.38 1.28
0.5 1.65 2.03 1.23 2.03 1.23 2.14 1.29
0.6 1.65 1.88 1.14 1.88 1.14 1.98 1.20
0.7 1.64 1.81 1.10 1.81 1.11 1.91 1.16
0.8 1.71 1.82 1.07 1.82 1.07 1.92 1.13
0.9 1.77 1.92 1.08 1.92 1.08 2.03 1.14
1.0 1.84 2.10 1.14 2.10 1.14 2.22 1.21
Suburban exposure
0.0 4.40 3.37 0.77 3.08 0.70 4.91 1.12
0.1 3.94 2.95 0.75 2.69 0.68 4.28 1.09
0.2 3.37 2.54 0.75 2.32 0.69 3.69 1.10
0.3 2.79 2.20 0.79 2.01 0.72 3.20 1.15
0.4 2.22 1.93 0.87 1.76 0.79 2.81 1.26
0.5 1.98 1.73 0.88 1.58 0.80 2.52 1.27
0.6 1.98 1.61 0.81 1.47 0.74 2.34 1.18
0.7 1.97 1.55 0.79 1.41 0.72 2.25 1.14
0.8 2.05 1.56 0.76 1.43 0.70 2.27 1.11
0.9 2.12 1.64 0.77 1.50 0.71 2.39 1.13
1.0 2.20 1.80 0.82 1.64 0.75 2.61 1.19

Fct: normalized force acting on tower calculated using Eq. (8).


u: Solidity ratio.
Code ratio for tower forces CRFt using ASCE 74 [12] as a datum for normalization.
CRFt(0°) (AS/NZS) = Column (2)/Column (1).
CRFt(0°) (BS) = Column (4)/Column (1).
CRFt(0°) (IEC) = Column (6)/Column (1).

damping of the combined conductor-tower system have to be  Effect of terrain and topography on the downburst wind field
obtained, while accurately modeling the connection between needs further research to be considered in the design codes.
the tower and the conductor. Then, these dynamic properties  Along the line spans, and due to the localized nature of down-
can be used through the gust factor approach proposed by Daven- bursts, a nonsymmetrical distribution of the wind pressure
port [32] which is adopted by design codes for dynamically sen- along cables’ length can happen. This can lead to an unbalanced
sitive structures. An expression for the aerodynamic damping of longitudinal force causing an out-of-plane moment on the
the combined system is also required. tower cross arms. Such a load case needs to be considered in
the downburst analysis of the towers and thus in the codes of
Current guidelines related to response of TLs under non-synoptic practice.
winds are summarized below:  Span reduction factor is only given in the AS/NZS:7000 [13] for
 In most of the available codes, no clear definition is given for the the case of downburst event. However, the factor provided in
wind field associated with the downburst events. this code does not account for the terrain effect or the elevation
o Only AS/NZS:7000 [13] has defined the downburst wind of the conductor above the ground.
field as a cold air column that fall vertically from a great  For flexible tall towers, the dynamic effect should be considered
height and strike the ground causing the wind draft to through the gust factor approach proposed by Kwon and Kar-
radiate from the impact site. eem [77]. However, this approach requires further research.
o ASCE 74-2010 considers the downburst wind speed simi-
lar to the F2 tornado scale.
 Regarding the velocity profile along the height, AS/NZS:7000 Acknowledgment
[13] provides the vertical distribution of the wind speed. How-
ever, due to the localized nature of downbursts, the velocity The authors would like to thank the CEATI International (www.
profile will depend on the downburst size and its relative loca- ceati.com) and its members from the Transmission Overhead Line
tion to the tower. This will lead to a number of critical load & Extreme Events Mitigation program – 28 leading transmission
cases, which should be incorporated in the design codes. electrical utilities form all over the globe for their financial support
44 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

Table A.5
Normalized forces Ftn acting on a square tower panel for angle of attack 45°.

Col. no. 1 2 3 4 5 6 7
Code n u Ftn (45°) ASCE Ftn (45°) AS/NZS CRFt(45°) (AS/NZS) Ftn (45°) BS CRFt(45°) (BS) Ftn (45°) IEC CRFt(45°) (IEC)
Sea side exposure
0.0 3.46 4.88 1.41 5.56 1.61 4.61 1.33
0.1 3.10 4.26 1.37 4.86 1.57 4.03 1.30
0.2 2.65 3.67 1.38 4.19 1.58 3.47 1.31
0.3 2.20 3.34 1.52 3.63 1.65 3.01 1.37
0.4 1.75 3.07 1.75 3.19 1.82 2.64 1.51
0.5 1.56 2.88 1.85 2.86 1.84 2.37 1.52
0.6 1.56 2.55 1.64 2.65 1.70 2.20 1.41
0.7 1.55 2.35 1.52 2.55 1.65 2.12 1.37
0.8 1.61 2.26 1.40 2.58 1.60 2.14 1.33
0.9 1.67 2.38 1.42 2.71 1.62 2.25 1.35
1.0 1.73 2.60 1.50 2.97 1.71 2.46 1.42
Open exposure
0.0 3.67 4.37 1.19 4.73 1.29 4.99 1.36
0.1 3.29 3.82 1.16 4.13 1.26 4.36 1.33
0.2 2.81 3.29 1.17 3.56 1.27 3.75 1.34
0.3 2.33 2.99 1.28 3.08 1.32 3.25 1.40
0.4 1.85 2.75 1.48 2.71 1.46 2.86 1.54
0.5 1.65 2.58 1.56 2.43 1.47 2.56 1.55
0.6 1.65 2.29 1.39 2.25 1.36 2.38 1.44
0.7 1.64 2.11 1.28 2.17 1.32 2.29 1.39
0.8 1.71 2.02 1.19 2.19 1.28 2.31 1.35
0.9 1.77 2.13 1.20 2.31 1.30 2.43 1.37
1.0 1.84 2.33 1.27 2.52 1.37 2.66 1.45
Suburban exposure
0.0 4.40 3.74 0.85 3.70 0.84 5.89 1.34
0.1 3.94 3.27 0.83 3.23 0.82 5.14 1.31
0.2 3.37 2.82 0.84 2.78 0.83 4.43 1.32
0.3 2.79 2.56 0.92 2.41 0.86 3.84 1.37
0.4 2.22 2.36 1.06 2.12 0.95 3.37 1.52
0.5 1.98 2.21 1.12 1.90 0.96 3.03 1.53
0.6 1.98 1.96 0.99 1.76 0.89 2.80 1.42
0.7 1.97 1.80 0.92 1.70 0.86 2.70 1.37
0.8 2.05 1.73 0.85 1.71 0.84 2.73 1.33
0.9 2.12 1.83 0.86 1.80 0.85 2.87 1.35
1.0 2.20 2.00 0.91 1.97 0.90 3.14 1.43

Fct: normalized force acting on tower calculated using Eq. (8).


u: Solidity ratio.
Code ratio for tower forces CRFt using ASCE 74 [12] as a datum for normalization.
CRFt(45°) (AS/NZS) = Column (2)/Column (1).
CRFt(45°) (BS) = Column (4)/Column (1).
CRFt(45°) (IEC) = Column (6)/Column (1).

of this research. Also, the authors would like to thank the anon- [8] Aboshosha H, Bitsumalak G, El Damatty A. LES of lower ABL flow in the built-
environment using roughness modeled by fractal surfaces. J Sust Cities Soc
ymous reviewers for their insightful comments which enhanced
2015;19:46–60.
the quality of the paper. [9] Aboshosha H, El Damatty A. Engineering method for estimating the reactions
of transmission line conductors under downburst winds. J Eng Struct
2015;142:198–216.
[10] Aboshosha H, El Damatty A. Dynamic effect for transmission line conductors
Appendix A
under downburst and synoptic winds. Wind Struct Int J 2015;21(2):241–72.
[11] Ahmadi-Kashani K, Bell AJ. The analysis of cables subject to uniformly
Results for the synoptic winds obtained using various codes. distributed loads. Eng Struct 1988;10(3):174–84.
See Tables A.1–A.5. [12] American Society of Civil Engineers (ASCE). Guidelines for electrical
transmission line structural loading. ASCE manuals and reports on
engineering practice, No. 74, New York, NY, USA; 2010.
[13] Australian Standard (AS) 1170.2. Australian standard for minimum design
References loads on structures – Part 2: wind loads, Standard Australia, North Sydney,
Australia; 2002.
[1] Abd-Elaal E, Mills JE, Ma X. An analytical model for simulating steady state [14] Australian Standard/New Zealand Standard (AS/NZS) 7000. Overhead line
flows of downburst. J Wind Eng Ind Aerodyn 2013;115:53–64. design detailed procedures. Standards Australia Limited/Standards New
[2] Abd-Elaal E, Mills JE, Ma X. Empirical models for predicting unsteady-state Zealand, North Sydney, Australia; 2010.
downburst wind speeds. J Wind Eng Ind Aerodyn 2014;129:49–63. [15] Bachmann et al. Vibration problems in structures. Basel: Birkhauser; 1995.
[3] Aboshosha H, El Damatty A. Capacity of electrical transmission towers under [16] Battista R, Rodrigues R, Pfeil M. Dynamic behavior and stability of
downburst loading. In: ASEA-SEC-1, Perth, November 28–Decmber 2; 2012. transmission line towers under wind forces. J Wind Eng Ind Aerodyn
[4] Aboshosha H, El Damatty A. Downburst induced forces on the conductors of 2003;91:1051–67.
electric transmission lines and the corresponding vulnerability of towers [17] Behncke RH, Ho TCE. Review of span and gust factors for transmission line
failure. In: CSCE 2013 conference, 29 May–1 June, Montreal, Canada; 2013. design. In: Proceedings of the 2009 electrical transmission and substation
[5] Aboshosha H, El Damatty A. Effective technique for the reactions of structures conference, 8–12 November, Fort Worth, TX, USA, published in
transmission line conductors under high intensity winds. Wind Struct ‘‘Electrical transmission and substation structures 2009: Technology for the
2014;18(3):235–52. next generation”, 2010, edited by M.W. Vogt, sponsored by Electrical
[6] Aboshosha H, El Damatty A. Span reduction factor of transmission line Transmission Structures Committee of the Structural Engineering Institute
conductors under downburst winds. J Wind Eng 2014;11(1):13–22. of ASCE; 2009. p. 209–220.
[7] Aboshosha H, Bitsuamlak G, El Damatty A. Turbulence characterization of [18] Blocken B, Stathopoulos T, Carmeliet J. CFD simulation of the atmospheric
downbursts using LES. J Wind Eng Ind Aerodyn 2015;136:44–61. boundary layer: wall function problems. Atmos Environ 2007;41(2):238–52.
H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46 45

[19] British Standard (BS) 8100. British standard for lattice towers and masts: Part [54] Gurunga CB, Yamaguchia H, Yukinob T. Identification and characterization of
1 – code of practice for loading, London, UK; 1986. galloping of Tsuruga test line based on multi-channel modal analysis of field
[20] British Standard (BS) 50341. Overhead electrical lines exceeding AC data. J Wind Eng Ind Aerodyn 2003;91:903–24.
45 kV: Part 1 – General requirements – Common specifications, London, [55] Hadžiabdić M. LES, RANS and combined simulation of impinging flows and
UK; 2001. heat transfer. Ph.D thesis. University of Sarajevo; 2005. 185p.
[21] Boddy D, Rice J. Impact of alternative galloping criteria on transmission line [56] Hangan H, Kim J. Numerical simulations of impinging jets with application to
design. In: Proc ASCE electrical transmission and substation structures 2009, downbursts. J Wind Eng Ind Aerodyn 2007;95(4):279–98.
ASCE, Reston, VA; 2009. [57] Hangan H, Roberts D, Xu Z, Kim J. Downburst simulation. Experimental and
[22] Chay MT, Albermani F, Wilson R. Numerical and analytical simulation of numerical challenges. In: Proceedings of the 11th international conference on
downburst wind loads. Eng Struct 2006;28(2):240–54. wind engineering, 1–5 June, Lubbock, TX, USA; 2003.
[23] Chay MT, Wilson R, Albermani F. Gust occurrence in simulated non- [58] Harikrishna P, Shanmugasundaram J, Gomathinayagam S, Lakshmanan N.
stationary winds. J Wind Eng Ind Aerodyn 2008;96(10):2161–72. Analytical and experimental studies on the gust response of a 52 m tall steel
[24] Chabart O, Lilien JL. Galloping of electrical lines in wind tunnel facilities. J lattice tower under wind loading. Comput Struct 1999;70:149–60.
Wind Eng Ind Aerodyn 1998;74–76:967–76. [59] Havard DG, Pohlman JC. Control of galloping conductors by detuning Paper
[25] Chen L, Letchford CW. A deterministic–stochastic hybrid model of 22–05. Paris (France): CIGRE; 1980.
downbursts and its impact on a cantilevered structure. Eng Struct 2004;26 [60] Hjelmfelt MR. Structure and life cycle of microburst outflows observed in
(5):619–29. Colorado. J Appl Meteorol 1988;27:900–27.
[26] Chen L, Letchford CW. Parametric study on the along-wind response of the [61] Holmes JD. Recent developments in the specification of wind loads on
CAARC building to downbursts in the time domain. J Wind Eng Ind Aerodyn transmission lines. J Wind Eng 2008;5(1):8–18.
2004;92(9):703–24. [62] Holmes JD, Oliver SE. An emperical model of a downburst. Eng Struct
[27] Cook NJ. The designer’s guide to wind loading of building structures Part 1. 2000;22:1167–72.
Building research establishment, London, UK; 1985. [63] Holmes JD, Hangan HM, Schroeder JL, Letchford CW, Orwig KD. A forensic
[28] Choi ECC. Wind characteristics of tropical thunderstorms. J Wind Eng Ind study of the Lubbock-Reese downdraft of 2002. Wind Struct 2008;11
Aerodyn 2000;84:215–26. (2):137–52.
[29] Darwish MM, El Damatty AA. Behavior of self supported transmission line [64] Holmes JD. Along-wind response of lattice towers: Part I – derivation of
towers under stationary downburst loading. Wind Struct 2011;14(5):481–4. expressions for gust response factors. Eng Struct 1994;16:287–92.
[30] Darwish M, El Damatty A, Hangan H. Dynamic characteristics of transmission [65] Holmes JD. Along-wind response of lattice towers: II – aerodynamic damping
line conductors and behaviour under turbulent downburst loading. Wind and deflections. Eng Struct 1996;18:483–8.
Struct 2010;13(4):327–46. [66] Holmes JD. Along-wind response of lattice towers – III. Effective load
[31] Davenport AG. Buffeting of a suspension bridge by storm winds. ASCE J Struct distributions. Eng Struct 1996;18:489–94.
Div 1962;88(3):233–64. [67] Houle S, Ghannoum E, Hardy C. Static and dynamic testing of transmission
[32] Davenport AG. Gust loading factors. ASCE J Struct Div 1967;93(3):11–34. lines subjected to real wind conditions. In: CIGRE symposium, Leningrad;
[33] Davenport AG. Gust response factors for transmission line loading. In: 1991.
Proceedings. 5th international conference on wind engineering, Fort Collins, [68] Indian Standard (IS) 875. Indian standard code of practice for design loads
Colorado. Pergamon Press; 1979. p. 899–909. (other than earthquake) for buildings and structures, Part 3: Wind loads, New
[34] Davenport AG. Gust response factors for transmission line loading. In: Delhi, India; 1987.
Cermak JE, editor. Proc of the fifth international conference on wind [69] International Electrotechnical Commission (IEC) 60826. Design criteria of
engineering, Fort Collins, CO, July 1979. New York (NY): Pergamon Press; overhead transmission lines. Geneva, Switzerland; 2003.
1980. [70] Irvine HM. Cable structures. Cambridge (MA, USA): MIT Press; 1981.
[35] Davenport AG. How can we simplify and generalize wind loads?. In: [71] Ivan M. A ring-vortex downburst model for flight simulations. J Aircraft
Presented at the third asia-pacific symposium on wind engineering. 1986;23:232–6.
December 13–D15. Hong-Kong, Keynote Lecture; 1993. [72] Jones KF. Coupled vertical and horizontal galloping. ASCE J Eng Mech
[36] Davison et al. Alcoa overhead engineering data no. 4. Alcoa. 1992;118(1):92–107.
[37] Dempsey D, White H. Winds wreak havoc on lines. Transm Distrib World [73] Kim J, Hangan H. Numerical simulations of impinging jets with application to
1996;48(6):32–7. downbursts. J Wind Eng Ind Aerodyn 2007;95(4):279–98.
[38] Desai YM, Yu P, Popplewell N, Shah AH. Finite element modelling of [74] Kim JD, Hangan H, Ho TCE. Downburst versus boundary layer induced loads
transmission line galloping. Comput Struct 1995;57(3):407–20. for tall buildings. Wind Struct 2007;10(5):481–94.
[39] De Gaetano P, Repetto MP, Repetto T, Solari G. Separation and classification of [75] Kishnasamy S. Wind and ice loading research in ontario hydro. In:
extreme wind events from anemometric records. J Wind Eng Ind Aerodyn Proceedings of the fifth US national conference on wind engineering,
2014;126:132–43. Lubbock, Texas; 1985.
[40] Den Hartog JP. Transmission line vibration due to sleet. Trans AIEE [76] Koziey B, Mirza F. Consistent curved beam element. Comput Struct 1994;51
1932;51:1074–86. (6):643–54.
[41] Didden N, Ho CM. Unsteady separation in a boundary layer produced by an [77] Kwon DK, Kareem A. Gust-front factor: new framework for wind load effects
impinging jet. J Fluid Mech 1985;160:235–56. on structures. ASCE J Struct Eng 2009;135(6):717–32.
[42] Donaldson CD, Snedeker RS. A study of free jet impingement, Part 1. Mean [78] Ladubec C, El Damatty AA, El Ansary AM. Effect of geometric nonlinear
properties of free and impinging jets. J Fluid Mech 1971;45:281–319. behaviour of a guyed transmission tower under downburst loading. In:
[43] Durst CS. Wind speeds over short periods of time. Meteorol Mag Proceedings of the international conference on vibration, structural
1960;89:181–6. engineering and measurement (ICVSEM 2012), 19–21 October 2012,
[44] Dyke PV, Laneville A. Galloping of a single conductor covered with a D-section Shanghai, China. Trans Tech Publications; 2012. p. 1240–9.
on a high-voltage overhead test line. J Wind Eng Ind Aerodyn 2008;96: [79] Li CQ. Stochastic model of severe thunderstorms for transmission line design.
1141–51. Probab Eng Mech 2000;15(4):359–64.
[45] El Damatty A, Aboshosha H. Capacity of electrical transmission towers under [80] Li C, Li QS, Xiao YQ, Ou JP. A revised empirical model and CFD simulations for
downburst loading. In: Proceedings of the 1st Australasia and South-East Asia 3D axisymmetric steady-state flows of downbursts and impinging jets. J
structural engineering and construction conference, 28 November–2 Wind Eng Ind Aerodyn 2012;102:48–60.
December, Perth, Australia; 2012. p. 317–22. [81] Lin WE, Savory E, McIntyre RP, Vandelaar CS, King JPC. The response of an
[46] El Damatty A, Elawady A. Critical load cases for lattice transmission line overhead electrical power transmission line to two types of wind forcing. J
structures under downbursts. In: The 14th international conference on wind Wind Eng Ind Aerodyn 2012;100(1):58–69.
engineering, ICW14, June 21–25; 2015. [82] Lin WE, Orf LG, Savory E, Novacco C. Proposed large-scale modelling of the
[47] El Damatty A, Hamada A, Elawady A. Development of critical load cases transient features of a downburst outflow. Wind Struct 2007;10(4):315–46.
simulating the effect of downbursts and tornadoes on transmission line [83] Loredo-Souza AM, Davenport AG. The effects of high winds on transmission
structures. In: Proceedings of the 8th Asia-Pacific conference on wind lines. J Wind Eng Ind Aerodyn 1998;74–76:987–94.
engineering, 10–14 December, Chennai, India; 2013. [84] Loredo-Souza AM, Davenport AG. A novel approach for wind tunnel
[48] Fujita TT. ‘‘The downburst: microburst and macroburst”, SMRP research modelling of transmission lines. J Wind Eng Ind Aerodyn 2001;89:1017–29.
paper 210. USA: University of Chicago; 1985. [85] Loredo-Souza AM, Davenport AG. Wind tunnel aeroelastic studies on the
[49] Fujita TT. Downbursts: meteorological features and wind field characteristics. behaviour of two parallel cables. J Wind Eng Ind Aerodyn 2002;90:407–14.
J Wind Eng Ind Aerodyn 1990;36:75–86. [86] Loredo-Souza AM, Davenport AG. The influence of the design methodology in
[50] Gast KD, Schroeder JL. Supercell rear-flank downdraft as sampled in the 2002 the response of transmission towers to wind loading. J Wind Eng Ind Aerodyn
thunderstorm outflow experiment. In: Proceedings of the 11th international 2003;91(8):995–1005.
conference on wind engineering, Lubbock, Tex; 2003. p. 2233–40. [87] Lombardo FT, Smith DA, Schroeder JL, Mehta KC. Thunderstorm
[51] Gant SE. Reliability issues of LES-related approaches in an industrial context. characteristics of importance to wind engineering. J Wind Eng Ind Aerodyn
Flow Turbul Combust 2009;84:325–35. 2014;125:121–32.
[52] Gerges RR, El-Damatty AA. Large displacement analysis of curved beams. In: [88] Lou W, Sun B, Tang J. Aeroelastic model investigation and spectral analysis of
Proceedings of the Canadian society of civil engineering conference, 3–5 June, a tall lattice tower. Adv Struct Eng 2000;3(2):119–30.
Montreal, Canada, ST 100; 2002. [89] Lilien JL, Erpicum M, Wolfs M. Overhead line galloping. Field experience
[53] Gunter WS, Schroeder JL. ‘‘High-resolution full-scale measurements of during one event in Belgium on last February 13th, 1997. In: 8th int conf
thunderstorm outflow” winds. J Wind Eng Ind Aerodyn 2015;138:13–26. atmospheric icing of structures, Reykjavik, Iceland; 1998.
46 H. Aboshosha et al. / Engineering Structures 112 (2016) 23–46

[90] Macdonald JHG, Griffiths PJ, Curry BP. Galloping analysis of stranded [118] Shehata AY, El Damatty AA. Failure analysis of a transmission tower during a
electricity conductors in skew winds. Wind Struct 2008;11(4):303–21. microburst. Wind Struct 2008;11(3):193–208.
[91] Ma WY, Liu QK, Du XQ, Wei YY. Effect of the Reynolds number on the [119] Shehata AY, Nassef AO, El Damatty AA. A coupled finite element-optimization
aerodynamic forces and galloping instability of a cylinder with semi-elliptical technique to determine critical microburst parameters for transmission
cross sections. J Wind Eng Ind Aerodyn 2015;146:71–80. towers. Finite Elem Anal Des 2008;45(1):1–12.
[92] Manitoba Hydro. Bipole 1 & 2 HVDC transmission line wind storm failure on [120] Shehata AY, El Damatty AA, Savory E. Finite element modeling of
September 5, 1996 – review of emergency response, restoration and design of transmission line under downburst wind loading. Finite Elem Anal Des
these lines. Transmission and Civil Design Department, 98-L1/1-37010- 2005;42(1):71–89.
06000; 1999. 54 p. [121] Simiu E, Scanlan RH. Wind effects on structures. 3rd ed. New York: Wiley;
[93] Mara TG, Galsworthy JK, Savory E. Assessment of vertical wind loads on 1996.
lattice framework with application to thunderstorm winds. Wind Struct [122] Sissenwine N, Tattleman P, Granthan DD, Gringorten II. Extreme wind speeds,
2010;13(5):413–31. gustiness and variations with height. Air Force Cambridge Research
[94] Mara TG, Hong HP. Effect of wind direction on the response and capacity Laboratories. Technical report 73-0560. Aeronomic Laboratory, Project
surface of a transmission tower. Eng Struct 2013;57:493–501. 8624; 1972.
[95] Mason MS, Wood GS, Fletcher DF. Numerical simulation of downburst winds. [123] Solari G, Burlando M, De Gaetano P, Repetto MP. Characteristics of
J Wind Eng Ind Aerodyn 2009;97(11–12):523–39. thunderstorms relevant to the wind loading of structures. Wind Struct
[96] Mason MS, Fletcher DF, Wood GS. Numerical simulation of idealized three- 2015;20(6):763–91.
dimensional downburst wind fields. Eng Struct 2010;32(11):3558–70. [124] Solari G, De Gaetano P, Repetto MP. Thunderstorm response spectrum:
[97] Mason MS, Wood GS, Fletcher DF. Numerical investigation of the influence of fundamentals and case study. J Wind Eng Ind Aerodyn 2015;143:62–77.
topography on simulated downburst wind fields. J Wind Eng Ind Aerodyn [125] Solari G. Gust buffeting. I: peak wind velocity and equivalent pressure. J
2010;98(1):21–33. Struct Eng 1993;119(2):365–82.
[98] Mehta KC, Kadaba R, Kempner L. Conductor response to extreme winds. In: [126] Solari G. Gust buffeting. II: dynamic along-wind response. J Struct Eng
Proceedings. Second international symposium on probability methods 1993;119(2):383–97.
applied to electric power systems, September 20–23, 1988, Oakland, CA; [127] Sparling B, Wegner L. Estimating peak wind load effects in guyed masts.
1988. p. 9-1–9-11. Wind Struct 2007;10(4):347–66.
[99] Mehta KC, Kadaba R. Wind induced response of transmission line conductors. [128] Su Y, Huang G, Xu Y. Derivation of time-varying mean for non-stationary
In: Sun TF. Recent advances in wind engineering. Proceedings of the second downburst winds. J Wind Eng Ind Aerodyn 2015;141(2015):39–48.
Asia-Pacific symposium on wind engineering, Beijing, China. vol. 2. Pergamon [129] Vermeire BC, Orf LG, Savory E. Improved modelling of downburst outflows for
Press; 1989. p. 748–55. wind engineering applications using a cooling source approach. J Wind Eng
[100] Mehta KC, Kadaba R. Field data analysis of electrical conductor response to Ind Aerodyn 2011;99(8):801–14.
winds. J Wind Eng Ind Aerodyn 1990;36:329–38. [130] Vermeire BC, Orf LG, Savory E. A parametric study of downburst line near-
[101] Momomura Y, Marukawa H, Okamura T, Hongo E, Ohkuma T. Full-scale surface outflows. J Wind Eng Ind Aerodyn 2011;99(4):226–38.
measurements of wind-induced vibration of a transmission line system in a [131] Vicroy DD. ‘‘A simple, analytical, axisymmetric microburst model for
mountainous area. J Wind Eng Ind Aerodyn 1997;72:241–52. downdraft estimation” NASA-TM-104053. Hampton (VA, USA): NASA; 1991.
[102] Orf L, Oreskovic C, Savory E, Kantor E. Circumferential analysis of a simulated [132] Vicroy DD. Assessment of microburst models for downdraft estimation. J
three-dimensional downburst producing thunderstorm out flow. J Wind Eng Aircraft 1992;29:1043–8.
Ind Aerodyn 2014;135(2014):182–90. [133] Volpe H. Wind effects on conductors for span factors. Final project report, line
[103] Orf LG, Kantor E, Savory E. Simulation of a downburst-producing design section, division of facilities engineering. Portland (OR): Bonneville
thunderstorm using a very high-resolution three-dimensional cloud model. Power Administration; 1989.
J Wind Eng Ind Aerodyn 2012;104–106:547–57. [134] Wang L, McCullough M, Kareem I. A data-driven approach for simulation of
[104] Orf LG, Oreskovic C, Savory E. Circumferential analysis of simulated 3-D full-scale downburst wind speeds. J Wind Eng Ind Aerodyn 2013;123
downburst-producing thunderstorm outflows. In: Proceedings of the 6th (2013):171–90.
European and African conference on wind engineering, Cambridge, UK, July; [135] Wang X, Lou W, Li H, Chen Y. Wind-induced dynamic response of high-rise
2013. transmission tower under downburst wind load. J Zhejiang Univ 2009;43
[105] Orwig KD, Schroeder JL. Near-surface wind characteristics of extreme (8):1520–5.
thunderstorm outflows. J Wind Eng Ind Aerodyn 2007;95:565–84. [136] Wei P, Bingnan S, Jinchun T. A catenary element for the analysis of cable
[106] Oseguera RM, Bowles RL. ‘‘A simple, analytics 3-dimensional downburst structures. Appl Math Mech 1999;20(5):0253–4827 [English Edition].
model based on boundary layer stagnation flow” NASATM- 100632. Hampton [137] Wieringa J. Updating the Davenport roughness classification. J Wind Eng Ind
(VA): NASA Langley Research Center; 1988. Aerodyn 1992;41–44:357–68.
[107] Ohkuma T, Marukawa H. Galloping of overhead transmission lines in gusty [138] Wilson JW, Roberts RD, Kessinger C, McCarthy J. Microburst wind structures
wind. Wind Struct 2000;3(4):243–53. and evaluation of Doppler radar for airport wind shear detection. J Climate
[108] Peil U, Nölle H. Guyed masts under wind load. J Wind Eng Ind Aerodyn Appl Meteorol 1984;23:898–915.
1992;41–44:2129–40. [139] Winkelman PF. Sag-tension computations and field measurements of
[109] Rawlins CB. Analysis of conductor galloping field observations single Bonneville Power Administration. AIEE paper 59-900; 1959.
conductors. IEEE Trans Power Apparatus Syst 1981;PAS-100(8):3744–51. [140] Wolfson MM, DiStefano JT, Fujita TT. Low-altitude wind shear characteristics
[110] Savory E, Parke GAR, Disney P, Toy N. Wind-induced transmission tower in the Memphis, TN area. In: Preprints of the 14th conference on severe local
foundation loads: a field study-design code comparison. J Wind Eng Ind storms, October 29–November 1, Indianapolis, IN, USA. American
Aerodyn 2008;96:1103–10. Meteorological Society; 1985. p. 322–7.
[111] Savory E, Parke GAR, Zeinoddini M, Toy N, Disney P. Modelling of tornado and [141] Wood GS, Kwok KCS, Motteram NA, Fletcher DF. Physical and numerical
microburst-induced wind loading and failure of a lattice transmission tower. modelling of thunderstorm downbursts. J Wind Eng Ind Aerodyn
Eng Struct 2001;23(4):365–75. 2001;89:535–52.
[112] Selvam RP, Holmes JD. Numerical simulation of thunderstorm downdrafts. J [142] Yan Z, Li Z, Savory E, Lin W. Galloping of a single iced conductor based
Wind Eng Ind Aerodyn 1992;41–44:2817–25. on curved-beam theory. J Wind Eng Ind Aerodyn 2013;123(Part A):
[113] Sengupta A, Sarkar PP. Experimental measurement and numerical simulation 77–87.
of an impinging jet with application to thunderstorm microburst winds. J [143] Yasui H, Marukawa H, Momomura Y, Ohkuma T. Analytical study on wind-
Wind Eng Ind Aerodyn 2008;96:345–65. induced vibration of power transmission towers. J Wind Eng Ind Aerodyn
[114] Shan L. Characteristics study of transmission line mechanical research center 1999;83:431–41.
(TLMRC) wind tower data. Interim report no. EPRI TR-100906. Project 2016- [144] Yu P, Wong P, Kaempffer F. Tension of conductor under concentrated loads. J
03. Palo Alto (CA): Electric Power Research Institute; 1992. Appl Mech – Trans Am Soc Mech Eng 1995;62(3):802–9.
[115] Shan L. Overview of EPRI conductor wind loading experiments. In: [145] Zhang Y, Hu H, Sarkar PP. Modeling of microburst outflows using impinging
Proceedings of the seventh US national conference on wind engineering. jet and cooling source approaches and their comparison. Eng Struct
Los Angeles, California; 1993. 2013;56:779–93.
[116] Shan L. Evaluation of the results of several full-scale conductor wind loading [146] Zhou Y, Kareem A. Gust loading factor: new model. J Struct Eng 2001;127
experiments. EPRI TR-100479; 1994. (2):168–75.
[117] Shehata AY, El Damatty AA. Behaviour of guyed transmission line structures [147] Zhu S, Etkin B. Model of the wind field in a downburst. J Aircraft
under downburst wind loading. Wind Struct 2007;10(3):249–68. 1985;22:595–601.

You might also like