You are on page 1of 19

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

net/publication/305376890

Collapse mechanisms of power towers under wind loading

Article  in  Structure and Infrastructure Engineering · July 2016


DOI: 10.1080/15732479.2016.1190765

CITATIONS READS

17 852

3 authors:

Edgar Tapia-Hernández Santiago de Jesus Ibarra-González


Metropolitan Autonomous University 5 PUBLICATIONS   22 CITATIONS   
81 PUBLICATIONS   309 CITATIONS   
SEE PROFILE
SEE PROFILE

David De Leon
Universidad Autónoma del Estado de México (UAEM)
84 PUBLICATIONS   223 CITATIONS   

SEE PROFILE

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

Estudio paramétrico del modelado inelástico de contravientos de acero View project

Losses on steel buildings under Hurricane winds in Mexico 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

Collapse mechanisms of power towers under wind


loading

Edgar Tapia-Hernández, Santiago Ibarra-González & David De-León-


Escobedo

To cite this article: Edgar Tapia-Hernández, Santiago Ibarra-González & David De-León-
Escobedo (2016): Collapse mechanisms of power towers under wind loading, Structure and
Infrastructure Engineering, DOI: 10.1080/15732479.2016.1190765

To link to this article: http://dx.doi.org/10.1080/15732479.2016.1190765

Published online: 15 Jul 2016.

Submit your article to this journal

Article views: 60

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: [Universidad Autonoma Metropolitana] Date: 22 July 2016, At: 20:44
Structure and Infrastructure Engineering, 2016
http://dx.doi.org/10.1080/15732479.2016.1190765

Collapse mechanisms of power towers under wind loading


Edgar Tapia-Hernándeza  , Santiago Ibarra-Gonzáleza and David De-León-Escobedob
a
Departamento de Materiales Universidad Autónoma Metropolitana-Azc, Mexico City, Mexico; bCivil Engineering Department, School of Engineering,
Universidad Autónoma del Estado de México, Estado de México, Mexico

ABSTRACT ARTICLE HISTORY


In this paper, nonlinear static pushover analyses were carried out to assess the capacity and collapse Received 1 August 2015
mechanism of two existing transmission towers under wind loading. Different load distribution patterns Revised 24 March 2016
were adopted from the application of the codified design wind load of the following countries: Mexico, Accepted 11 April 2016
United States, India, Japan and the New Zealand – Australian Code. Three-dimensional inelastic response
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

KEYWORDS
analyses were performed using open access software. Analyses were employed to define the capacity Transmission tower; collapse
curves, stress-strain curves for structural elements, the yielding mapping sequences and the collapse mechanisms; pushover
mechanisms, and to evaluate the influence of the tower body deformation and to assess the theoretical analyses; wind load; ductility;
overstrength and ductility capacities. Results show a damage concentration, which leads to a fragile overstrength
collapse mechanism with important strength reserves and a non-uniform distribution of yielding within
the tower height. Since the collapse mechanism is not compatible with the desired performance inherent
to the design philosophy, recommendations for the design stage are proposed, which pretend to ensure
that the inelastic behaviour be consistent with the goals implicit in a code-based design to prevent tower
collapses.

1. Introduction during full-scale testing of transmission line towers, as presented


by Prasad et al. (2012).
Recently, the number of partial or total collapse of transmission
The desirable collapse mechanism of a transmission tower
lines towers, under wind loading, has been increasing in Mexico
must provide enough energy dissipating capability as well as
(Figure 1). In spite of the fact that these damages are supposed adequate strength and stiffness. It is desirable to take advantage
to be controllable, the vulnerability of tower transmission lines of the inelastic energy dissipation capacity by avoiding fragile
under severe wind hazards is one of the major concerns for failure modes and, with the development of ductility, by the
designers, owners, managers and society. However, there are still excursion through the inelastic force–displacement curve. Even
several uncertainties related to the exposed structures, storms though a prescribed mechanism cannot be totally achieved, the
and the particular conditions of each tower, etc.; that makes it ductile behaviour can be promoted, and brittle mechanisms can
difficult to quantify the probability of failure or the numbers of be avoided by prescribing balanced safety factors and ductility
failures of a line over a period of time. provisions. It is known that the leg element buckling is less desir-
Even though the design of transmission line towers follows able than brace buckling failure, as the former will lead to a total
the code provisions, their collapse mechanism depends on many collapse earlier than the latter one.
factors such as improper detailing, material defects, fabrica- Nevertheless, as a common practice, the required codified
tion errors, forces during erection, variation in bolt strength, design check for strength is carried out at structural element level
detailing, connection failures, etc. (Prasad, Knight, Mohan, & rather than at the structural system level. While it is possible and
Lakshmanan, 2012). It is known, however, that the structural desirable to design so that damage occurs in the bracing system
behaviour of a power transmission tower may depend not only rather than in the leg elements, it is not possible to completely
on structural elements, but also on the wind loading source avoid the damage in vertical elements. In general, the number
characteristics as Savory, Parke, Zeinoddini, Toy, and Disney of possible failure modes increase with the number of elements.
(2001), De Oliveira et al. (2006) and Banik, Hong, and Kopp Detailing for damage control may not be sufficient based on lin-
(2008) pointed out. Mara and Hong (2013) investigated the ine- ear frame analysis of transmission towers, because damage does
lastic response of a lattice tower under different wind events; the not necessarily occur only at the locations of maximum demands
analytical results from nonlinear static pushover show that the as indicated in a linear analysis. So, this panorama shows the
yield stress and capacities depend on the wind direction. And importance to keep developing studies to enhance the failure
completely different types of premature failures were observed prediction, given the high direct or indirect losses because of

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


© 2016 Informa UK Limited, trading as Taylor & Francis Group
2    E. Tapia-Hernández et al.

collapses, in order to improve the knowledge acquired and pro- Until 2006, in Mexico there were 18,144  km of 400-kV
pose better design recommendations. two-circuit power transmission lines, which represents about
The suggested design wind load is calibrated based on the 40% of the total extension of the high transmission available lines
assumption that the tower behaves elastically during a wind in the country (27,148 km of 230-kV and 475 km of 165 kV)
event. However, this design approach does not provide any (Bolivar-Villagómez, 2006). At the metropolitan area of Mexico
insight into the inelastic tower behaviour under extreme wind City (one of the largest cities in the world by population and
loading, even though the consideration of the inelastic response density), the importance of such structures is based on the fact
can be important (Lee & McClure, 2007). Hence, this paper deals that the power is provided for 24 million of people through an
with the results of nonlinear analysis for two tension-type towers external circuit of 400-kV power transmission lines.
designed to fulfil all the recommendations of the wind and steel Additionally, the Mexican power grid exports about 1.3
guidelines of the Mexico’s Federal District Code (MFDC-04, Terawatt per hour of electricity to the United States per year;
2004). Specifically, the main thrust of this paper is focused on whereas, the 400-kV interconnection line to Guatemala and
the performance of existing towers, in order to advance on the Belize has a transmission capacity of 200 Megawatts per hour
knowledge about failure mechanisms and design recommenda- (IEA, 2012). So, studied towers are essential on the local elec-
tions to prevent collapses on these towers. tric power system. The configuration of the studied towers is
Nonlinear static pushover analyses were carried out for frequently used in Mexico, both for tangent towers with suspen-
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

two transmission line towers under wind loading in order to sion string (2 degrees of line diversion) and also for towers with
study the collapse mechanism. For the analyses, different load tension string from 10 to 60 degrees.
distribution patterns were adopted, through the application Herein, transmission line towers were analysed by using
of the specified design wind load in pushover analyses, for commercial software (CSI, 2009) (Figure 3(a)) through a lin-
several wind codes. In particular, six international regulations ear static analysis in 3D model under: (a) self-weight of cables,
were considered: North-America ASCE SEI 7-10 (ASCE-10 insulators and tower; (b) mechanical tension in cables due to the
2010); Japan AIJ-06 (AIJ-06, 2006, chapter 6); Australian/New line deviation and (c) wind loads acting on the cables and the
Zealand 1170.2 (AS/NZS-11, 2011); India IS 875 (IS875-95 tower body. For the analytical model, only braces were modelled
1992) and two Mexican Codes: (a) the Manual of the Mexican as a pin jointed space truss subject to axial forces only (Figure
Federal Electricity Commission (MOC-CFE-08) and (b) the 3(c)), where hinges were placed at both ends of the member. Leg
Guidelines for Wind Design of the Mexico’s Federal District elements were considered as continuous element in agreement
Code (NTCV-04, 2004). Results of the analysis were performed with the actual structure’s condition (Figures 3(d) and (e)). Given
to: (a) identify load–strain curves for the structural elements, that the leg elements members were modelled using frame ele-
(b) draw global load–deformation curves, (c) determine ments, because of the potential formation of plastic hinges and,
yielding mapping sequences and collapse mechanisms by using a nonlinear force–deformation relation for the mem-
and (d) assess theoretical overstrength factors and ductility bers, which results from the interaction between the axial and
capacities. flexural stresses. In contrast, for the individual brace members,
it was assumed a buckling failure mode. Splice connections were
modelled and they were also assumed to exhibit full flexural
2.  Models description
strength and stiffness continuity.
The 400-kV tangent type steel lattice towers (Figure 2) were mod- Concentric loads at both ends of members were assumed to
elled as three dimensional structures according to the design and take into account the geometric symmetry of the actual struc-
construction drawings. Studied structures (Table 1) are repre- ture (Figure 3(f)). Buckling strength of the member about its
sentative of typical line transmission towers in Mexico; especially minor axis was considered for elements made of single-angle
for the metropolitan area of Mexico City. section whereas, for elements made out of double angle sections,

(a) Collapse in Tulancingo, Mex. Nov, 2008 (b) Collapse in Huehuetoca, Mex. April, 2014.

Figure 1. Damage in transmission towers caused by wind loading in Mexico in recent years.
Structure and Infrastructure Engineering   3

52.53m Legs
5.45 1.5 Braces

8.0 m
L3"x1/4"
44.53m 5.45 L2"x1/4"
42.68m Legs
5.45 1.5 Braces
L4"x1/4"

8.5 m
8.0 m
L3"x1/4" L3"x1/4"
L2"x1/4" 36.03m 5.45
34.68m 5.45
L4"x3/8"
L4"x1/4"

8.5 m
8.5 m

L3"x1/4" 27.53m 5.45 2L4x1/4"


26.18m 5.45

52.53 m
L4"x3/8"
42.68 m
8.5 m

2L4"x3/8"
17.68m 5.45 2L4x1/4"
L4"x1/4"
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

27.53 m
2L4"x3/8"
17.68 m

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

2L4"x1/2"
0.00m L3"x1/4" 0.00m L3"x1/4"

5.12 m 5.92 m

(a) Model 43T10 (b) Model 53T10

Figure 2. Characteristics of studied towers.

Table 1. Structural specifications.


Model 43T10 53T10
Structure height 42.68 m 52.53 m
Square base area 5.12 m × 5.12 m 5.92 m × 5.92 m
Number of elements 879 980
Voltage 400 kV
Span (both sides) 400 m
Line angle 10 degrees
Insulator Tension type (3.8 kN each)
Conductor ACSR 1113 Bluejay. Mechanical tension: 36.3 kN
Ground wire Galvanised wire 3/8”. Mechanical tension: 9.3 kN
Steel sections Equal-legged angles. ASTM A-572
Bolts ASTM A394–00 with standard hole

it was assumed that the buckling strength is around a principal behaviour and hysteretic energy dissipation may reduce the lat-
axis. According to the results, the structures fulfil all the require- eral force demand to the structure (Raychowdhury, 2011); and
ments of the wind and steel guidelines of MFDC-04 (2004). The the foundation flexibility might modify the input lateral demand.
considered load combinations and the load factors are shown Thus, it is expected that the obtained results are modified when
in Table 2; whereas the demand/capacity ratios for the main leg the soil foundation dynamic stiffness is introduced in the model.
elements are presented in Figure 4. Dynamic properties are sum- According to the results, the maximum demand is linked to
marised in Table 3 and reported with further details in Ibarra- the wind patterns perpendicular to the line direction (Table 2),
González (2014). which is in agreement with the findings presented in other exper-
The influence of the motion of the transmission lines in the imental and analytical researches (Liang, Zou, Wang, & Cao,
neighbouring spans and the interaction effect between the tower 2015). For this load combination, the main source of demand
and its foundation were not considered in the analysis. In par- is the wind pressure in the tower body, which represents 50%,
ticular, soil – foundation – structure interaction might affect the independently of the tower’s height. To illustrate this, the contri-
response of steel structures in several ways: foundation move- bution rates of the main participating components are calculated
ment can modify the period of a system increasing its flexibility for the critical supports as a function of the axial load demand
(Tapia-Hernández, De Jesús, & Fernández-Sola, 2016); nonlinear as shown in Table 4.
4    E. Tapia-Hernández et al.

(c) Brace connection detail (e) Splice connection


Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

(a) 3D (b) Existing 53T10 (d) Splice connection (f) Base view
model

Figure 3. Studied towers details.

Table 2. Considered load combinations.

Mechanical tension
Self-weight of tower Self-weight of cables in cables due the line Wind loads acting on Wind loads acting on
Load Combination elements and insulators deviation tower body cables
1. Gravitational 1.15 1.50 1.15 --- ---
2. Longitudinal direction 1.10 1.10 1.10 1.10 ---
3. Perpendicular direction 1.10 1.10 1.10 1.10 1.10

Load combination 1 50
Load combination 2
Load combination 3
40
40

30 30
Height (m)
Height (m)

20 20

10 10

0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Demand/Capacity ratio Demand/Capacity ratio
(a) Model 43T10 (b) Model 53T10

Figure 4. Demand/capacity ratio for the main leg elements.

Table 3. Natural dynamic properties of towers.


Model Mode Main direction Period (s) Frequency (cycle/s) Angular Freq. (rad/s) Eigenvalue (rad2/s2)
43T10 1 Translation in line direction 0.493 2.03 12.75 162.6
2 Translational perpendicular to line direction 0.488 2.05 12.87 165.7
3 Rotational 0.177 5.65 35.52 1261.7
53T10 1 Translation in line direction 0.605 1.65 10.38 107.7
2 Translational perpendicular to line direction 0.601 1.66 10.45 109.3
3 Rotational 0.197 5.08 31.89 1016.8
Structure and Infrastructure Engineering   5

Table 4. Contribution rate by demand at the critical support.


Model 43T10 53T10
Component Load (kN) Contribution (%) Load (kN) Contribution (%)
Cable self-weight 11.8 1.8 11.8 1.4
Tower self-weight 34.0 5.1 43.2 5.0
Mechanical tension in cables 118.6 17.9 130.3 15.0
Wind pressure in cables 167.2 25.2 260.8 30.0
Wind pressure in tower body 332.9 50.0 422.4 48.6
Σ= 664.5 100.0 868.5 100.0

Solidity ratio 50
Proposal
40 40

Height, z (m)
Solidity ratio
30 30
Height, z (m)

Proposal

20 20
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

10 10

0 0
0.10 0.15 0.20 0.25 0.30 0.10 0.15 0.20 0.25 0.30
Solidity ratio, φ Solidity ratio, φ
a) Model 43T10 b) Model 53T10

Figure 5. Solidity ratio ϕ.

4 div. 1.5

P/Py
t=10 mm 1
5 div.

plate (typ.)
0.5
4 div. 5 div.
4 div

0
0Strain (%)
4 div

-0.6 -0.4 -0.2 0.2


5 div.

-0.5
5 div.

P/Rc
-1
4 div

5 div. 4 div. -1.5


Single angle section Double angle section (b) Typical hysteretic response of the
(a) Section discretisation modelled elements

Figure 6. Section characteristics for the plastic analysis.

2.1.  Solidity ratio ϕ 𝜙(z) = z2


+ 0.17 0.0 ≤ z ≤ 0.55H ⎫
3H 2 ⎪
The solid areas of leg and diagonal members were carefully ⎬ (1)
computed from the construction drawings, so that the solidity 𝜙(z) = H−z
5H
+ 0.17 0.55H < z ≤ H ⎪ ⎭
ratios ϕ (effective solid area to total area of the surface) were
calculated along the height (Figure 5). From these results, lateral 3.  Nonlinear models
nodal forces were obtained.
Three-dimensional inelastic analyses were performed by using
Since solidity ratio values ϕ vary widely along the tower height
the open access software (Mazzoni, McKenna, Scott, & Fenves,
(Figure 5), they are difficult to determine in the preliminary
2006). Nonlinear beam-column elements, with plasticity spread
design of new structures. For this reason, a proposal to roughly along the element length, were considered. Two rectangular
assess an initial estimation of the solidity ratio ϕ(z) at height (z) patches were used to generate the cross section of angles: one for
for such structures was developed (Equation (1)) and it is also each leg. Patches were discretised into fibres with quadrilateral
summarised in Figure 5. The proposed expression is expressed as shapes and four integration points per element (Figure 6(a)).
a function of the tower height H and it crudely serves to estimate Figure 6(b) illustrates the hysteretic curve of a typical single angle
a preliminary magnitude of the wind forces in new designs. The element, where P is the axial load obtained along the analysis, Py
proposed curve was developed from the identified trend in two is the yielding capacity in tension (=AFy) and Rc is the buckling
ranges of the obtained solidity ratios ϕ(z): capacity in compression.
6    E. Tapia-Hernández et al.

b= 500 mm 0 (typ.)
t=10 mm
plate (typ.)

0 (typ.)
Bolt A394 Pinned

h/6 h/6 h/6 h/6 h/6 h/6


Joint detail joint
Leg
member Bracing
h member Node

Joint (see detail)


.)
typ
19 mm Guset 0(
plate (typ.) 0(
typ
.)
Variable Variable

(a) Tower elevation (b) Structure model

Figure 7. Typical elevation.


Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

In the model, torsional restraint of the element was also algorithm was selected, in the software library (Mazzoni et al.,
included when out-of-plane buckling is studied. Element 2006), to achieve rapid convergence. The solution algorithm uses
torsional properties have been added to the fibre nonlinear an energy increment test, which checks the positive force conver-
beam-column element by using the corresponding aggregation gence if half of the inner-product of the displacement increment
tool in the software (Mazzoni et al., 2006). In addition, leg mem- and the unbalanced force is smaller than a tolerance equal to
bers were modelled with a set of six nonlinear beam-column 10−5. The magnitude of the convergence tolerance depends on
elements to reproduce the response of an axially loaded element the computational capacity, units, time step, etc. Here, this mag-
including large translational displacement and P-delta effects as nitude is used by taking into account that the models are built
shown in Figure 7. with approximately 1000 elements (Table 1). The equations are
Based on the drawing specifications, a steel with nominal formed using a SparseGeneral scheme and they are numbered
properties fy = 345 MPa, E = 200,000 MPa, and G = 77 GPa was using the reverse Cuthill-McKee numbering, which optimises the
used for all members; except for the cross section L 2′′x1/4′′, node numbering in order to reduce the storage bandwidth. This
where Fy  =  248  MPa was considered. The uniaxial Giuffre- method displays an error message when the structure is discon-
Menegotto-Pinto material is used for steel fibres with extensions nected (collapsed). The constraints were represented with a Plain
included for kinematic and isotropic hardening (Filippou, Popov, constraint handler, which enforces homogeneous single-point
& Bertero, 1983). The inelastic model used herein accounts for constraints (homogenous boundary conditions).
the response along the element by integration of the uniaxial A single load step is performed by using a load control integra-
hysteric steel material model over the cross section. This for- tor for the vertical loads: self-weight of the structure, self-weight
mulation offers numerical efficiency, and it has a reasonable of cables and insulators and components of the mechanical ten-
agreement with experimental results and provides a more real- sion in cables due the line deviation. Here, the origin in the time
istic representation of the elements response as compared to the domain is set to 0.0 and this load is kept constant. Subsequently,
simpler bi-linear model (Uriz & Mahin, 2008). load wind patterns on the tower body, on the insulators and on
Thus, element model was found to give realistic predictions the cables, were added as static incremental loadings. Elements
of the inelastic response of a member. Reasonable agreement can (legs and brace) in tower model were subjected only to end
be observed between the model in the software (Mazzoni et al., forces, with no concentrated loads along the member length.
2006), and an experimental test used to predict the capacity of a Pushover analyses are useful to assess the structural capacity
member under axial load in compression. It has been observed or to characterise the capacity curve by monotonically increasing
that axially loaded steel elements give realistic predictions of the the lateral loads. The application of inelastic analyses (static and
inelastic response of a member (Hsiao, Lehman, & Roeder, 2013). dynamic) in wind engineering is not common because the design
Reasonable agreement can be observed between the model and an for wind actions usually requires the structures to remain into
experimental test to predict the capacity of a member under axial the elastic range. Therefore, this is used to evaluate the structural
compression loads (Uriz & Mahin, 2008). This is by far superior to capacity under seismic excitations due to the fact that, in the
the use of empirical or semi-empirical models, for which specific seismic design philosophy, the structures are allowed to undergo
test data are needed to adequately reproduce key response prop- the inelastic range under severe excitations (Banik et al., 2008).
erties. Further details and validation of this modelling technique In order to do that, pushover analyses were carried out in order
may be found elsewhere in Ibarra-González (2014). to assess the nonlinear response of the structures under wind
loading through the identification of the collapse mechanism
and the inelastic capacities.
4.  Pushover analyses
The tower model with inelastic material properties is examined
Because of the material nonlinearities, an iterative solution under a nonlinear static analysis and its internal deformations
is required and a Newton method with line search solution and forces are calculated. The target displacement is intended to
Structure and Infrastructure Engineering   7

100 100 100


90 90 90

Towers height (m)


Towers height (m)

Towers height (m)


80 80 80
70 70 70
60 60 60
50 50 50
40 40 40
30 30 30
20 20 20
10 10 10
1.0 1.2 1.4 1.6 1.8 2.0 1.0 1.2 1.4 1.6 1.8 2.0 1.0 1.2 1.4 1.6 1.8 2.0

IS 875-95 VD/V R IS 875-95 VD/V R IS 875-95 VD/V R


NTCV-04 NTCV-04 NTCV-04
MOC CFE-08 MOC CFE-08 MOC CFE-08
AIJ-06 AIJ-06 AIJ-06
AS NSZ-11 AS NSZ-11 AS NSZ-11
ASCE-05 ASCE-05 ASCE-05
Note: Standards defined in References
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

(a) Before the hill (b) On the crest (c) After the hill

Figure 8. Wind velocity ratios for different positions around a hill.

(a) (b) (c)

θ
θ

Figure 9. Structure location in a hill-shape condition.

represent the maximum displacement likely to be experienced In addition, the complexity of the formulation makes neces-
during the design demand. Because the mathematical model sary the use of advanced knowledge to understand the concepts
accounts directly for the effect of material inelastic response, the and limitations behind the theoretical wind load patterns. For
calculated internal forces will be reasonable approximations of example, estimated gusts associated with the theoretical wind
those expected during the design demand. Recent studies (Banik, load patterns are estimated from the statistical outgrowth of the
Hong, & Kopp, 2010; Mara, 2013) demonstrated that the capac- turbulence and they are not a measured quantity (Peyrot, 2009).
ity curves of a transmission line tower obtained from pushover It seems that many designers of transmission towers believe that
analysis, and from the Incremental Dynamic Analyses (IDA), are a better design can be achieved through the precision of complex
similar. Thus, the dynamic wind load has little or no effect on formulas that depend, sometimes, on unknown and intricate
the capacity curve and, hence, the pushover analysis is adequate parameters, instead of promoting common sense and design
to capture the essential features of the structural response that simplicity.
significantly represents its performance. In spite of the fact that the loading patterns in codes are based
on theoretical wind patterns developed more than 40–50 years
ago, they have evolved independently due to local adaptations
4.1.  Loading protocols
and, therefore, extremely different load patterns can be obtained
The load patterns of wind provisions of the majority of the codes even for equal conditions depending on the specification. To
or guidelines are based on a very sophisticated mathematical pro- illustrate this, in Figure 8 the velocity ratios of theoretical wind
posal developed in the 1960’s (Davenport, 1961). Modern codes profiles are shown for structures at a height z. Load velocity
propose a theoretical wind load pattern based on local observa- patterns VD are calculated for an inclination of θ  =  7.5° and
tions and, given the need for simplifications they assume that, H = 26.3 m in three different structure locations around the hill
below the gradient height, the average wind velocity decreases by (Figure 9): (a) 200 m before the hill, (b) at the crest (0 m) and
following a predefined profile as a function of the terrain condi- (c) 200 m after the hill.
tions (somehow arbitrarily defined for 3–5 categories depending Wind forces were calculated at the top of structures between
on the code). The selection of terrain category is made consid- 10 m and 100 m in height in order to illustrate the velocity var-
ering the presence of obstructions that constitute the surface iation for each code. Mean basic wind speed at 10  m height
roughness. Nevertheless, in actual situations, terrain categories was assumed to be equal to VR = 39 m/s (≈140 km/hr), for an
are difficult to set because of the uncertainty in the assumed open terrain condition and equal circumstances of temperature,
shapes of the load pattern in height. atmospheric pressure, air density, importance factor, return
8    E. Tapia-Hernández et al.

50
MOC CFE-08
ASCE-05
40 MOC-CFE 40 AIJ-06
ASCE-05 AS/NSZ-11
AIJ-06 IS 875-95
30

Height (m)
AS/NSZ-11 30
IS 875-95
Height (m) 20 20

10 10

0 0
0.7 0.8 0.9 1.0 1.1 0.7 0.8 0.9 1.0 1.1
Force/Force MOC CFE-08 Force/ForceMOC CFE-08
(a) Load patterns for 43T10 model (b) Load patterns for 53T10 model
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

Figure 10. Normative load distribution patterns considered.

period factor, wind direction factor, shielding coefficient, etc. Based on the above, five load distribution patterns obtained
In this way, the theoretical wind load patterns depend only on the from the application of the codified design wind load on tow-
terrain roughness, height factor Fα and the calculated topography ers were considered in this research (NTCV-04 was excluded,
factor FT. The terrain roughness and height factor Fα are the main because of a physical inconsistency, where gradient wind veloc-
parameters that link the wind speed with height in the atmos- ities depend on terrain categories). Patterns were obtained con-
pheric boundary layer for different terrain roughness; whereas sidering open terrain, with scattered obstructions, and they are
the topography factor FT reflects the change of the mean wind shown as normalised with respect to the MOC-CFE-08 load
speed that occurs as wind passes at right angles over escarpments pattern, for the force at the top node in order to emphasise their
or ridge-shaped topography (Figure 9). differences (Figure 10).
The wind speed profiles considered represent a traditional The incipient collapse was defined by the point where the
mean atmospheric boundary layer wind profile, where it is axial shortening exceeds the theoretical ultimate deformation
assumed that the wind is a stationary and ergodic stochastic capacity of the elements due to buckling, based on the results
process (De Oliveira et al. 2006). In MOC-CFE-08 (2008), it is of an experimental research (Kemp, 1996), although the anal-
specified that wind speed is spatially fully correlated and that yses showed a numerical convergence for a larger lateral load.
the adopted power law leads to a 3-s gust wind speed V3 s at a Therefore, incipient collapse was not defined in terms of the point
height of 10 m. Then, the wind loads were calculated in height where there is no-convergence (in the iterative algebraic process)
according to the methodology specified by the compared codes. for an increased lateral load but instead of that, by the theoretical
Nevertheless, the wind load in some cases was based on a 10-min limits for the actual element capacities. The described procedure
mean wind speed V10 min at 10 m. Therefore, a conversion of the was applied to the compression leg members, since they led to
analytical results for a direct comparison was applied through the collapse mechanism, as it is discussed later.
the ratio V3 s/V10 min = 1.43 for an open terrain condition (Durst, According to the procedure suggested by Kemp (1996), the
1960; Mara, 2013). When necessary, and depending on the code effective lateral slenderness ratio is computed by taking into
procedure, the following values were assumed: mean annual account: (a) the slenderness ratio in lateral-torsional buckling,
temperature 16.5 °C, air density ρ = 1.22 kg/m3, return period (b) a flange slenderness factor in local buckling and, (c) a dis-
of 50 years and the importance factor, the wind directionality tortional restraint factor. The procedure proposed by Kemp
factor, the shielding factor and the spatial correlation effect were is based on test results for conditions at which the strain, on
set to 1.0. the tension side of the lateral buckle, first falls under yielding
Regulatory theoretical wind profiles are shown in Figure 8, in compression, allowing for amplification of the maximum
where it might be observed that they are completely different lateral deflection due to interaction between local and lateral
although they were calculated for similar conditions. These wind buckling. The available inelastic capacity is computed by sim-
profiles have a high dependency on the topography factor, and plifying the theoretical predictions based on the stress–strain
this factor is defined for fluctuating wind speed, by parameters curve. Further details can be found elsewhere (Ibarra-González,
based on the results of wind tunnel experiments conducted 2014).
in each country. This suggests that, even though the structure
could have exactly the same basic wind demands, different design
5.  Nonlinear static analyses
sections could be obtained according to its location around the
world. Despite this, some studies (Savory, Parke, Disney, & Toy, Pushover analyses were carried out under normative load pat-
2008) indicate an excellent agreement between the code calcu- terns. Analyses were conducted for a wind direction perpen-
lations and measures from the field monitoring in transmission dicular (transverse) to the transmission line. In the structure,
towers, within the overall accuracy of the file data. the leg elements in the left-hand side were in tension with axial
Structure and Infrastructure Engineering   9

1.0 L+50m 1.0 L+50m

P/Py
0.5

P/Rc
0.5
0.0 0.0
0 1 2 0 1
ε (103) ε (103) 2
1.0 L+45m 1.0 L+45m
P/Py

P/Rc
0.5 0.5

0.0 0.0
0 1 2 0 1 ε (103) 2
ε (103)
1.0 L+35m 1.0 L+35m
P/Py

P/Rc
0.5 0.5

0.0 0.0
0 1 ε (103) 2 0 1 ε (103) 2

1.0 L+25m 1.0 L+25m


P/Py
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

P/Rc
0.5 0.5

0.0 0.0
0 1 2 0 1 ε (103) 2
ε (103)

1.0 L+15m 1.0 L+15m


P/Py

P/Rc
0.5 0.5

0.0 0.0
0 1 ε (103) 2 0 1 ε (103) 2

1.0 L+5m 1.0 L+5m


P/Py

P/Rc
0.5 0.5

0.0 0.0
0 1 ε ( 10 3) 2 0 1
ε (103) 2

Figure 11. Capacity curves for elements along the height.

extensions, while the leg elements in the right-hand side were wind patterns considered in this research. Thus, collapse mech-
in compression with axial shortenings. anism does not match the assumptions inherent on the code
No dependency on the lateral load pattern has been noticed design philosophy, where it is expected that a lot of elements
in the developed final mechanism. This observation is in agree- shall undergo an inelastic excursion.
ment with similar studies, where, for any given wind direction,
the capacity curves are approximately the same regardless of the
5.1.  Stiffness of the gusset plate joints
applied loading profile (Banik et al., 2008). Figure 11 summarises
the results obtained from the pushover analyses for the 53T10 Results of experimental studies, on elastic buckling of braces steel
model, where P is the axial load obtained along the analysis. elements axially loaded in compression, demonstrated the accu-
These graphs show capacity curves for elements at different racy of the theoretical effective lengths whose values are, by far,
heights under axial load normalised with respect to the inelastic smaller than those used for the calculation in the current design
capacity (buckling in compression Rc and yielding in tension Py) practice in members. The above is due to the restraint provided
against the strain deformation rate (ε = δ/L). For the single-angle by the complementary brace, whose stiffness is increased by the
sections, the compression capacity Rc was calculated by following presence of a tensile action (Metelli, 2013). Other studies (Da
the criteria of the flexural-torsional buckling (singly symmetric Silva, Vellasco, De Andrade, & De Oliveira, 2005; Jankowska-
section); whereas, for the double-angle sections, the strength Rc Sandberg & Kolodziej, 2013) suggest that the possible explana-
was revised in terms of the wall thickness under (a) the torsional tion, for the braces stability (reduction of the buckling length),
buckling condition and/or (b) the flexural buckling condition is related to the fact that the actual behaviour of the element
(doubly symmetrical cross sections). connection is closer to semi-rigid connections, instead of the
Results show a local damage concentration with non-uniform assumed hinge connection. These sources of additional capaci-
distribution of plastic yielding through the height. The energy ties, not included in the specialised codes, and commonly used
dissipation is mostly produced by buckling in compression or for structural analysis for the design of transmission towers, may
yielding in tension of the leg elements (at height of 25 m approx- explain why the bracing system usually does not contribute to
imately), without subsequent damages in brace elements. In fact, the energy dissipation in transmission towers under lateral wind
the bracing system remained elastic along analyses under all patterns.
10    E. Tapia-Hernández et al.
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

Figure 12. Influence of the stiffness of gusset plate joint Kr in bracing system.

The influence of the stiffness of the gusset plate joints Kr in 5.2.  Damage concentration
the axial capacity of the bracing system were studied as a func-
It was observed that a small inelastic excursion always started
tion of the maximum deformation v of the unbraced length
with an incipient buckling of leg elements in compression located
(Figure 12(a)). Here, the model with Kr → ∞ represents an ele- at the middle-height of the tower. Peak plastic deformations were
ment with a fixed support condition (theoretical rigid connec- reported in leg elements where the cross-section changes from
tion), whereas the model with Kr = 0 represents an element with 2L4′′ × 3/8′′ to 2L4′′ × 1/4′′ at a height of approximately 25 m
a hinge connection; so that a stiffness in the range 0 < Kr < ∞ (Figure 13). The overlapping on leg elements was evaluated and
represents models with semi-rigid connections. included in studied models through a splice connection with
Incremental static inelastic analyses were developed for critical bolts (welding process is forbidden for transmission towers in
slenderness ratio of the bracing system (for cross section L4′′ × Mexico) and two additional angles L4′′ × 3/8′′, as it is shown in
1/4′′, kL/r = 65 and kL/r = 77) in the studied models (Figure 2). the view A-A in Figure 13(a).
In the three-dimensional analyses, axial load in compression was Based on the displacement of the nodes and capacity
added as static incremental load, and elements were modelled as curves of the fibres for the cross section of the member leg
above discussed by using the open access software (Mazzoni et al., elements (Figure 13(a)), it seems that the connection is far
2006). Figure 12b displays typical capacity curves, namely axial enough rigid to reduce the unbraced length, which would
load P normalised with respect to the compression buckling capac- be beneficial to the compression capacity. Nevertheless, the
ity for the hinge connection RcH, against the axial strain deforma- change of section from 4L 4′′ × 3/8′′ to 2L 4′′ × 1/4′′ leads to
tion rate (ε = δ/L). In the analyses, the maximum deformation v at a local increment of the stress, which produces the damage
the mid-span (y = L/2) was calculated in the analyses in order to concentration. This position of the splice connection is typ-
identify the stability condition of the modelled elements. ical in towers in Mexico, because, according to the assembly
The influence of the gusset plate stiffness Kr for a joint on engineers, it facilitates the erection operation. Usually, tow-
the axial capacity of the bracing system is shown in Figure ers are partially assembled by workers at ground level and
12(c) related with maximum deformation v(y  =  L/2) at the then installed through an erecting crane. Nevertheless, the
mid-span. According to the results, the assumptions adopted described failure has been previously reported on structures
in the bracing system of the studied models (Kr = 0) are related under intense winds in similar connections, for example in
to the limit of instability. So, in spite of the fact that there is Huehuetoca, Mexico in May 2014 (Figure 13(b)).
wide range of semi-rigid connections (0 < Kr < ∞) that could The reported damage concentration in the change of the cross
lead to slightly different results, this research takes into account section produced, then, a fragile collapse in both studied cases;
the physically possible most critical condition. Additionally, although the structure had a near-elastic response and important
it worth to mention that, although the brace elements were strength reserves with a non-uniform inelastic response along
modelled with pinned connections, according to the results, the height (Figure 13). Therefore, a weakness in the local design
the bracing system does not contribute to the energy (and assembly) process was identified; it is preferable to change
dissipation. the cross section at the end of the unbraced length, instead of
Structure and Infrastructure Engineering   11
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

Figure 13. Concentration of the damage.

0.00-0.25
0.25-0.50
0.50-0.75
0.75-1.00
Normalised
inelastic
deformations

(a) Initial yielding (b) δglobal= 1.53% (c) Final collapse


δglobal= 1.49% (Step 1.96) (Step 2.00) δglobal= 1.68% (Step 2.20)

Figure 14. Behaviour of the 53T10 model for the analysis under MOC-CFE-08 load pattern.

doing it at the middle length, in order to avoid an undesirable shown (with a colour scale), normalised with respect to the max-
abrupt change of stiffness along of the element. imum axial elements with deformations by step. For each case,
the per cent of global drift (deformation at the top divided by
the tower height) is also included.
5.3.  Evolution of the collapse mechanisms
It can be observed from Figure 14 that, indeed, the inelastic
Progressive collapse mechanisms for the 53T10 model are response starts with the incipient buckling of the leg elements
mapped in Figure 14. Magnitudes of inelastic deformations are in compression at the middle height. In general, for the collapse
12    E. Tapia-Hernández et al.

Figure 15. Base shear versus global drift curves.


Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

Table 5. Deformation capacities.

Elastic analyses Drift at yielding Final drift


Model Drift (%) δy (%) Step δu (%) Step
43T10 0.64 1.19 2.09 1.39 2.48
53T10 0.85 1.49 1.96 1.68 2.20

mechanism (Figure 14(c)) the energy dissipation is mostly for the service limit state, equal to δy≈1.20%. Moreover, as the
obtained by the yielding of a few elements in compression and average final drift is 1.54%, it may be suggested a code restric-
tension. Although the structure may have enough strength tion for the collapse limit state equal to 1.50%. The large diver-
capacity reserves, studied models exhibited a non-uniform dis- sity of tower configurations around the world makes it difficult
tribution of yielding along the height, which is associated to a to establish a unique deformation limit which may be applied
damage concentration mechanism. This mechanism is not con- to the particular structure configurations considered here.
sidered in the design process and, in fact, it is difficult to predict. Nevertheless, the obtained results for the considered config-
Figure 15 summarises the capacity evaluation obtained urations are useful and represent a reasonable estimation of
from the nonlinear static pushover analyses. The results are the deformation limit that a designer may take into account in
plotted as the base force in the dominant wind direction (wind order to develop reliable designs in the revision of new and/or
patterns perpendicular to the line direction) versus the global existing towers.
drift (displacement at the top of the tower divided by the struc- In order to explore the potential effect of proposing drift limits
ture height) in the same direction. The base shear is the total on the tower performance, the influence of the deformation of
reaction component from all supports in the dominant wind the tower body was evaluated through the rotation at the base as
direction. it is shown in Figure 16. The theoretical rotation ϕ(t)T was calcu-
Capacity curves are associated with a non-ductile behaviour, lated by Equation (2), which can be found by static equilibrium
where the collapse mechanism is formed as soon as the axial at the top of the tower (Figure 16(c)). Whereas, the analytical
capacity in the compression leg member is reached at the tower rotation ϕ(t)A was calculated from the displacement at the top
middle height (Figure 14). In addition, note that the global (Equation (3)) during pushover analysis:
capacity curve (Figure 15) is not enough to describe the ine- ∑ ∑ ∑
fi yi + wi xi + Ti yi − VH
lastic response of the structure, since a local concentration of 𝜙(t)T = (2)
damage (an unexpected redistribution of loads or a non-uniform PH
distribution through the tower height), is not reported at them.
Thus, only specific observations, valid for these particular cases,
v(t)
may be made duly from these results. 𝜙(t)A = (3)
H
where Σ fi is the wind loads acting on the cables and the tower
6.  Inelastic capacities
body; Σ wi is the self-weight of cables, insulators and tower body;
From pushover global curves (Figure 15), obtained drift values Σ Ti is the mechanical tension in cables due to the line deviation;
(deformation at the top divided by the tower height) were studied y is the vertical lever arm and x is the horizontal lever arm and
through three different profiles (Table 5): (a) the elastic analyses v(t) is the deformation at the top along the time of the analyses.
drift, which is considered in a typical design process, (b) the drift According to the results, along the analyses, the evolution
at first yielding δy and c) the drift at the collapse mechanism δu. of the analytical rotation ϕ(t)A obtained along the analysis is
Given the minimum obtained drift at yielding (Table 5), it more critical than the one calculated by static equilibrium ϕ(t)T.
is possible to conservatively propose a deformation restriction, Namely, the displacement at the top critically depends on the
Structure and Infrastructure Engineering   13

Theoretical Theoretical
Analytical Analytical φ
Yielding limit Yielding limit
0.06 0.06

y i+1
Context of Context of Fi+1

yi
0.05 0.05
Stability Stability Ti+1
0.04 0.04 Fi
Ti
φ 0.03 φ 0.03
0.02 0.02
wi+1
0.01 0.01
wi
0.00 0.00 xi
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0
Step(t)
V
Step(t)
(a) Model 43T10 (b) Model 53T10 (c) Variables

Figure 16. Influence of the deformation of the tower body.


Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

50
40
40

Tower height (m)


Tower height (m)

30
30
20
20
10 10
0 0
0 20 40 60 80 0 20 40 60 80
Deformation (cm)
Deformation (cm)
drift= 0.52%
drift= 0.56%
drift= 0.64% (Elastic design) drift= 0.83% (Elastic design)
drift= 0.98% drift= 1.18%
drift= 1.19% (Initial yielding) drift= 1.49% (Initial yielding)
drift= 1.39% (Collapse) drift=1.68% (Collapse)
(a) Model 43T10 (b) Model 53T10

Figure 17. Deformed configuration for different global drifts along the analyses.

(a) dg (b) Wind direction


L= 400m L= 400m T
d3
T Cables
d2

Deformed shape

10°

d1
dg

T
Undeformed
shape
Undeformed shape

T Defo
rmed ds T
dy

shap
e
s
θ

L'
T0
y

(c) W=ws
Semi-span
dx
x

Figure 18. Influence of the deformed shape in the increment of cables tension.

body deformation on the studied models. This implies that global The evolutions of these deformed configurations along the
parameters, or simplified analysis, might not correctly reflect the analyses are shown in Figure 17 for different global drifts (defor-
response of the overall structural system. mation at the top divided by tower height). According to these
14    E. Tapia-Hernández et al.

3.0 3.0

Overstrenght, Ω
43T10 43T10 2.48
2.5 53T10 2.5 53T10

Ductility, µ
2.2
2.0 2.0

1.5 1.5
1.17 1.13
1.0 1.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Tower period (sec) Tower period (sec)
(a) Ductility capacities (b) Overstrength capacities

Figure 19. Developed ductility and overstrength by studied towers.

Table 6. Evaluation of the deflection in the cables tension.


Crossarms Height (m) Deflection at collapse (cm) Elongation, L′ (cm) L′/L Catenary, y (m)
Guard wire, dg 52.5 88.7 409.7 1.024 9.022
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

Conductor 3, d3 44.5 65.2 405.3 1.013 10.686


Conductor 2, d2 36.0 43.4 402.3 1.006 10.686
Conductor 1, d1 27.5 25.7 400.8 1.002 10.686

results, flexural deformation configurations govern at the elastic According to the results, the elongation of the distance x from
stage (before the first yielding). This response is not strongly L to L′, and subsequent increment of the cable tension T has a
affected by the loading pattern or the structure height, which is in minimal influence in the final capacities of the tower because:
agreement with similar studies, where the responses are approx-
(1)  The increment of the span L′ is small L′/L → 1.0 (Table
imately the same regardless of the applied loading profile (Mara,
6), even assuming that the adjacent structures are
2013; Mara & Hong, 2013). This confirms the influence of the
infinitely rigid.
axial deformation (shortening or extension) of the leg elements
(2)  In a real scenario, the adjacent towers (Figure 18(b))
on the final collapse mechanism, and the poor participation of
would develop a deflection at the same direction,
the bracing system (shear stiffness contribution).
which would release the increment of cables tension.
(3)  The contribution rate on the critical support due the
6.1.  Influence of the deflection in the cables cables tension represents approximately 15% of the
total axial load demand (Table 4).
The deflection of the tower (Figure 18(a)) modifies the horizontal
cables length from L to L′ (Figure 18(b)), which causes a gauge
reduction of the catenary y and, therefore, an increment of the
tension T along the cable (Equation (4)). Here, c = T0/w, where 6.2.  Ductility and overstrength
w is the cable weight per unit length, T0 the cable tension and s Finally, the inelastic behaviour was also assessed through
the cable length as it is shown Figure 18(c): the ductility capacity μ  =  δu/δy and the overstrength capacity
� √ Ω  =  Vb Final/Vb Design (variables were previously defined in
T= T20 + w2 s2 = w c2 + s2 (4) Figure  15). Due to the non-ductile collapse mechanism that
was developed, both models presented a near-elastic behaviour
Then, the increment of cable tension T is a function of the hori- (μ ≈ 1, Figure 19(a)). These results are in agreement with the
zontal distance x (Figure 18 (c)), which is defined as a function obtained capacity curves, where the damage concentrations in leg
of the cable length s by Equation (5) for a cable that supports a members, by compression buckling in the middle height, lead the
uniform distributed load as self-weight: collapse, although the tower still has important strength reserves.
s � � The towers fulfil all the requirements of the wind and steel
ds s s s guidelines of the local Code (MFDC-04, 2004). However, it has
x=∫ √ = c sinh−1 = c sinh−1 (5)
0 2
1 + s ∕c2 c 0 c been shown that structures may have a considerable reserve
of strength that is not explicitly considered in current codes
In Table 6, the elongation L′ of the cable lengths for the model (Mitchel et al., 2003). This overstrength capacity in steel struc-
53T10 was calculated in terms of the maximum deflections tures arises from several factors that sometimes occur in the
at cross-arms (Figure 18(a)) and the span between towers design/construction process: (a) the restricted choices for sizes
(L = 400 m, Table 1). Here, it was supposed that adjacent struc- of members and elements (typification of sections) and rounding
tures remain without deformations in order to study the more of sizes and dimensions; (b) the difference between nominal and
unfavourable condition. The catenary y at the mid-span was also factored resistances equal to 1/γ, where γ is the material resist-
calculated. ance factor as defined in the codes; (c) the ratio of actual yield
Structure and Infrastructure Engineering   15

strength to minimum specified yield strength; (d) the develop- 7. Conclusions


ment of strain hardening; (e) the development of the collapse
This paper reports the results of nonlinear analysis for
mechanism; (f) structure redundancy, (g) the adjustments to
two tension-type steel lattice towers. Three-dimensional ine-
adapt the necessary design connections to the dimensions and
lastic analyses were performed, where nonlinear beam-column
geometry of the actual structure, among other parameters.
elements with plasticity spread along the length element, were
The assessed overstrength capacity ratios (nominal shear Vnom
used to model all elements. The cross section of angles was dis-
and maximum shear Vmax) for the analysed towers, were always
cretised into fibres with quadrilateral shapes. Leg members were
larger than two (Figure 19(b)); this could mean that the towers
modelled with a set of six nonlinear beam-column elements to
have been designed for a much larger theoretical wind profiles,
reproduce the response of an axially loaded element including
than the ones that cause damage in actual structures. In addi-
large translational displacement and P-delta effects.
tion, the results lead to the observation that, by following current
Several load regulatory wind profiles were adopted from the
codes and practices, the inelastic response does not necessarily
application of the codified design wind load. Calculations were
agree in many instances with the assumptions underlying the
performed to: (a) identify load–strain curves for elements, (b)
code’s or practical design expected behaviour, as it is reported
draw global load–deformation curves, (c) obtain yielding map-
in similar studies (Belloli, Rosa, & Zasso, 2014).
ping sequences and collapse mechanisms, (d) evaluate the influ-
Ductility and overstrength are presented in terms of the fun-
ence of the tower body deformation and (e) assess theoretical
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

damental period (or tower height), because it has been observed


overstrength and ductility capacities.
that there exists a relationship between the tower’s period (or
Results show a local concentration of the damage with
tower height) and the inelastic capacities. This observation is
non-uniform distribution of yielding within the height. The
related to the findings in previous studies, where the inelastic
distribution of energy dissipation is mostly produced by either
capacities (ductility and overstrength) under lateral load pat-
buckling in compression or yielding in tension of the leg ele-
terns, depend on the structure height (fundamental period)
ments, without subsequent generation of damage in brace
(Tapia-Hernández & Tena-Colunga, 2014). In fact, some interna-
elements. In fact, the bracing system remained elastic along anal-
tional codes (for example MFDC-04) establish the overstrength
yses under all wind patterns considered in this research. Thus,
factor for steel structures as a function of the fundamental period
collapse mechanism does not acceptably match the intended
for seismic design.
goals in the codes design philosophy, where it is expected that
Although the overstrength concept is routinely employed for
a large number of elements would develop a significant inelas-
predicting the load–deformation behaviour in seismic design,
tic response. This mechanism is not considered into the design
its application in wind engineering has not been sufficiently
process and, in fact, it is difficult to predict. No dependency on
explored. Perhaps this is partly due to the fact that the seismic
the recommended lateral load pattern has been noticed in the
design philosophy allows the structures to undergo a permanent
developed final mechanism.
inelastic deformation under severe seismic excitations, whereas
The reported damage concentration in the change of the cross
the design for wind actions requires the structures to remain into
section leads to a fragile collapse; although the structures have
the elastic range (Banik et al., 2008). According to the obtained
near-elastic responses and important strength reserves with a
tendency, if the tower is short in height (short period), a high
non-uniform inelastic response along the height. Therefore, on
overstrength ratio may be developed. In addition, this observa-
the basis of the two analysed cases, a weakness on the local design
tion notes that methods to obtain overstrength factors in codes,
process was identified. Similar damages on actual towers under
for steel structures, should consider the geometrical variation
intense wind were also reported. A proposal for a preliminary
of the section shape in order to avoid a very low probability of
assessment of the solidity ratio ϕ for such structures was sug-
yielding.
gested, which serves to estimate the magnitude of the wind forces
This relationship is not currently considered in structures
in new designs.
codes, because there are no guidelines for the designer to limit
The main contributions of this investigation to wind-resistant
the damage and prevent the subsequent collapse. More detailed
design of electrical transmission tower are as follows:
and conservative deformation limits are proposed with the
intention to avoid the fragile collapse and to guarantee that the • According to the analysis results of two transmission tow-
structure remains in a stable configuration, valid for the two ers under lateral wind loads, the response could not be
tower configurations studied. But further efforts are required in adequately estimated (collapse mechanism, overstrength
order to close the gap between the predicted behaviour in the and ductility capacities, maximum deformation, etc.) by
design stage and the actual behaviour, especially for the inelastic following only the current international design codes;
range. specifically they could not predict the actual capacity and
High-voltage transmission towers have a distinct structural some unfavourable effects. In fact, if towers were designed
behaviour under wind loads, as compared to building structures, with an enhanced capacity-based design approach, they
because they are typically designed considering relatively small could have a better inelastic performance; but their ine-
safety margins, despite their high exposure to intense winds due lastic capacities would also be difficult to predict from the
to the large transmission line extension (Holmes, 2007). So, in current normative tools and design practices.
addition to possible deficiencies in the design process, it is worth • Capacity curves and collapse mechanisms were very simi-
mentioning that the behaviour of transmission towers under lar regardless of the applied loading profile. Thus, the over-
such action is a complex task to assess due to different factors, strength factors (ratio of yielding to maximum capacity) are
as it is noted in this research. insensitive to the considered theoretical wind load patterns.
16    E. Tapia-Hernández et al.

• By following the current codes and practices, the designs wind loading. In Proceedings of BBAA VI, International Colloquium on
of the structures involve important unforeseen reserves Bluff Bodies Aerodynamics & Applications. Milano, Italy.
Banik, S., Hong, H., & Kopp, G. A. (2010). Assessment of capacity curves
of strength. This notes the importance to close the gap for transmission line towers under wind loading. Wind and Structures
between the predicted behaviour, in the design stage, and An International Journal, 13(1), 1–20.
the actual behaviour especially in the inelastic range of Belloli, M., Rosa, L., & Zasso, A. (2014). Wind loads on a high slender
transmission towers. tower: Numerical and experimental comparison. Engineering Structures,
• There was a fragile collapse mechanism developed in the 68, 24–32.
Bolivar-Villagómez E. (2006), Sistema Eléctrico Nacional ). International
towers, both models showed a near-elastic behaviour Electrotechnical Commitee. International Council on Large Electric
(ductility capacities are almost equal to μ ≈ 1). Despite Systems, Cigre.(pp. 1–57 www.cigre.org.mx/uploads/media/SIN_
this result being predictable, there are no guidelines for Mexico-CIGRE.pdf (in Spanish).
the designer to limit this kind of damage and to prevent CSI. (2009). SAP 2000 analysis software. Berkeley, CA: Computers and
the subsequent fragile collapse. The fact that the collapse Structures.
Da Silva, J. G. S., Vellasco, P. C. G., De Andrade, S. A. L., & De Oliveira, M.
is governed by a fragile response, for the considered con- I. R.. (2005). Structural assessment of current steel design models for
figuration types, might be taken into account by codes and transmission and telecommunication towers. Journal of Constructional
design practices through more detailed and conservative Steel Research, 61, 1108–1134.
deformation limits (as the proposed in this study) may Davenport, A. G. (1961). The application of statistical concepts to the wind
loading on structures. Proceedings of the Institution of Civil Engineers,
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

help to keep the structure safe and stable.


19, 449–472.
• Although several classical structures of power transmis- De Oliveira, A., Silva, E., De Medeiros, J. C. P., Loredo-Souza, A. M., Rocha,
sion tower are widely used all over the world, some design M. M., Rippel, L. I., Carpeggiani, E. A., & Núñez, G. J. Z. (2006). Wind
and construction procedures of the tower structures loads on metallic latticed transmission line towers (Paper B2-202). Paris:
seems to be unfavourable. The results of this investigation International Council on Large Electric Systems, Cigre.
argue the relevance of the capacity-based design approach Durst, C. S. (1960). Wind speeds over short periods of time. Meteorological
Magazine, 89, 181–186.
and that appropriate simple measures, for example the Filippou, F. C., Popov, E. P., & Bertero, V. V. (1983). Effects of bond
suggested connection between leg members (especially at deterioration on hysteretic behavior of reinforced concrete joints (Report
the middle height), could effectively increase the tower’s EERC 83-19). Berkeley, CA: Earthquake Engineering Research Center,
ability to survive under strong winds. University of California.
Holmes, J. D. (2007). Wind loading of structures (2nd ed.). London: Spon
Additional analyses of towers designed by following a detailed Press.
capacity-based design could complement, further improve the Hsiao, P.-C., Lehman, D. E., & Roeder, C. W. (2013). A model to simulate
understanding of the actual behaviour and prevent failures of special concentrically braced frames beyond brace fracture. Earthquake
Engineering & Structural Dynamics, 42, 183–200.
these structures. The influence of the soil – foundation – struc- IEA. (2012). Electricity information. Statistics. Paris: International Energy
ture interaction should also be explored. Caution should be Agency. ISBN 978-92-64-17468-9.
exercised, by designers, before the generalised application of the Ibarra-González, S. (2014). Estudio del mecanismo de colapso de torres de
above statements as they correspond to findings for the specific transmisión ante cargas laterales [Study of the collapse mechanism in
tower transmission lines under lateral loads] (Master thesis), Universidad
cases analysed here.
Autónoma Metropolitana – Azcapotzalco, Mexico City, Mexico (in
Spanish).
IS 875-95. (1992). Use of structural steel in over head transmission line
Acknowledgements
towers- code of practice (Materials, loads and permissible stresses). New
The master fellowship granted to the second author by the National Science Delhi: Bureau of Indian Standards.
and Technology Council of Mexico (Conacyt) is gratefully acknowledged. Jankowska-Sandberg, J., & Kolodziej, J. (2013). Experimental study of steel
truss lateral-torsional buckling. Engineering Structures, 46, 165–172.
Kemp, A. (1996). Inelastic local and lateral buckling in design codes.
Disclosure statement Journal of Structural Engineering, 22, 374–382.
Lee, P. S., & McClure, G. (2007). Elastoplastic large deformation analysis
No potential conflict of interest was reported by the authors. of a lattice steel tower structure and comparison with full-scale tests.
Journal of Constructional Steel Research, 63, 709–717.
Liang, S., Zou, L., Wang, D., & Cao, H. (2015). Investigation on wind
ORCID tunnel tests of a full aeroelastic model of electrical transmission tower-
Edgar Tapia-Hernández   http://orcid.org/0000-0003-4269-180X line system. Engineering Structures, 85, 63–72.
Mara T. (2013), Capacity assessment of a transmission tower under wind
loading (thesis and dissertation repository). University of Western
References Ontario, Ontario, Canada. Paper 1527.
Mara, T., & Hong, H. (2013). Effect of wind direction on the response and
AIJ-06. (2006). Recommendations for loads on building. Tokyo: AIJ capacity surface of a transmission tower. Engineering Structures, 57,
Architectural Institute of Japan. 493–501.
ASCE 7-10. (2010). Minimum design loads for buildings and other structures Mazzoni, S., McKenna, F., Scott, M., & Fenves, G. (2006). Open system for
(ASCE Standard ASCE/SEI 7-05). New York, NY: American Society of earthquake engineering simulation, user command-language manual
Civil Engineers. ISBN 07844-0809-2. (Report NEES grid-TR 2004–21). Berkeley, CA: Pacific Earthquake
AS/NZS-11. (2011). AS/NZS 1170.2. Structural design actions. Part 2: Wind Engineering Research, University of California.
actions (2nd ed.). Sidney: Australian/New Zealand Standard. ISBN 978- Metelli, G. (2013). Theoretical and experimental study on the cyclic
0-7337-9805-4. behavior of X braced steel frames. Engineering Structures, 46, 763–773.
Banik, S., Hong, H., & Kopp, G. A. (2008, July 20–24). Assessment of MFDC-04. (2004). Reglamento de Construcciones para el Distrito Federal
structural capacity of an overhead power transmission line tower under [Mexican Federal District Code]. Gaceta Oficial del Departamento del
Structure and Infrastructure Engineering   17

Distrito Federal [Official gazette of the Federal District Department]. Raychowdhury, P. (2011). Seismic response of low-rise steel moment frame
October. Mexico City, Mexico (in Spanish). (SMRF) buildings incorporating nonlinear soil-structure interaction.
Mitchel, D., Tremblay, R., Karacabeyli, E., Paultre, P., Saatcioglu, M., Engineering Structures, 33, 958–967.
& Anderson, D. (2003). Seismic force modification factors for the Savory, E., Parke, G. A. R., Zeinoddini, M., Toy, N., & Disney, P. (2001).
proposed 2005 edition of the National Building Code of Canada. Modelling of tornado and microburst-induced wind loading and failure
Canadian Journal of Civil Engineering, 30, 308–327. of a lattice transmission tower. Engineering Structures, 23, 365–375.
MOC-CFE-08. (2008). Manual de Diseño de Obras Civiles. Capítulo de Savory, E., Parke, G. A. R., Disney, P., & Toy, N. (2008). Wind-induced
Diseño por Viento. México: Instituto de Investigaciones Eléctricas, transmission tower foundation loads: A field study-design code
Comisión Federal de Electricidad (in Spanish). comparison. Journal of Wind Engineering and Industrial Aerodynamics,
NTCV-04. (2004). Normas Técnicas Complementarias para Diseño por 96, 1103–1110.
Viento [Design Guidelines for wind of the Mexican Federal District Tapia-Hernández, E., & Tena-Colunga, A. (2014). Code-oriented
Code]. Gaceta Oficial del Distrito Federal, décimo cuarta época, tomo II, methodology for the seismic design of regular steel moment resisting
octubre [Official gazette of the Federal District Department]. October. braced frames. Earthquake Spectra, 30(4), 1–27.
Mexico City, Mexico (in Spanish). Tapia-Hernández E., De Jesús Y., & Fernández-Sola L. (2016), Dynamic
Peyrot, A. H. (2009). Wind loading: Uncertainties and honesty suggest soil-structure interaction of ductile steel frames in soft soils. Advanced
simplification. In Proceedings, Electrical Transmission and Substation Steel Construction. Submitted for publication.
Structures Conference, (pp. 184–208). Fort Worth, TX: American Uriz, P., & Mahin, S. (2008). Toward earthquake-resistant design
Society of Civil Engineers. of concentrically braced steel-frames structures (Report of
Prasad, R. N., Knight, G. M. S., Mohan, S. J., & Lakshmanan, N. (2012). Pacific Earthquake Engineering Research Center, PEER 2008/08).
Studies on failure of transmission line towers in testing. Engineering Berkeley, CA: Pacific Earthquake Engineering Research, University of
Downloaded by [Universidad Autonoma Metropolitana] at 20:44 22 July 2016

Structures, 35, 55–70. California.

View publication stats

You might also like