You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/313815664

Structural behaviour of lattice transmission towers subjected to wind load

Article  in  Structure and Infrastructure Engineering · February 2017


DOI: 10.1080/15732479.2017.1290120

CITATIONS READS

10 1,868

2 authors:

Edgar Tapia-Hernández Emilio Sordo


Metropolitan Autonomous University Metropolitan Autonomous University
73 PUBLICATIONS   218 CITATIONS    13 PUBLICATIONS   23 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nonlinear behavior of steel structures View project

Seismic behavior of Braced Steel Frames View project

All content following this page was uploaded by Edgar Tapia-Hernández on 11 December 2017.

The user has requested enhancement of the downloaded file.


Structure and Infrastructure Engineering
Maintenance, Management, Life-Cycle Design and Performance

ISSN: 1573-2479 (Print) 1744-8980 (Online) Journal homepage: http://www.tandfonline.com/loi/nsie20

Structural behaviour of lattice transmission towers


subjected to wind load

Tapia-Hernández Edgar & Emilio Sordo

To cite this article: Tapia-Hernández Edgar & Emilio Sordo (2017): Structural behaviour of
lattice transmission towers subjected to wind load, Structure and Infrastructure Engineering, DOI:
10.1080/15732479.2017.1290120

To link to this article: http://dx.doi.org/10.1080/15732479.2017.1290120

Published online: 16 Feb 2017.

Submit your article to this journal

Article views: 17

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=nsie20

Download by: [148.206.91.180] Date: 23 February 2017, At: 10:33


Structure and Infrastructure Engineering, 2017
http://dx.doi.org/10.1080/15732479.2017.1290120

Structural behaviour of lattice transmission towers subjected to wind load


Tapia-Hernández Edgara  and Emilio Sordob
a
Materials Department, Universidad Autónoma Metropolitana-Azcapotzalco, Mexico City, Mexico; bUniversidad Autónoma Metropolitana-Lerma,
Principal, Estado de México, Mexico

ABSTRACT ARTICLE HISTORY


Hurricane Wilma hit Yucatan Peninsula in 2005, causing substantial damage to local electrical transmission Received 2 June 2016
towers. Based on the failure characteristics observed on such towers, an analytical study is performed Revised 13 November 2016
to reproduce such failures and assess their vulnerability. Two latticed transmission towers are analysed Accepted 4 December 2016
under the action of 14 different wind velocity patterns corresponding to several national and international KEYWORDS
wind design codes. Displacement-controlled pushover analyses are performed to reproduce impending Wind loads; transmission
failure mechanism for considered wind patterns, and associated gradient wind speed is computed. Results towers; non-linear analyses;
illustrate that consistent cyclonic wind speed patterns lead to better estimates of failure mechanism and hurricanes; structural safety
gradient wind values than traditional non-cyclonic patterns.

1. Introduction of the supporting tower systems, on the basis of probabilistic


In recent years, increased intense hurricane activity has resulted approaches.
in damage to electric transmission lines (Webster, Holland, Codes have different approaches to the lateral wind speed
Curry, & Chang, 2005). These damages are usually controllable, patterns that must be applied to structural systems. Some
as they are considered in risk analyses of such structural systems. include explicit considerations for the load pattern related to
However, it is important to improve technical knowledge about cyclonic wind speeds, based on reported field measurements
this phenomenon to reduce economic losses and interruptions that point out differences from standard continental wind
in electric service from damage to transmission towers. Typhoon patterns (Amano, Fukushima, Ohkuma, Kawaguchi, & Goto,
Number 21 in 2002 and Hurricanes Wilma, Rita and Katrina in 1999; Shanmugasundaram, Harikrishna, Gomathinayagam, &
2005, Ike and Gustav in 2008 and Earl and Alex in 2010 are just a Lakshmanan, 1999). The relevance of these observations on the
few examples (Figure 1). Vulnerability of high-voltage transmis- safety of designs that do not consider explicitly cyclonic or con-
sion lines and local distribution lines to severe natural hazards is tinental winds is an issue still under discussion. However, despite
therefore a major concern for designers, owners and managers modern wind hazard distribution studies for local code intensity
of these facilities. levels, transmission line failures are still occurring. Although
High-voltage transmission towers have a distinct structural international concern has been focused on downburst effects
behaviour under wind loads, as they are typically designed con- to transmission towers (Shehata, El Damatty, & Savory, 2005),
sidering relatively small safety margins, despite their high expo- failures of Mexican transmission line systems are mainly due
sure to intense winds due to the large transmission line extension to high intensity cyclonic winds (Muñoz, Sánchez, López, &
(Holmes, 2007). In addition, behaviour of transmission towers Pérez-Rocha, 2009) which cause important damages and losses
under such action can be complex to assess due to different fac- to transmission structures lines, as shown in Figure 1.
tors as connections (Jiang, Wang, McClure, Wang, & Geng, 2011) Near-failure performance under critical wind loading for
or cable supporting systems, among others. Inelastic response of transmission towers are sometimes assessed through wind tunnel
conductors and their specific vibration modes may have also an testing (e.g. Mara, 2013; Liang, Zou, Wang, & Cao, 2015; Yang,
important effect on the overall behaviour (Kaminski, Riera, de Yang, Niu, & Zhang, 2015). Since test results are only valid for
Menezes, & Miguel, 2008). Thus, specific manuals for the analysis a particular tower and loading conditions, they may not accu-
and design of these structural systems have been developed as rately predict tower behaviour under critical onsite wind loading
ASCE Manual of Practice No. 74 (ASCE 74, 1991) or the National conditions. It is known, however, that structural behaviour of
Electric Safety Code IEEE C2 (IEEE, 2007), which are commonly an electrical transmission tower may depend on wind loading
referenced by international codes for wind design. Given the source characteristics, as Savory, Parke, Zeinoddini, Toy, and
relative complexity of these manuals and the broad range of site Disney (2001) and Banik, Hong, and Kopp (2008) pointed out.
conditions where transmission lines have to be supported on, In addition, results from the scarce studies on tower behaviour
current codes establish simple criteria for the structural design under strong winds typically focus on the behaviour of mid-rise

CONTACT  Tapia-Hernández Edgar  etapiah@azc.uam.mx


© 2017 Informa UK Limited, trading as Taylor & Francis Group
2   E. TAPIA-HERNÁNDEZ AND E. SORDO

Figure 1. Damage in transmission towers caused by the Hurricane Wilma in Yucatan, Mexico.

Figure 2. Tension tower 53T10 and its typical connections (a) Global view, (b) Typical connection and (c) Typical support.

towers, i.e. 20 to 30 m in height (Banik et al., 2008; Lin et al., codes for lattice transmission towers contain only limited advice
2011; Jiang et al.,, 2011). Capacity curves based on non-linear, on the treatment of high-intensity wind effects (hurricanes) and
load-controlled static pushover analyses for five different static structural design is carried out using wind load profiles and
load patterns were obtained recently by Banik, Hong, and Kopp response factors derived for atmospheric boundary layer winds
(2010) for a 25-m-tall lattice tower. Although no consideration based on elastic response (Mara, 2013). In this regard, some
was made in Banik’s procedure to account for P − Δ effects, his references (e.g. CIGRÉ, 2008, 2012) suggest more load cases in
work pointed to a strong capacity dependence on the specific order to evaluate the response and to mitigate the damage of
adopted load pattern, particularly among seismic loading pat- transmission towers under severe wind demands.
terns and wind design loading ones. Hurricane winds are not equivalent to the synoptic wind
There is then need to study critical performance of structural events characterised by typical atmospheric boundary layer
systems when subjected to natural intense loading conditions. winds since the passage of the eye of a hurricane through a tower
In this regard, hurricane Wilma hit Yucatan peninsula in 2005, location is characterised by increasing wind speeds followed by
being the most powerful hurricane ever recorded in the area. In a relatively short period of calm winds, after which the wind
the span of just 24 h, Wilma had intensified from 111 kph tropical velocity increases again to a second maximum before gradually
storm to a 278 kph category 5 hurricane, an unprecedented event decreasing as the hurricane moves away from the tower. So, a var-
for an Atlantic tropical cyclone. Wilma reached a peak sustained iation of the mean wind distribution over time is usually reported
wind speed of  kph with an estimated minimum central pres- in exposure measurements. For example, the maximum observed
sure of 88.2 kPa. Sustained surface wind speed over Cancun area 15-min speeds at the 10-m level varied between 23.0 and 36 m/s,
reached 160 kph, getting to gust values up to 209 kph, as Pasch, while the maximum 60 seg wind speeds varied between 29.2 and
Blake, Cobb, and Roberts (2006) report, among other detailed 41.9 m/s during the monitoring of hurricanes Katrina, Rita and
characteristics of Wilma Hurricane. It is estimated 23 million Wilma in 2005 (Masters et al.,, 2010).
dollar cost of damages to electrical sector in Quintana Roo State This paper presents a comparative overview of four inter-
from the effects of Wilma (CENAPRED-CEPAL, 2007). national codes on their application to the analysis and design
While numerous field experiments have been conducted to of transmission towers under the action of hurricane winds.
characterise wind in neutral conditions, the literature is scarce in In particular, three international regulations are evaluated;
addressing surface-level winds occurring over land in hurricanes North-American ASCE SEI 7 (ASCE 7, 2010), Australian/New
(Masters, Tieleman, & Balderrama, 2010). In fact, current design Zealand 1170.2 (AS/NZS, 2011) and two Mexican Codes: (a) the
STRUCTURE AND INFRASTRUCTURE ENGINEERING   3

(a) (b)
Figure 3. Typical information of the studied lattice transmission towers (a) Transverse load H due to the tension and definition of the line angle Δ and (b) Configuration
of 53T10 tower.

Wind Design Chapter of the Civil Engineering Manual of the are standard towers and typologies commonly used in Mexico
Mexican Federal Electricity Commission (CFE- 2008) and (b) by the Federal Commission of Electricity.
the Guidelines of Mexico’s Federal District Code (NTCV, 2004). Following typical design specifications, a 400-m span was
Non-linear static pushover analyses, which are routinely used for considered (Figure 4). Conductors were assumed to be ACSR
characterisation of capacity curves for structures under seismic 1113 (Bluejay) with a rated strength in tension equal to 36.28
excitations, were employed here to estimate the impending fail- kN (3700 kg) and 18.34 N/m (1.87 kg/m), and the wire guard
ure behaviour of transmission towers. Non-linear analyses are with a yielding tension capacity equal to 9.32 kN (950 kg) with
performed for 14 load patterns derived from these codes (12 3.98  N/m (0.406  kg/m). Weight of the insulator string was
for continental winds and 2 for cyclonic regions) to determine assumed to be 3.08 (314 kg) (Tapia & Valdepeña, 2002). Wind
wind speed associated to impending failure mechanism of two load calculation considers terrain effects, structure height, wind
high-voltage transmission tower models. Differences on code gust and structure shape, under three main load conditions;
criteria and code wind speed gradients are also discussed as a longitudinal (parallel to the line) wind acting on insulators and
result of the study. tower structure, transverse (perpendicular to the line) wind
acting on conductors, insulators and tower structure together
with the transverse component of the line tension and a broken
2.  Models description
conductor wire condition.
Two 400 kV high-voltage transmission lattice towers are studied. Models of transmission line towers were analysed using
These 53-m height towers are structured with equal-legged angle the commercial software SAP2000 (CSI, 2010) by linear static
high resistance commercial A-572 steel sections and ASTM A394 analysis under: (a) self-weight of cables, isolators and tower, (b)
bolts (Figure 2). These are typical configurations of Mexican lat- mechanical tension in cables due the line deviation and (c) wind
tice towers. They support an overhead power line on two cross- loads acting on cables and tower body. In models, each member
arms carrying a three-phase electric circuit each, and two extra of the modelled structure is being idealised as a three-dimen-
cross-arms at the tower’s top for the ground wires (guard wires). sional steel frame element, where P-Delta effects were included
Maximum deviator line angle (Figure 3a) is Δ = 10° for model in order to account the second-order effect due the deforma-
53T10 with 5.92 m × 5.92 m square base area and Δ = 60° for tion of the structure produced by gravitational and lateral forces.
model 53T60 with 8.36 m × 8.36 m square base area. Both models According to the results, the structures fulfil all the requirements
4   E. TAPIA-HERNÁNDEZ AND E. SORDO

52.53m Legs 52.53m Legs


5.45 1.5 Braces 5.45 1.5 8.10 Braces

8.0 m

8.0 m
L3"x1/4" L3"x1/4"
44.53m 5.45 L2"x1/4" 44.53m 5.45 8.10 L2"x1/4"

L4"x1/4" 2L3"x1/4"

8.5 m

8.5 m
36.03m 5.45 L3"x1/4"
36.03m 5.45 8.10

L4"x3/8"
8.5 m

8.5 m
8.10 L4"x1/4"
27.53m 5.45 2L4x1/4" 27.53m 5.45
52.53 m

2L4"x3/8"

2L4"x3/8"
L4"x1/4"
L4"x3/8"
27.53 m

27.53 m
2L4"x1/2"
5.00m 2L4"x1/2" 5.00m 2L3"x1/4"

0.00m L3"x1/4" 0.00m L4"x1/4"

5.92 m 8.36 m

(a) (b)
Figure 4. Cross sections of studied towers (a) Model 53T10 and (b) Model 53T60.

(a) Undeflected (b) Mode 1 (c) Mode 2 (d) Mode 3


shape (T1 = 0.79 Sec.) (T2 = 0.79 Sec). (T3 = 0.43 Sec.)
Parallel to line Normal to line Torsional
direction direction

Figure 5. Modal configuration without conductors and main periods of the tower 53T10.

of the wind and steel guidelines of the Mexican Code, where the applied periodic loading will coincide with a modal response
maximum demand is linked to the wind patterns perpendicular and hence cause resonance, which leads to large oscillations.
to the line direction. Characteristics of first three modes of studied towers are shown
In tower analysis, it is useful to determine the modes’ shape in Table 1. Since towers have square cross section and identical
and natural frequencies (which depend on the mass and stiff- element cross sections (Figures 3a, 4), it is not surprising that
ness distributions) as it allows knowing if the frequency of any the first two modes are translational with a significantly similar
STRUCTURE AND INFRASTRUCTURE ENGINEERING   5

(a) (b)

Figure 6. Structural elements contribution to total solid area (a) Tower 53T10 and (b) Tower 53T60.

(a) (b)

Figure 7. Solidity Ratio ϕ (a) Tower 53T10 and (b) Tower 53T60.

Figure 8. Drag force coefficient values.

magnitude. These results (frequencies) are in agreement to other contribution from those member types to the total solid area
similar analytical studies (Savory et al.,, 2001) and ambient vibra- is noted. Areas were carefully computed from the construction
tion measurements (Yin, Lam, Chow, & Zhu, 2009). Figure 5 drawings. From these results, solidity ratios ϕ (effective solid
illustrates the schematic mode configurations of tower 53T10. area on which wind acts divided by the total area enclosed by
The reported natural periods in Table 1 do not consider here cou- the perimeter of the exposed surface) were calculated and sum-
pling with the suspended conductors. In this regard, according marised in Figure 7. Lateral nodal forces were calculated from
to Pengyun et al., (2012), since the conductors might have an solidity ratios ϕ.
important contribution to the reactive mass in tension towers, It can be observed in Figure 7 that tower solidity ratio values
the dynamic properties are not accurately assessed in a single do not vary much along the tower height. The two towers have
tower in comparison with those of coupled tower line systems an average value of around ϕ  ≈ 0.207. In the analyses of new
when the response perpendicular to the line is evaluated. As for towers, this magnitude might be useful in the initial estimation
the parallel to the line vibration, the effects of power line on the of the wind forces at the design stage. Since drag coefficient
dynamic properties of the tower are relatively small. is roughly constant, code effective–load distributions can be
Solid area graphs for leg and diagonal members as a ratio applied directly through a static structural analysis to determine
to the total solid area are depicted in Figure 6, where a similar member forces at impending collapse (Holmes, 1995). Values for
6   E. TAPIA-HERNÁNDEZ AND E. SORDO

(a) (b)

(c) (d)

Figure 9. Design V(z) normalised to the site wind speed VR, for areas having flat or open terrain (a) AS/NZS- 2011, (b) ASCE 7- 2010, (c) CFE- 2008 and (d) NTCV- 2004.

Figure 10. Wind speed patterns for analysed codes up to tower height.

the drag force coefficient Cd are depicted in Figure 8 for the range varied with the member spacing ratio; so, the solidity ratio ϕ
of computed values of the solidity ratio ϕ shown in Figure 7, alone may not be sufficient for describing the drag coefficient of a
which include the Japanese Code (AIJ, 2006), the Canadian code lattice frame having a non-uniform solid area distribution (Mara,
(CSA, 2010), the U.S. code (ASCE 7, 2010), the Australian/New 2014). In Figure 8, similar values for all codes can be noted, with
Zealand (AS/NZS, 2011), the International Electro-technical a maximum difference of 6% at ϕ ≈ 0.2 which is precisely the
Commission (IEC, 2003), the Manual of the Mexican Federal average solidity ratio of this type of structural systems, as it was
Electricity Commission (CFE- 2008, 2008), the Guidelines of corroborated in this study (Figure 7).
Mexico’s Federal District Code (NTCV, 2004) and results of wind It is worth mentioning that the experimental drag coefficient
tunnel tests (Bayar, 1986; Carril, Isyumov, & Reyolando, 2003; for towers in wind tunnel test is noticeably lower than would be
Mara & Hong, 2013). estimated using codes (Tapia-Hernández & Cervantes-Castillo,
According to design codes, the recommended values of the 2016). The difference may be attributed to the fact that the expres-
drag force coefficient Cd for generic lattice frame geometries sions in codes are not intended for this type of geometry (four-
(Figure 7) are based on the solidity ratio ϕ. Nevertheless, it was sided pyramid), and that the distribution of solid area within
observed in wind tunnel tests that the drag force coefficient Cd the frame is not uniform (Mara, 2013). Based on these results,
STRUCTURE AND INFRASTRUCTURE ENGINEERING   7

Figure 11. Wind forces acting on cables and tower body; Tower 53T60 (a) Flat, unobstructed terrain conditions and (b) Open terrain with scattered obstructions.

the values given in the CSA- 2010 and IEC- 2003 may be more
appropriate for lattice towers, with exception of 0.18 < ϕ < 0.30,
where the lower limit is defined by CFE- 2008, NTCV- 2004
and AS/NZS- 2011. It is important noting that differences have
been reported between the results of wind–structure interactions
recorded in field, tunnel test and computational fluid dynamics
analyses (e.g. Momomura, Marukawa, & Ohkuma, 1992), where
typically the effective drag forces on conductors are reduced.

3.  Loading patterns


Figure 12.  Cyclonic and non-cyclonic patterns acting on cables and tower body Wind load patterns for transmission-latticed towers as given
according to AS/NZS- 2011; tower 53T60. by three international codes are employed for the structural

Figure 13. Total load and overturning moment for VR = 100 kph by code (tower 53T60) (a) Wind Total Drag load (VR = 100 km/hr) and (b) Overturning moment (VR = 100 km/
hr).
8   E. TAPIA-HERNÁNDEZ AND E. SORDO

(a) (b)

Figure 14. Pushover curves for an element L4″ × 3/8″ (a) Comparative curves and (b) Modelling details in OpenSees.

pushover analysis: Mexican Federal Electricity Commission


Wind Design Manual (CFE- 2008); American Society of Civil
Engineers Structural Engineering Institute Code ASCE/SEI 7
(ASCE 7, 2010) and Australian/New Zealand 1170.2 Standards
(AS/NZS, 2011), which is largely based on the format set out in
ISO 4354 (ISO 4354, 2009) adding provision for different types
of wind, such as synoptic, thunderstorm, downburst and cyclone
(AS/NZS Supp, 2011). All these codes consider wind solicitation
as a vertical pattern of horizontal effective loading, as given by
Equation (1), which is based on the relationship between pres-
sure and atmospheric flow velocity from the Bernoulli’s Equation
for non-aerodynamic bodies (Holmes, 2007):
1
F(z) = 𝜌V (z)2 GCd Af (1)
2
Here, F(z) is the total force at height z over the projected sur-
face area Af normal to the wind direction due to wind speed
at the same height V(z). Cd is the drag force coefficient which
accounts for the tower shape and structural elements interfer-
ence. G is the gust effect factor which accounts for additional
dynamic amplification of loading in the wind direction due to
Figure 15.  Failure mechanism developed at studied towers (a) Tower 53T10; all
load patterns and (b) Tower 53T60; all load patterns. wind turbulence and structure interaction. ρ is the air density at

Figure 16. Collapse of 53T10 tower in Mexico (a) Global view and (b) Lower elements.
STRUCTURE AND INFRASTRUCTURE ENGINEERING   9

both extra tropical and tropical storm winds (Fleck-Fadel, Fleck-


Fadel, Riera, Kaminski, & Ramos, 2012). As the studied towers
have natural frequencies greater than 1 Hz (Table 1), dynamic
resonant contribution is neglected.

3.1.  Boundary layer


The most significant difference among codes lies in the wind
speed variation at the boundary layer. The variation of V(z)
Figure 17. Overstrength definition. normalised to the site wind speed VR (3 s gust speed at 10 m
above ground with scattered obstructions) is given by parameters
Table 1. Main periods of the studied towers.
depicted in Table 2. In AS/NZS- 2011, the definitions of reference
pressure and exposure factor for different types of wind (such as
Mode Tower 53T10 (s) Tower 53T60 (s) Main direction synoptic, thunderstorm, downburst or cyclone) are incorporated
1 0.79 0.73 Translational in line into the definition of the exposure factor Mz,cat (AS/NZS Supp,
direction
2 0.79 0.7 Translational perpendicu- 2011).
lar to line direction In Table 3, the terrain category definition for areas having flat,
3 0.43 0.49 Rotational open terrain and suburban areas is shown, whereas in Figure 9,
the variations are illustrated. Despite that load patterns for areas
Table 2. Code parameters for wind velocity vertical pattern.
of suburban housing were calculated and they are included in the
results of the non-linear analyses, the graphs focus their attention
Code Parameter Definition only on flat and open terrain conditions since these categories
AS/NZS- 2011 Mz,cat Mz,cat = Terrain/height multiplier
ASCE 7- 2010

Kz =  Velocity pressure exposure coefficient
are more related with cyclonic regions and this will be useful
Kz
CFE- 2008 Frz Frz = Local exposure factor for the following discussion. U.S. and Australian/New Zealand
NTCV- 2004 Fα Fα = height variation factor codes establish a wind speed gradient to site wind speed ratio
of V(z)/VR ≈ 1.40 (Figure 9a and b), regardless of its cyclonic
or synoptic genesis, which is somehow consistent with Sparks’
Table 3. Terrain category definition for analysed codes.
(2003) conclusions. On the other hand, CFE- 2008 establishes
AS/NZS- ASCE 7- NTCV- V(z)/VR ≈ 1.56 (Figure 9c).
Terrain Herein 2011 2010 CFE- 2008 2004 It is worth noting that a physical inconsistency of the wind
Flat, unob- Flat Category 1 Category Category 1 Category 1 profile is noticed in Figure 9d (NTCV- 2004) because the wind
structed D
areas, velocity gradient is not adjusted to the same magnitude regard-
water less the terrain categories. For this reason, this code was not
surfaces included in the following discussion. Table 4 shows normalised
Open ter- Open Category 2 Category C Category 2 Category 2
rain with wind velocity values up to the tower’s height. The corresponding
scattered best fit for mean power law exponent associated to these values is
obstruc- shown in Table 5, where significant differences among codes are
tions
Areas of Suburban Category 3 Category B Category 3 Category 3 observed for this exponent at synoptic values. However, AS/NZS-
suburban 2011 hurricane-specific expression has a very good agreement
housing with the 0.16 value recommended by Fu, Li, Wu, Xiao, and Song
(2008) for a typical open country terrain during the passage of a
tropical storm on the basis of field measurements.
Table 4. Variation of normalised wind velocity profile factor with height for differ- Code wind speed patterns up to the height of studied tow-
ent codes. ers are also graphed together in Figure 10. A great similarity
AS/NZS- between ASCE 7- 2010 and AS/NZS- 2011 for open, non-cy-
2011 clonic wind patterns can here be readily observed. It is important
(Hurri- AS/NZS- 2011
cane) (Synoptic) ASCE 7- 2010 CFE- 2008 to note that the AS/NZS criterion for cyclonic and non-cyclonic
events is consistent with the evidence found on the shorten-
Height z Open
(m) and Flat Open Flat Open Flat Open Flat ing of the boundary layer in these events (Amano et al.,, 1999;
10 1.00 1.00 1.12 1.00 1.09 1.00 1.14 Shanmugasundaram et al.,, 1999). It is also observed that wind
15 1.07 1.05 1.16 1.04 1.13 1.05 1.18 patterns for Mexican Code (CFE- 2008) are quite conservative
20 1.13 1.08 1.19 1.08 1.15 1.09 1.22 for heights lower than 10 m.
30 1.20 1.12 1.22 1.12 1.20 1.15 1.27
40 1.25 1.16 1.24 1.16 1.23 1.19 1.30
50 1.29 1.18 1.25 1.19 1.25 1.23 1.33
3.2.  Dynamic response factor
It is recognised that the resonant dynamic response is usually
sea level and 25 °C, taken here as ρ = 1.23 kg/m3, as suggested not a major problem in these structural systems and that the
by Kasperski (2009). These models constitutes a reasonable potential structural resonance of cable elements is cancelled by
approximation for extended pressure systems, characteristic of the typical high aerodynamic damping (Holmes, 2008; Lin et
10   E. TAPIA-HERNÁNDEZ AND E. SORDO

Table 5. Best normalised wind velocity parameters fit for mean power law expo- tower body for model 53T60. According to the formulations, in
nent.
order to be conservative, the exposure factor is constant until a
Hurri- height above ground zmin. The height zmin varies from one code to
cane the other (Table 5) and therefore modifies the initial magnitude
specific Synoptic, general
of the load patterns. Thus, the most significant differences in load
AS/
Best fit Exponent α AS/NZS- NZS- ASCE 7- CFE- 𝛼max patterns are detected at the base of the tower.
for (z∕10m)𝛼 2011 2011 2010 2008 𝛼
min Comparison of AS/NZS- 2011 load patterns for cyclonic and
Exponent α Open 0.158 0.102 0.105 0.128 1.25 non-cyclonic regions for unobstructed terrain conditions acting
Flat 0.158 0.069 0.087 0.099 1.43 on cables and tower body is illustrated in Figure 12. Again, they
Factor c Open 1.00 1.00 1.00 1.00 1.00
Flat 1.00 1.13 1.09 1.14 1.05
are similar, exception made at the base of the tower. This result is
Height Open 3.0 m 5.0 m 2.13 m (7 ft) 10.0 m somehow consistent with the observation made by Sparks (2003)
above Flat 3.0 m 4.57 m(15ft) in the sense that boundary layer characteristics of cyclonic winds
ground
zmin
do not substantially affect the practice of common wind engi-
neering. Total drag shear force FD and its resultant vertical loca-
tion h/H for a site wind speed equal to VR = 100 kph are depicted
Table 6. Code parameters for gust dynamic effects. in Table 7. The location h/H is a reference of the position in
height of the total drag force FD in order to obtain an equivalent
Code Parameter (value for towers studied herein)
overturning moment.
AS/NZS- 2011 Cdyn (= 1.0) Cdyn = dynamic response
factor It can be noted in Table 7 that Mexican Code (CFE- 2008)
ASCE 7- 2010 G (= 0.85) G = gust effect factor, rigid leads to the greatest drag forces in the unobstructed terrain con-
structure (>1.0 Hz) dition, whereas ASCE 7- 2010 leads to the smallest. AS/NZS- 2011
CFE- 2008 FAD (= 1.0*) FAD = dynamic amplification
factor leads to the same force magnitude for unobstructed and scattered
obstructed terrains for cyclonic winds, representing more than
*Although CFE- 2008 introduces a limiting aspect ratio not satisfied by the studied
cases in order to neglect dynamic effects, for code comparison purposes.
110% of US Code (ASCE 7- 2010) for same terrain conditions. It
is worth noting the importance of the wind pressure on cables,
which represents around 44% of the total wind drag force for
al., 2011). Consistently, dynamic effects related to this factor are tower 53T10 and around 40% for tower 53T60. According to these
typically one since structures have natural frequencies greater results, the drag force can be located at the middle height (h/H
than 1 Hz (Table 1). It is worth noting that ASCE 7- 2010 gust ≈ 0.5) in order to have a good approximation of the overturning
effect factor is even lower than 1.0 (Table 6), recognising that moment in both studied towers. This observation does not depend
the design wind already includes the effects of gustiness (Solari on the line angle Δ, solidity ratio ϕ or terrain condition.
& Kareem, 1998). Additionally, wind gust and hence local wind Additionally, total drag forces and overturning moments for a
forces on an overhead line do not peak simultaneously over a site wind velocity equal to VR = 100 kph are shown in Figure 13a
long horizontal distance normal to the mean wind direction. and b, respectively, for flat unobstructed terrain and open terrain
Namely, turbulent eddies in strong wind flow have finite sizes with scattered obstructions. It is worth noting the importance of
and will not envelope a complete transmission line span (Holmes, the wind pressure on cables, which represents 40% of the total
2008). Then, for simplicity, dynamic effects of cable elements are force and around 50% of the total overturning moment.
also ignored in this study, although it is known that these effects
and the cables fixing systems (for suspension and tension tow-
4.  Non-linear static pushover analysis
ers) may have an influence on the ultimate strength of the tower
(Kaminski et al.,, 2008). Non-linear, step-by-step, displacement-controlled pushover
Because CFE- 2008 and ASCE 7- 2010 do not include an analyses were carried out following the code-based horizontal
explicit methodology for helically wound unwrapped cables of force vertical patterns. The non-linear static pushover analysis
transmission towers, the drag factor coefficients for cables were method is a popular and routinely used method for seismic per-
computed following the criteria of AS/NZS- 2011 (Equation (2)), formance evaluation of structures by monotonically increasing
where Vdes,θ is the tower orthogonal design wind speed and b is the lateral loads with a predefined lateral load distribution.
the outside diameter equal to b  =  3.2  cm for conductors and Studies indicate that non-linear static analysis is quite superior to
b = 0.95 cm for ground wires. For values of bVdes,θ between 0.5 conventional linear static analysis with truss elements for under-
and 5.0, a linear interpolation was considered: standing the behaviour, load-carrying capacity and structural
stability (Prasad, Knight, Mohan, & Lakshmanan, 2012).
CD = 1.2 For bVdes, < 0.5 (2) Pushover analyses were carried out in SAP2000 (CSI, 2010)
in order to evaluate the non-linear behaviour of the two ana-
CD = 1.0 For bVdes, > 5.0
lysed towers under the action of the code load patterns. Bracing
elements were modelled with a yield in tension and elastic buck-
3.3.  Load patterns
ling in compression. Although the SAP2000 program does not
Wind loads acting on tower structure and cable elements are include a particular good modelling option for simulating the
calculated according to code criteria for flat unobstructed terrain cyclic behaviour of braces buckling in compression, the elastic
and open terrain with scattered obstructions. Figures 11 illus- buckling modelling option is a reasonable approximation for
trates flat and open terrain load patterns acting on cables and pushover analyses when global parameters are assessed, as shown
STRUCTURE AND INFRASTRUCTURE ENGINEERING   11

Table 7. Total drag shear force and vertical location for a site wind speed of 100 km/hr.

AS/NZS- 2011
Cyclonic Synoptic ASCE 7- 2010 CFE- 2008
Open-Flat Flat Open Flat Open Flat Open
Tower 53T10
Total drag force 0.48 0.55 0.51 0.51 0.51 0.51 0.51
location (h/H)
Total drag force 14.1 14.7 12.3 12.5 11.0 16.1 13.4
FD(ton)
% Tower 54 56 55 57 56 56 55
% Cables 46 44 45 43 44 44 45

Tower 53T60
Total drag force 0.54 0.47 0.50 0.50 0.50 0.50 0.50
location (h/H)
Total drag force 15.4 16.3 13.6 13.8 12.1 17.8 14.7
FD(ton)
% Tower 59 60 60 61 60 60 59
% Cables 41 40 40 39 40 40 41

in Figure 14a. In this figure, axial load normalised with the plas- Roeder, 2013; Tapia-Hernández & Tena-Colunga, 2014; Uriz
tification capacity in compression against the strain deformation & Mahin, 2008). In these analyses, load wind patterns on the
rate (ε = δ/L) obtained originally with SAP2000 is compared tower, insulators and cables were considered as static incremental
to the curve obtained in a more detailed model in OpenSees loadings, whereas weight of the structure, conductors and insu-
(Mazzoni, McKenna, Scott, & Fenves, 2006). lators and the component of the line tension remained constant.
The comparative study of axially loaded elements was carried Structural members are modelled as beam column elements and
out in 3D for the element L4″ × 3/8″ with h = 1.50 m, which is a plastic hinge post-yielding behaviour was considered through
typical cross section and length used in both towers (Figure 4). a simple axial force and bending moment interaction model.
In OpenSees, two rectangular patches were used to generate the Two failure mechanisms (Figure 15) were induced by the ine-
cross section of angles: one for each leg. Patches were discretised lastic pushover analyses, one for each of the two studied towers.
into fibres with quadrilateral shapes and four integration points Mechanism was shown to be independent of the specific loading
per element (Figure 14b). Member was modelled with a set of pattern. Failure is defined here as the point in the analyses where
10 non-linear beam column elements to reproduce the response there is no convergence for an increased lateral load. In fact, the
of an axially loaded element. Further details of the modelling final mechanism of tower 53T10 is well related with physical
technique of angles axially loaded using OpenSees can be found damage in towers as the one shown in Figure 16. This structure’s
in Tapia-Hernández, Ibarra-González, and De-León (in press). collapse occurred in grassland terrain with few scattered obstruc-
It can be observed that, for practical purposes, the approxi- tions at night in the fall 2006 in the Puebla, Mexico. As can be
mations obtained with the SAP2000 modelling are reasonable observed, the damage was concentrated only in the elements
enough to asses global design parameters when performing push- connected to the foundation.
over analyses. Nevertheless, different results could be obtained Total wind forces associated to impending failure mechanism
when the hysteretic behaviour is assessed (Hsiao, Lehman, & for the considered code load patterns are shown in Table 8 for
Table 8. Total drag shear force and vertical location for a site wind speed of 100 km/hr.

AS/NZS- 2011
Cyclonic Synoptic ASCE 7- 2010 CFE- 2008
Open
and Flat Suburban Flat Open Suburban Flat Open Suburban Flat Open Suburban
Tower 53T10
Wind design Vel at failure 178 185 177 192 216 192 203 227 168 184 202
(km/hr)
Drag force at failure (ton) 44.2 43.2 46.2 45.5 44.5 45.9 45.4 44.1 45.7 45.1 44.6
Drag force overstrength, 3.16 3.42 3.13 3.69 3.16 3.67 4.13 5.15 2.83 3.37 4.06
Ω
Safety factor 1.27 1.32 1.26 1.37 1.54 1.37 1.45 1.62 1.20 1.31 1.44

Tower 53T60
Wind design Vel at failure 178 178 177 192 190 192 202 227 168 183 202
(km/hr)
Drag force at failure (ton) 48.5 48.5 50.5 50.0 47.0 50.5 50.0 49.0 50.0 49.0 49.0
Drag force overstrength, 3.15 3.15 3.10 3.67 3.40 3.66 4.12 5.21 2.81 3.37 4.07
Ω
Safety factor 1.27 1.27 1.26 1.37 1.35 1.37 1.44 1.62 1.20 1.30 1.44
12   E. TAPIA-HERNÁNDEZ AND E. SORDO

flat, open terrain and suburban areas. According to these results, • Wind speed patterns for Mexican codes (NTCV- 2004
the most conservative design criterion is that of CFE- 2008 for and CFE- 2008) are identical for the terrain with scat-
flat terrain, as the tower’s failure is related to the lowest wind tered obstruction and suburban areas. However, the wind
velocity. In contrast, maximum site speed wind VR resulted from pattern is completely different for unobstructed terrain
ASCE 7- 2010 load patterns for open and suburban terrains. condition. Both codes have a conservative criterion for
With exception of the CFE- 2008 criterion for suburban areas, heights lower than 10 m regardless the terrain condition.
the obtained overstrength ratio is always bigger than   >  3.0 • A physical inconsistency of the wind profile suggested by
for both towers, where the overstrength was defined according NTCV- 2004 is noticed because the wind velocity gradient
to Figure 17. In the figure, Fd is the nominal design force, Fy is is not adjusted to the same magnitude regardless the ter-
the yielding force and Fmax is the maximum force resisted by the rain categories. In addition, this is the only code that leads
structure. Note that because the design for wind actions requires to a lower force for an unobstructed terrain condition than
the structures to remain in the elastic range (Banik et al.,, 2008), the one obtained for a terrain with scattered obstructions.
it is expected that Fd < Fy in a rational analysis. • The total drag force FD can be located at the middle height
There is no clear tendency between results of both towers, (h/H ≈ 0.5) in order to have a good approximation of the
which probably means that in the original design they were cal- overturning moment. This observation does not depend
culated under the same wind velocity and, therefore, the line on the deflection angle Δ, solidity ratio ϕ or terrain
angle  (10° and 60°) does not substantially affect the wind speed condition.
that leads to impending failure. Finally, according to the Wind • According to the results, the ASCE 7- 2010 is related with
Chapter of Civil Engineering Manual (CFE- 2008), the site wind the highest overstrength factor Ω. This observation has a
design velocity for the studied towers is 140.4 kph. Then, a safety strong relationship with the fact that this code is also related
factor was computed for the scenarios studied in this research with the maximum wind velocity proposed between the
(Table 8). These results represent the risk of collapse of the tower compared codes the for all the terrain conditions.
under intense wind loads. • The line angle  of the studied lattice towers does not sub-
stantially affect the wind speed that leads to impending
5. Conclusions failure.
• In general, there is a good correspondence among the
Two existing 400  kV high-voltage transmission lattice towers
compared codes.
53 m in height were non-linearly analysed under the action of
load patterns from four wind design regulations: two Mexican Additional analyses of existing lattice transmission towers,
(CFE- 2008, 2008 and NTCV- 2004), the US code (ASCE 7, 2010) which collapse under intense wind loads, and applying different
and the Australian/New Zealand Code (AS/NZS- 2011) for three numerical codes could complement this research, and further
different terrain conditions: (a) flat, unobstructed areas, (b) open improve the understanding of the actual behaviour and prevent
terrain with scattered obstructions and (c) areas of suburban failures of these structures.
housing. Particularly, only AS/NZS- 2011 code states a specific
load pattern for cyclonic, which is different from the traditionally
considered standard continental winds in accordance with some Disclosure statement
recent measurements. Basic parameters related to load patterns No potential conflict of interest was reported by the authors.
such as solidity ratios ϕ for studied towers were also presented.
Pushover analyses were carried out to evaluate tower inelastic ORCID
response under intense load wind. Thus, 14 code load patterns
Tapia-Hernández Edgar   http://orcid.org/0000-0003-4269-180X
(12 for continental winds and 2 for cyclonic regions) were applied
in order to determine the final force and the site wind speed
VR leading to impending failure mechanism. According to the References
results, the failure mechanism was independent to the applied AIJ-06. (2006) Recommendations for loads on building. Tokyo: AIJ
load pattern. Wind loading models contain many simplifications, Architectural Institute of Japan. English translation, pp. 1–56.
but there are some general conclusions that can be drawn, and Amano, T., Fukushima, H., Ohkuma, T., Kawaguchi, A., & Goto, S. (1999).
need to be refined in future research work. Main observations The observation of typhoon winds in Okinawa by Doppler sodar.
from the results of this study are: Journal of Wind Engineering and Industrial Aerodynamics, 83, 11–20.
ASCE 74. (1991). Guidelines for electrical transmission line structural
• The tower drag force coefficients have the largest variation loading. ASCE Manuals and Reports on Engineering Practice No. 74,
New York, NY: American Society of Civil Engineers Press.
at solidity ratios ϕ close to that of typical lattice transmis- ASCE 7. (2010). Minimum design loads for buildings and other structures.
sion towers (ϕ ≈ 0.20). ASCE/SEI 7–10. New York, NY: American Society of Civil Engineers,
• AS/NZS- 2011 leads to similar wind forces between the ASCE Press.
cyclonic and non-cyclonic regions for the unobstructed AS/NZS. (2011). AS/NZS 1170.2. Structural design actions. Part 2: Wind
terrain condition. However, the difference is noticeable actions. Australian/New Zealand Standard.
AS/NZS Supp (2011). AS/NZS 1170.2 Supplement 1: 2002. Structural
for other terrains, where cyclonic wind forces are greater design actions, wind actions – Commentary. Australian/New Zealand
than the continental wind ones. This observation reflects Standard. March.
the implicit assumption that cyclonic wind criterion con- Banik, S., Hong, H., & Kopp, G.A. (2008). Assessment of structural capacity
siders a lower surface roughness (lower boundary layer). of an overhead power transmission line tower under wind loading.
STRUCTURE AND INFRASTRUCTURE ENGINEERING   13

Proceedings of BBAA VI (pp. 20–24). Milano: International Colloquium Liang, S., Zou, L., Wang, D., & Cao, H. (2015). Investigation on wind
on Bluff Bodies Aerodynamics & Applications, July. tunnel tests of a full aeroelastic model of electrical transmission tower-
Banik, S., Hong, H.P., & Kopp, G.A. (2010). Assessment of capacity curves line system. Engineering Structures, 85, 63–72.
for transmission line towers under wind loading. Wind and Structures Masters, F., Tieleman, H., & Balderrama, J. (2010). Surface wind
An International Journal, 13, 1. measurements in three Gulf Coast hurricanes of 2005. Journal of Wind
Bayar, D.C. (1986). Drag coefficients of latticed towers. Journal of Structural Engineering and Industrial Aerodynamics, 98(10–11), 553–547.
Engineering, 112(2), 417–430. Mara T.G. (2013). Capacity assessment of a transmission tower under
Carril, C.F., Isyumov, N., & Reyolando, M.L.R.F. (2003). Experimental study wind loading (University of Western Ontario – Electronic thesis and
of the wind forces on rectangular latticed communication towers with dissertation repository). Paper 1527.
antennas. Journal of Wind Engineering and Industrial Aerodynamics, 91, Mara, T.G., & Hong, H.P. (2013). Effect of wind direction on the response
1007–1022. and capacity surface of a transmission tower. Engineering Structures, 57,
CENAPRED-CEPAL. (2007). Características e impacto socioeconómico 493–501.
de los huracanes “Stan” y “Wilma” en la República Mexicana en el 2005 Mara, T.G. (2014). Influence of Solid area distribution on the drag of a
[Characteristics and socio-economic impact of Hurricanes “Stan” and two-dimensional lattice frame. Journal of Engineering Mechanics, 140,
“Wilma” in the Mexico in 2005]. Mexico City: Centro Nacional de 644–649.
Prevención de desastres (CENAPRED) (in Spanish). Mazzoni, S., McKenna, F., Scott, M., & Fenves, G. (2006). Open system for
CFE- 2008. (2008). Manual de Diseño de Obras Civiles. Capítulo de Diseño earthquake engineering simulation. User Command-Language Manual,
por Viento [Civil works manual. Chapter of wind design]. México City: Report NEES grid-TR 2004-21. Berkeley, CA: Pacific Earthquake
Instituto de Investigaciones Eléctricas, Mexican Federal Electricity Engineering Research, University of California. Retrieved http://
Commission (in Spanish). opensees.berkeley.edu
CIGRÉ. (2008). How overhead lines respond to localized high intensity Momomura, Y., Marukawa, H., & Ohkuma, T. (1992). Wind-induced
winds basic understanding. International Council on Large Electrical vibration of transmission line system. Journal of Wind Engineering and
Systems. Working Group B2.06.09. 2008. Technical Brochure 350, 77p. Industrial Aerodynamics, 43, 2035–2046.
ISBN: 978-2-85873-037-7. Muñoz J.C., Sánchez J., López A. & Pérez-Rocha L.E. (2009), Design wind
CIGRÉ. (2012). Overhead line design guidelines for mitigation of severe speeds for Mexico: An optimum approach of wind design of structures,
wind storm damage. International Council on Large Electrical Systems. Proceedings: 11th Americas Conference on Wind Engineering, American
Working Group B2.39. Technical Brochure 485, 38 p. ISBN: 978-2- Association for Wind Engineering, San Juan, Puerto Rico.
85873-177-0. NTCV. (2004). Normas Técnicas Complementarias para Diseño por
CSI. (2010). SAP 2000 V14 integrated finite element analysis and design of Viento [Wind design guidelines of Mexico’s City Code]. Gaceta Oficial
structures. Berkeley, CA: Computers and Structures, CSI. del Distrito Federal, décimo cuarta época, tomo II, October. México
Fleck-Fadel, L., Fleck-Fadel, L., Riera, J.D., Kaminski, J., & Ramos, R.C. City. (in Spanish).
(2012). Assessment of code recommendations through simulation of Pasch, R.J., Blake, E.S., Cobb, H.D., & Roberts, D.P. (2006, September 29).
EPS wind loads along a segment of a transmission line. Engineering Hurricane Wilma Tropical Cyclone Report. National Hurricane Center.
Structures, 43, 1–11. Pengyun L., Jiedong L., Ming N., Wanli Z., & Anguo H. (2012). Dynamic
Fu, J.Y., Li, Q.S., Wu, J.R., Xiao, Y.Q., & Song, L.L. (2008). Field response of power transmission towers under wind load. Energy
measurements of boundary layer wind characteristics and wind- Procedia, 17, 1124–1131. Retrieved from http://dx.doi.org/10.1016/j.
induced responses of super-tall buildings. Journal of Wind Engineering egypro.2012.02.217
and Industrial Aerodynamics, 96(8–9), 1332–1358. Prasad, R.N., Knight, G.M.S., Mohan, S.J., & Lakshmanan, N. (2012).
Holmes, J.D. (1995). Along-wind response of lattice towers – II. Studies on failure of transmission line towers in testing. Engineering
Aerodynamic damping and deflections, Engineering Structures, 18(7), Structures, 35, 55–70.
483–488. Savory, E., Parke, G.A.R., Zeinoddini, M., Toy, N., & Disney, P. (2001).
Holmes, J.D. (2007). Wind loading of structures. (2nd ed.). London: Spon Modelling of tornado and microburst-induced wind loading and failure
Press. of a lattice transmission tower. Engineering Structures, 23, 365–375.
Holmes, J.D. (2008). Recent developments in the specification of wind Shanmugasundaram, J., Harikrishna, P., Gomathinayagam, S., &
loads on transmission lines. Wind and Engineering an International Lakshmanan, N. (1999). Wind, terrain and structural damping
Journal, 5, 8–18. characteristics under tropical cyclone conditions. Engineering
Hsiao, P.-C., Lehman, D.E., & Roeder, C.W. (2013). A model to simulate Structures, 21, 1006–1014.
special concentrically braced frames beyond brace fracture. Earthquake Shehata, A.Y., El Damatty, A.A., & Savory, E. (2005). Finite element
Engineering & Structural Dynamics, 42, 183–200. modeling of transmission line under downburst wind loading. Finite
IEC. (2003). International Electro-technical Commission (IEC). Design Elements in Analysis and Design, 42, 71–89.
criteria of overhead transmission lines (3rd ed.). IEC Standard 60826. Solari, G., & Kareem, A. (1998). On the formulation of ASCE7-95 gust
Geneva: Author. effect factor. Journal of Wind Engineering and Industrial Aerodynamics,
IEEE. (2007). National electric safety code, C2. 7 CFR 1755.901(b), 77–78, 673–684.
Washington, DC: Institute of Electrical and Electronics Engineers Press. Sparks, P.R. (2003). Wind speeds in tropical cyclones and associated
ISO 4354. (2009, June). Wind actions on structures (2nd ed.). International insurance losses. Journal of Wind Engineering and Industrial
Organization for Standardization, 1, 68 p. Aerodynamics, 91, 1731–1751.
Jiang, W.Q., Wang, Z.Q., McClure, G., Wang, G.L., & Geng, J.D. (2011). Tapia F., & Valdepeña M. (2002, October), Summary of transmission lines.
Accurate modeling of joint effects in lattice transmission towers. 230 kV and 400 kV towers. Light and Power Company, Civil Engineering
Engineering Structures, 33, 1817–1827. Department (1st ed.), pp. 49– 52 (in Spanish).
Kaminski, J., Riera, J.D., de Menezes, R.C.R., & Miguel, L.F.F. (2008). Model Tapia-Hernández, E., & Tena-Colunga, A. (2014). Code-oriented
uncertainty in the assessment of transmission line towers subjected to methodology for the seismic design of regular steel moment-resisting
cable rupture. Engineering Structures, 30, 2935–2944. braced frames. Earthquake Spectra, 30, 1683–1709.
Kasperski, M. (2009). Specification of the design wind load – A critical Tapia-Hernández, E., Ibarra-González, S., & De-León, D. (in press).
review of code concepts. Journal of Wind Engineering and Industrial Collapse mechanisms of power towers under wind loading. Structure and
Aerodynamics, 97, 335–357. Infrastructure Engineering. doi:10.1080/15732479.2016.1190765
Lin W.E., Savory E., McIntyre R.P., Vandelaar C.S., & King, J.P.C.. (2011, July). A Tapia-Hernández E., & Cervantes-Castillo J.A. (2016). Influence of the
single-span aeroelastic model of an overhead electrical power transmission drag coefficient on communication towers. International Journal of Civil
line with guyed lattice towers. In Proceedings: 13th International Conference Engineering. doi: 10.1007/s40999-017-0157-z
on Wind Engineering. Amsterdam, The Netherlands.
14   E. TAPIA-HERNÁNDEZ AND E. SORDO

Uriz P., & Mahin S. (2008, November). Toward earthquake-resistant Yin, T., Lam, H.F., Chow, H.M., & Zhu, H.P. (2009). Dynamic reduction-
design of concentrically braced steel-frames structures. Report of Pacific based structural damage detection of transmission tower utilizing
Earthquake Engineering Research Center. PEER 2008/08. ambient vibration data. Engineering Structures, 31, 2009–2019.
Yang, F., Yang, J., Niu, H., & Zhang, H. (2015). Design wind loads for Webster, P.J., Holland, G.J., Curry, J.A., & Chang, H.R. (2005). Changes
tubular-angle steel cross-arms of transmission towers under skewed in tropical cyclone number, duration, and intensity in a warming
wind loading. Journal of Wind Engineering and Industrial Aerodynamics, environment. Science, 309, 1844–1846.
140, 10–18.

View publication stats

You might also like