You are on page 1of 9

Engineering Structures 143 (2017) 549557

Contents lists available at ScienceDirect

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

Performance-based design of steel towers subject to wind action


Rodolfo K. Tessari , Henrique M. Kroetz, Andr T. Beck
Department of Structural Engineering, University of So Paulo, Av. Trabalhador So-carlense 400, 13566-590 So Carlos, SP, Brazil

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

Article history: Performance-based Wind Engineering (PBWE) is a novel design philosophy that aims to identify and
Received 21 September 2016 quantify the uncertainties involved in structural design, in order to ensure predictable performance levels
Revised 23 March 2017 to engineering structures. Due to the recent proposal of the methodology and formulation complexity,
Accepted 26 March 2017
there are few studies related to PBWE, each presenting different limitations. This paper proposes an
Available online 29 April 2017
application of the Performance-based Wind Engineering methodology to the probabilistic analysis of
steel towers, evaluating different calculation models for the estimation of wind forces on this type of
Keywords:
structure. Uncertainties involved in the characterization of the wind field and the structural strength
Performance-based Engineering
Performance-based Wind Engineering
were investigated, and two procedures of the Brazilian winds standard NBR6123:88 for the estimation
Structural reliability of wind forces on steel towers were analyzed. A case study concerning the reliability estimation of a
Probabilistic analysis telecommunication tower was also conducted. It was found that both studied calculation models lead
Latticed telecommunication tower to similar safety levels, and that the design of towers considering that wind always blows from the worst
direction is too conservative. It is also shown that, in PBWE, minimum cost design can be guided by
assigning same target reliability, but different mean recurrence intervals for different performance levels.
2017 Elsevier Ltd. All rights reserved.

1. Introduction Classically, risk management in structural design problems is


addressed through deterministic or semi-probabilistic approaches
Inherently random and knowledge-based uncertainties are part (encoded in design standards). In the deterministic scenario, load
of the structural design process. Aleatory uncertainty is often asso- and resistance parameters are admitted as perfectly known and
ciated to environmental hazards (e.g. earthquakes, storms, land- invariant. In the latter, factors of safety calibrated according to tra-
slides and tornadoes), whose intensities and frequencies are ditional engineering practice are applied to the nominal or charac-
difficult to predict, while epistemic uncertainty encompasses lim- teristic values of design variables. Despite the advances in the field
ited databases and limited models that cannot assure perfect rep- of structural engineering during the past decades, both approaches
resentation of structural responses. Uncertainties can be reduced still present some flaws [25].
through research, but cannot be eliminated, giving rise to struc- The search for more effective methods for quantifying and mit-
tural risks. igating risks led to development of the Performance-based Engi-
Risk represents a function of the likelihood of a hazardous event neering (PBE) approach. The definition of PBE was first outlined
taking place, and its social and economic consequences. It strongly in SEAOCs Vision 2000 report [6] and soon became widespread
depends on location and typology of the concerned facility, since in the literature [79]. The PBE design philosophy represents a
any structure is subjected to a certain range of hazards. Therefore, paradigm in which the prescriptive approach imposed by struc-
selection of the relevant threats to each structure and acknowledg- tural standards is replaced by the quantitative assessment of
ment of the uncertainties involved is important and demands design alternatives against performance objectives, described in
attention. To neglect or to underestimate risks may compromise probabilistic terms [10].
structural safety, and may expose people to avoidable, dangerous The first publications regarding performance-based procedures
situations; whereas overestimating risks can lead to misallocation were developed in the United States for seismic design and retrofit
of resources [1]. of buildings [6,11]. These implementations were driven by the
1994 Northridge and 1995 Kobe earthquakes, that together caused
a total estimated loss of U$120 billion [12], even though affected
structures complied with seismic codes, based on traditional
Corresponding author. design philosophies, prevailing at the time [9].
E-mail addresses: tessari.rodolfo@usp.br (R.K. Tessari), henrique.kroetz@usp.br
(H.M. Kroetz), atbeck@sc.usp.br (A.T. Beck).

http://dx.doi.org/10.1016/j.engstruct.2017.03.053
0141-0296/ 2017 Elsevier Ltd. All rights reserved.
550 R.K. Tessari et al. / Engineering Structures 143 (2017) 549557

Naturally, the first-generation PBE procedures presented some vi. assessing the structural risk based on the probabilistic
shortcomings [13]. In order to fill these gaps, a more robust description of the DVs;
methodology was developed in the Pacific Earthquake Engineering vii optimizing design (i.e. minimizing risk or maximizing a util-
Research (PEER) Center, that turned out to be the most popular and ity function) by appropriate decision strategies.
currently applied Performance-based Earthquake Engineering
(PBEE) formulation [14]. The first step consists in choosing a set of intensity measure
PEERs PBEE methodology was built over the foundations of parameters (IMs) that are sufficient to describe the site-specific
structural reliability theory. This provided a theoretical framework hazard (e.g. mean wind velocity, turbulence intensity, direction).
to the probabilistic treatment of uncertainties and an explicit Attention must be paid to the selection of each IM, since they have
system-level performance assessment [9]. Its successful applica- direct impact on the output of all the subsequent stages.
tions made the technical and scientific community consider PEERs Step ii. corresponds to the choice and probabilistic characteriza-
rational approach ideal for the design of structures subjected to tion of a set of interaction parameters (IPs) able to represent the inter-
natural hazards. In recent years, many studies have been carried action between the environment and the structure. Aerodynamic
out on the performance assessment of structures subject to tsuna- coefficients and aeroelastic derivatives are proper examples of IPs.
mis [15] and hurricanes [2,16,17]. The process continues (step iii.) with selection of the most sig-
Another branch of great interest for further developments deals nificant random structural parameters (SPs) and engineering
with the application of PBE concepts to wind engineering. demand parameters (EDPs) that influence the structural behavior.
Performance-based Wind Engineering (PBWE) was first proposed Material and geometrical properties are normally chosen as SPs,
in 2004 [18] as the result of the Italian PERBACCO project, and whereas EDPs are represented by the acceleration, stress or dis-
assumed a concise and general format in 2009, when its methodol- placement at selected points. Such parameters are then probabilis-
ogy was established [19]. tically assessed through structural response analyses.
Recently, studies illustrating the applicability of PBWE method- In step iv., damage measures (DMs) and decision variables are
ology and proposing incremental improvements have been pub- specified in order to quantify the structural damage and the build-
lished: Ciampoli, Petrini and Augusti [20] evaluated the ing performance, respectively. Both are strongly interconnected
performance of a long span suspension bridge; Ciampoli and Pet- and depend on the considered facility type and usage. Typical
rini [21] analyzed the structural behavior of a 74 floor building; DMs are the loss of occupant comfort and the damage to structural
and Petrini, Gkoumas and Bontempi [22] performed an extensive and non-structural components in a building, due to excessive
literature review on damage and loss analysis in order to expand vibration or displacements. DVs include the number of casualties,
the PBWE procedure. Griffis et al. [23] also briefly proposed an the economic losses or some threshold that represents the collapse
alternative procedure for PBWE, focusing on nonlinear dynamic or loss of serviceability condition during windstorms.
analysis of structures. DVs which quantify the performance objectives must distin-
The rapid evolution and diffusion of the PBWE concept, over the guish between low- and high-performance levels [1]: the former
last years, is remarkable. However, due to the novelty and com- are related to ultimate limit states (ULS) and imply possible conse-
plexity of the methodology, there are still few studies related to quences on structural integrity and personal safety; the latter are
the theme, each presenting different limitations. No studies associated to serviceability limit states (SLS) and affect operability
addressing PBWE design of steel towers were found in the pub- and comfort.
lished literature. In this paper, the PBWE framework is adapted Regardless of the target performance, in the original PBWE pro-
to analyze the structural behavior of such structures in probabilis- cedure (as in the PEERs approach), the structural risk is defined as

tic terms. Special attention is given to the choice of proper param- the probability of a relevant DV exceeding a threshold value dv :
eters characterizing the wind field and structural behavior, as well  
as to the comparison of different wind action models for steel tow-
Gdv PDV > dv
RRRRR 
ers. The developed methodology is demonstrated in application to Gdv jDM  f DMjEDP  f EDPjIM; SP; IP
a popular telecommunications steel tower. f IPjIM; SP  f IM  f SP  dDM  dEDP  dIP  dIM  dSP
1

2. The PBWE methodology where f  symbolizes the probability density function, f j the
conditional probability density function, and G the complemen-
The present study is based on the Performance-based Wind tary cumulative distribution function.
Engineering framework proposed in [19,24]. The central objective Through the multiple integral expressed in Eq. (1), it is possible
of the PBWE procedure is to assess if a structural facility fulfills to separate the risk assessment into the previously stated elemen-
specific performance requirements (usually related to safety, func- tary steps:
tionality and comfort), as specified by end-users, stakeholders or
society. It consists of the following steps [25]:  site-specific hazard analysis and structural characterization, i.e.
the assessment of f IM and f SP, respectively;
i. characterizing the wind hazard at the candidate location of  interaction analysis, that corresponds to the estimation of
the structure; f IPjIM; SP;
ii. probabilistic modeling of the wind-structure interaction  structural analysis, aimed at assessing the probability density
phenomena; function of the structural response f EDPjIM; IP; SP, conditional
iii. analyzing the structural response within the scope of on the output of the previous steps;
stochastic mechanics;  damage analysis, that gives the damage probability density
iv. characterizing and evaluating the variables that express the function f DMjEDP conditional on EDP;

structural damage and govern the considered performance  finally, loss analysis, that matches the value of Gdv jDM.
measures;
v. defining the decision variables (DVs) that are appropriate to With respect to the PEER approach, the major advance provided
quantify the intended structural performances (mainly in by the PBWE methodology is the decoupled analysis of parameters
terms of consequences of damage); that characterize the structural behavior from parameters that
R.K. Tessari et al. / Engineering Structures 143 (2017) 549557 551

qualify the wind field. This is possible due to the strict dependency connected to the reference averaging time. Shorter averaging times
of the former on the candidate location of the structure, whereas (e.g. the 3 s gust speed provided by ASCE 798) have flatter distri-
the latter depends only on the design configuration. Moreover, butions than those based on 1 h mean wind velocity (like the Cana-
unlike earthquakes, winds have instability phenomena related to dian and Australian standards). This leads to a large scatter in the
fluidstructure interaction [26]. Thus, the introduction of the inter- predicted values of the wind load effects.
action analysis stage is essential to determine the relevant features Other studies related to the lack of harmonization and align-
of the wind effects. Hence, the current procedure allows to assess ment of wind codes and standards have similar findings [31,32].
sensitivity of each stochastic variable in the propagation of uncer- However, the coefficients of variation for both along-wind and
tainties, identifying its contribution to the failure probability of the cross-wind responses of high-rise building were surprisingly small
structure. It is also possible to check the response variability occa- (in the range of 1418%), considering the complexity of the calcu-
sioned by consideration of different aerodynamic and aeroelastic lations involved. The most plausible cause points to the fact that
phenomena. many procedures are inter-related and have a common source of
Given the complexity of the multi-dimensional integrals in Eq. origin.
(1), the complete PBWE procedure is rarely applied to practical As reported in [33], several international standards (e.g. NBC
design problems. Studies dealing with decision variables are usu- and ASCE 7-95) employ the Gust Response Factor (GRF) method
ally related to optimization problems. In such cases, DVs are suggested by Davenport [34], with slight modifications to deter-
rewritten in terms of the total expected cost of failure or the finan- mine the dynamic response of slender structures subjected to wind
cial loss rate for discrete levels of damage [27,28]. Most researches action. The method uses a statistical approach using influence
adopt assumptions that limit the quantification of risk to the like- lines, in which the total wind response is formed by a mean and
lihood of the structural response exceeding a certain threshold, a fluctuating component.
defined by particular engineering parameters. Ciampoli and Petrini The Brazilian wind code, NBR6123:88 [35], is also based on the
[24] argue that these simplifications are generally applied due to method of Davenport, but has minor differences in the parameters
the ethical and practical difficulties in accounting for intangibles that define the static wind loads. In addition, the structural vibra-
aspects of, for example, the casualties resulting from structural tion (in its natural modes) takes place around the deformed posi-
collapse. tion defined by the pressures caused by the static wind
Therefore, if the performance is expressed by the fulfillment of a component (mean speed).
limit state quantified in terms of EDP, the whole procedure can be NBR6123:88 provides two alternative design approaches: the
simplified. The structural risk is thus evaluated as the probability equivalent static wind loads (ESWL) method, which indirectly con-
of exceedance given by: siders dynamic features; and the discrete dynamic method. Fol-
Z Z Z lowing, both models are briefly presented.
 
Gedp Gedp jIM; SP; IP  f IPjIM; SP  f IM
3.1. ESWL method
 f SP  dIP  dIM  dSP 2
The procedure starts with calculation of the reference wind
speed V 0 , defined as the 3 s gust speed, determined in open terrain
3. Design alternatives for wind actions on towers exposure at an elevation of 10 m, with a mean return period of
50 years. The characteristic wind speed V k is then given by:
International design standards contain different, alternative
V k z V 0 S1 S2 z S3 m=s 3
procedures for calculating wind loads on structures. They follow
a general framework that usually begins with definition of the risk where S1 accounts for topographic features; S2 depends on the wind
category for a given structure. Different mean recurrence intervals exposure of the structure; and S3 considers the safety level required
are associated to each risk category, leading to the selection of a for the structure during its life-time.
proper basic wind speed according to the site of the facility. Based The S2 factor applies only to the ESWL method and describes the
on wind speed, wind loads are estimated incorporating factors that profile of the mean wind speed according to a power law:
account for wind exposure and topographic features. Once the
S2 z b C g z=10p 4
wind loads are defined, material codes provide detailed prescrip-
tive requirements which should be complied with. in which z is the height above ground, b and p are constants that
Although similarities can be identified among wind design depend on terrain category and building class; and C g is a gust fac-
codes, each one follows more or less complex, specific calculation tor that equals 1.0 for open terrains and class A buildings.
procedures. As Lungu, Gelder and Trandafir [29] pointed out, in With V k , one calculates the wind dynamic pressure q:
their comparative study of Eurocode 1, ISO DIS 4354 and ASCE
1
795, each employs unique definitions of wind field characteris- qz q V 2 z N=m2  5
2 a k
tics. The main differences between such documents include:
with air density qa being taken as 1.225 kg/m3 (at NTP).
 the mathematical law used to describe the mean wind speed Before proceeding, it is necessary to divide the structure of the
profile in the atmospheric boundary layer (logarithmic law or tower into panels or modules (usually defined by the region com-
power law); prised between consecutive horizontal bracing members), and
 the different averaging time intervals used to determine the ref- determine the aspect ratio of each module. The aspect ratio / is
erence or basic wind speed; the ratio between the total frontal area of all individual members
 the power spectral density functions of Davenport, Solari and (effective area), Ae , and the frontal envelope area (contour area), Ac .
von Karman for along-wind gustiness, related to wind The longitudinal drag forces F A on the towers modules are then
turbulence. determined by the product of the dynamic pressure, the effective
area and the drag coefficient C d , which is a function of the aspect
Regarding the wind speed, similar conclusions were drawn by ratio:
Zhou, Kijewski and Kareem [30]. Comparing five major interna- F A Ae  C d /  q N 6
tional codes, they noticed that the wind-speed profile is closely
552 R.K. Tessari et al. / Engineering Structures 143 (2017) 549557

Besides the forces acting directly on the tower, the wind load on ence height, taken as 10 meters; b and p are tabulated coefficients
additional structures such as ladders, platforms, cables and anten- that depend exclusively on the terrain roughness.
nas must be estimated. NBR6123:88 does not specify how these The fluctuating component, in turn, is given by:
additional forces should be determined, but the internal codes of
the former Brazilian telecommunications company specify guideli- X
n  p
nes for determining wind effect on accessories. According to TELE- C Ai Ai zzi lm
i
ref

F^i q0 b n mi lm
2 i1
BRAS codes [36,37], wind forces acting on accessories are simply N 9
i X
n
added to those acting on the tower, without considering any mi lm 2
i
mutual interference. Such common approach has the benefit of i1
simplicity, but raises some criticism [38,39].
Once the wind loads acting on each module and its accessories
where n is the dynamic amplification factor, mi is the lumped mass
have been calculated, the resulting force must be distributed
of module i and lmi is the modal coordinate i corresponding to the
according to the panel height and the wind incidence angle (a).
natural frequency of vibration m.
The decomposition process varies according to tower geometry,
When more than one mode of vibration is retained in the solu-
with square and triangular cross-sections being the most common
tion (m > 1), the superposition of the fluctuating responses
shapes for latticed steel towers. Squared section towers have only th
two critical wind incidence directions, while triangular ones are obtained for the j mode (j 1; . . . ; m) must be computed using
more sensitive to a. Therefore, the decomposition process is illus- the root mean square (rms) criteria:
trated herein for a tetrahedral module. As shown in Fig. 1, F T is the v
total drag force acting on the tower module, with an incidence uX
u m
angle a. Half of F T is then applied to the upper nodes and half to F i F i t F^2j 10
the lower ones, as recommended by [36]. If the module has a pyra- j1

midal shape, F T can be decomposed proportionally to the influence


area of each node. F T =2 is further decomposed into forces F X and F Y
acting on each vertex of the lattice section, aligned with x and y
axes. These forces, in turn, depend on f X a and f Y a, which 4. Numerical application of PBWE
account, in an approximate way, for shielding caused by the lattice
structure itself [40]. As the need for cellular coverage increases, freestanding latticed
steel towers to support cellular and microwave antennas become
omni-present. A large number of such towers is located in highly
3.2. Discrete dynamic method
populated urban areas, raising concerns about the possible conse-
quences of structural collapse. Due to the lightweight of these
An alternative procedure presented by NBR6123:88 is the dis-
structures, and the rare occurrence of earthquakes in Brazil, wind
crete dynamic method. It is recommended for very flexible slender
forces are the primary concern in the design. In this context,
(especially weakly damped) structures which can develop signifi-
telecommunication towers are an ideal case-study for implemen-
cant fluctuating responses due to wind gust. In this method, the
tation of the PBWE methodology. Further advantage is the low
mean wind response F i is separated from the fluctuating response interference of other accidental loadings. Hereafter, the procedure
F^i , caused by the turbulent wind. Thus, the total wind force is given summarized in Section 2 is applied to the risk assessment of a
by: telecommunication tower, in the framework of PBWE.

F i F i F^i ; N 7

and the mean response expressed by: 4.1. Structural model


 2p
2 zi The numerical case-study example considered herein is a free-
F i q0 b C Ai Ai N 8
zref standing latticed steel tower, used to support cellular antennas.
The geometry of the structure, the position of the antennas and
where: q0 12 qa V 2p is the dynamic pressure, given in [N/m2 ]; other accessories, and the self-weight of the whole system were
V p 0; 69 V 0 S1 S3 is the design wind speed, in [m/s]; C Ai ; Ai and zi supplied by a manufacturing company.
correspond to the drag coefficient, the effective area and the height The tower is located in Florianpolis, SC, Brazil. It has a total
above ground of the tower module i, respectively; zref is the refer- height of 50 meters and triangular cross-section. It is formed by
two distinct segments: the bottom part (up to 42 m) forms a pyra-
mid frustum with an opening of 5.70 m at its base and 1.50 m at
the top; the upper part of the tower is 6.0 m high and has parallel
legs. The legs were made with omega shape profiles and the
bracing members with symmetric single angles, both in hot-
rolled ASTM A36 steel. The tower is divided into 26 sections, each
comprehending the region between consecutive horizontal
bracings.
The structural analyses were carried out on a 3D finite element
(FE) model, consisting of 267 nodes and 798 truss elements. In
total, 13 different cross-sections were used to discretize the bars,
in correspondence to the original design. Additional structures
were replaced by nodal loads.
The plan view of the tower with a schematic division of sections
Fig. 1. Resulting drag force on each module vertex according to the panel height and the layout of the antennas is shown in Fig. 2, along with the
and the wind incidence angle. implemented three-dimensional model.
R.K. Tessari et al. / Engineering Structures 143 (2017) 549557 553

where E and F y are the Young modulus and the yield strength of the
steel; A; Imin e Lef represent the cross-section area, the minimum
moment of inertia and the effective bar length; and Nbar is the nor-
mal force in the bracing member.
In fact, for low slenderness (i.e., short members), plastic buck-
ling can control the behavior of the bar. The interval for which
inelastic buckling shall be considered is defined to be
p
Lef =r 6 p 2E=F y , according to ASCE 10:97 [41]. In the present
work, this corresponds to a limiting value of 127. As almost all
members of the structural model have a higher slenderness, only
elastic buckling was considered.
Finally, the low-performance level is related to global stability
of the tower. It is assumed that global instability occurs if the buck-
ling or tensile stresses in a leg member are exceeded. The limit
state equations are similar to Eqs. (12) and (13), but N bar now rep-
resents the normal force in leg members:
Unfortunately, we could not find proper references to quantify
the costs associated to loss of performance. Hence, our study does
not include the DM integral in Eq. (1). The problem is solved in
terms of the probabilities associated to the three performance
levels described in this section. Eq. (2) is solved numerically by
means of common reliability analysis.

5.2. Uncertainty parameters

5.2.1. Wind field parameters


In the numerical analyses, the deterministic unpredictability of
both the magnitude and incidence direction of mean wind speeds
have been considered, as well as the roughness length of the
terrain.
Wind speeds (V 0 ) associated to a given mean return period are
nominal values. However, the maximum wind speed V max to be
observed in a fixed period (design live) is a random variable, whose
statistics are evaluated in reference to V 0 . In this work, discrete
time periods, varying from 1 to 50 years, were considered. Table 1
shows mean values and coefficients of variation (COV) of the max-
imum wind speeds associated to different return periods. These
Fig. 2. Three-dimensional and plan views of the tower.
statistics were calculated from wind speed records available in
the Wolfram Research database [42], corresponding to the city of
Florianpolis, SC (SBFL Station, 27.58 S 48.56 W). The data com-
prises a period of 42 years, extending from 1973 to 2015. Maxi-
5. Structural performance analysis
mum wind speeds were found to follow Gumbel (Type I)
extreme value distributions.
5.1. Performance requirements
The wind incidence angle a with respect to the global x-axis
(counted counterclockwise) is characterized by a uniform distribu-
The structural performance of the antenna-carrying tower is
tion, with lower and upper bounds equal to p=6 and p=2.
divided in discrete levels. Hence, for computational purposes, the
The variability of the roughness length z0 is indirectly consid-
EDP integral in Eq. (2) is replaced by a sum over three levels, as
ered by the exponent p of the power law that describes the mean
follows.
wind speed profile Eqs. (4), (8), and (9). It is modeled by a non-
The full operability of the antennas is characterized by the
dimensional Gaussian random variable with unitary mean and
allowable deflection of 1 400 0000 , according to [36]. This high-
COV equal to 0.047. This random variable becomes dimensional
performance level characterizes a service failure, and is written
by multiplication by the nominal value recommended in
in terms of the lateral displacement at the top of tower, repre-
NBR6123:88, according to the adopted wind model: 0.125 in the
sented by d:
static case; 0.23 in the dynamic case. More details can be found
g 1 X d  dX 1:45 m  dX 0 11 in [40].

An intermediate performance level is given by yielding or buck-


ling of any diagonal or horizontal bracing members. This is also Table 1
considered a service failure, requiring maintenance with the even- Statistical data of V max according to different mean return periods.
tual replacement of individual failed bars. This performance level is Variable Mean COV
expressed by:
V1 25.86 m/s 0.256
g 2 X Fy A  Nbar yielding or 12 V2 27.75 m/s 0.238
V 10 37.46 m/s 0.177
p2 EImin V 25 42.35 m/s 0.156
g 2 X Nbar buckling; 13 V 50 45.98 m/s 0.144
L2ef
554 R.K. Tessari et al. / Engineering Structures 143 (2017) 549557

5.2.2. Structural characterization


Two mechanical properties of the tower are considered as ran-
dom elements of the SP vector: the Young modulus E, assumed to
be Gaussian with mean lE =197.4 GPa and COV = 0.076; and the
yield stress F y , characterized by a Lognormal distribution, mean
lFy = 277.5 MPa and COV = 0.068 [43].
In addition, a dead load factor, D, was also incorporated in the
analysis to account for possible deviations in estimating the struc-
tural self-weight. According to [44], D is modeled by a Normal dis-
tribution, with biased mean of 1.05 and COV equal to 10%.

5.2.3. Windstructure interaction


The last primary source of uncertainties is related with the
physical mechanisms (aerodynamic and aeroelastic phenomena)
which arise when the structure is immersed in the wind field.
The gust factor C g was chosen as the only random component of
the interaction parameter vector. It is normally distributed, with
nominal mean equal to 0.98 (as given by NBR6123:88) and COV Fig. 3. Reliability indexes for incidence angle of 30.
of 0.125 [45]. It reflects the epistemic uncertainty involved in the
conversion of dynamic wind effects into static equivalents, in the
ESWL method.

6. Numerical results

A set of 40 scenarios were analyzed, combining the two wind


load models described in Section 3, five mean wind speed values,
and four different assumptions w.r.t. wind incidence angle: first,
a was considered random; latter, the deterministic values of 30,
60 and 90 (critical directions for triangular section towers) were
assumed. Table 2 summarizes the considered scenarios.
For each scenario, reliability index and corresponding failure
probability of the structure were estimated using the First Order
Reliability Method (FORM) and crude and importance sampling
Monte Carlo simulation techniques (both combined with Latin
Hypercube stratified sampling). Twenty numerical analyses were
carried out for each load model, each composed of 15104 crude
Monte Carlo and 3104 importance sampling Monte Carlo simula-
tions. Sensitivity analyses were also conducted to assess the most Fig. 4. Reliability indexes for incidence angle of 60.
important variables to engineering problems involving wind action
on steel towers.

6.1. Results

Results of the numerical simulations are reported in Figs. 36.


The reliability index b is plotted against maximum wind speeds,
according to wind incidence angles. In the graphics, SLS1 refers
to the deflection serviceability limit state, SLS2 to the SLS related
to buckling and yielding of bracing members, and ULS refers to
the ultimate limit state. Distinction is also made between the wind
load models analyzed: the static (STA) model is represented by
gray curves, while the dynamic model (DYN) is indicated by black
curves. The output of each reliability method is indicated by the
following symbols: () FORM, (+) Crude Monte Carlo and ()
Importance Monte Carlo.

Table 2
Numerical simulation scenarios.

Load model Wind max. speed Incidence angle Fig. 5. Reliability indexes for incidence angle of 90.

V1
Static V2 a 30
V 10 a 60 Since the towers design was based on a basic wind speed of
Dynamic V 25 a 90 30 m/s for a return period of 50 years (much lower than the nearly
V 50 a = R.V.
46 m/s used in this study), the estimated failure probabilities are
R.K. Tessari et al. / Engineering Structures 143 (2017) 549557 555

ing performance may sometimes be equivalent to minimizing costs


and the optimal structure will be the one that minimize all costs
over its life-cycle (costs of construction, operation, inspection,
maintenance, and disposal), including expected costs of failure
[46,47].
The discussion in this session is limited to the balance between
expected failure costs for different performance measures. As fail-
ure costs can vary substantially depending on the location of the
tower (e.g., densely populated urban areas vs. uninhabited rural
areas) and the size of the structure, the proportionality between
ultimate and service failure costs is problem-specific. As men-
tioned above, no specific references to quantify expected costs of
failure for this communications tower were found. Assume, how-
ever, the cost of ultimate failure to be 25 times higher than the cost
for a bracing element service failure (SLS2). If reliability indexes for
these two performance levels were of similar magnitude, then
similar expected failure costs would be obtained for inversely
proportional return periods. Consider, for instance, the mean
Fig. 6. Reliability indexes for random incidence angle. return period of two years for a SLS2 service failure, and the mean
50-year return period for the ultimate failure, for a random inci-
dent angle (Fig. 7). By comparing the reliability indexes estimated
high for a civil structure. However, the obtained data supports the for these performance levels, it is evident that the probability of
comparisons drawn. operational failures is much lower than the probability of
structural collapse (b2yr 50yr
SLS2 2:46 and bULS 1:59). Hence, the ULS
6.1.1. Discussion of wind load models
dominates expected costs of failure. This situation is typical of
Results presented in Figs. 36 show that the static load model is
structures designed in accordance to ultimate limit states, follow-
more conservative, leading to lower reliability indexes, especially
ing normative prescriptions. Such difference allows concluding
for the high-performance requirement (deflection limit state). This
that it could be cheaper to reduce the cross-sectional area of brac-
means that structures designed according to the ESWL method, in
ing bars, in order to attain a better design (lighter structure with
the deflection limit state, will be generally more conservative,
more balanced expected costs of failure).
comparing to those designed by the dynamic wind load model.
However, the difference in results reduces for lower performance
6.1.5. Sensitivity analysis
levels. Hence, it may be concluded that both NBR6123:88 proce-
Through the use of FORM, a sensitivity analysis is conducted, in
dures lead to structures of similar safety levels.
order to investigate propagation of the uncertainties affecting the
input parameters (IMs, SPs and IP) to the structural response.
6.1.2. Discussion of wind directionality
Figs. 811 show the sensitivity indexes for the most significant
From Figs. 36, one observes that the low performance level
variables, for each analysis scenario (remaining variables are omit-
was the only one to show large variability depending on the inci-
ted for clarity). It is clear that the relative importance of the vari-
dent wind direction. As expected, the angle of 90 (Fig. 5) proved
ables remained almost constant over the various scenarios for
to be the most critical to structural stability, presenting the lowest
wind incidence direction.
reliability levels amongst all cases. However, the assumption that
A preponderance of the maximum wind speed over the other
the wind component with the highest speed blows from the most
variables can be observed, especially for the dynamical load model.
unfavorable direction is shown to be quite conservative: from
In the static model, the gust factor appears with growing impor-
Fig. 6, which admits a random incidence direction, it can be seen
tance, as the wind speed increases, becoming the second parame-
that the calculated reliability indexes resemble the case where
ter of greater contribution to failure probabilities. Uncertainties in
a 60 (Fig. 4).
The limit state SLS1 remained invariant w.r.t. the wind inci-
dence direction, since calculation of the peak displacement at the
Incidence Angle (=RV)
top of the tower was defined as the vector sum of the translations 3.3
of each degree of freedom of the top nodes.
Reliability index,

For SLS2, the instability of either element types masked the


importance of a, since no distinction was made between buckling 2.6
or yielding of horizontal or diagonal bracing bars. 2.46
1.9
6.1.3. Discussion of largest wind speed 1.59
Figs. 36 show that there is an approximately linear relation-
ship between reliability indexes and demand wind speeds. This 1.2
implies that b indexes for different wind speeds could be estimated
from a simple (almost linear) regression.
0.5
V2 V50
24 30 36 42 48
6.1.4. Discussion of performance levels
Wind speed, m/s
The last step of the Performance-based Design procedure corre-
SLS2 - STA SLS2 - DYN
sponds to structural design optimization. This can be attained by
ULS - STA ULS - DYN
minimizing risk or maximizing a utility function tied to the build-
ing performance. For structural engineering systems, (reduced) Fig. 7. Comparison between the SLS2 reliability index for a period of 2 years (V 2 )
cost is a frequently used measure of performance. Hence, maximiz- with the ULS value for a period of 50 years (V 50 ) when aw =RV.
556 R.K. Tessari et al. / Engineering Structures 143 (2017) 549557

Fig. 8. Random parameter sensitivity for the static (left) and dynamic (right) load
models. Wind incidence direction equal to 30.
Fig. 11. Random parameter sensitivity for the static (left) and dynamic (right) load
models. Wind incidence direction taken as aleatory.

100 %
11.7 12.4 18.3 23.0
80 27.3
6.4
8.1
60

40 83.9 82.7
74.3 67.6 61.1
20

0
V1 V2 V10 V25 V50
Speed (V) Angle () Gust (Cg)
(a) ESWL method

Fig. 9. Random parameter sensitivity for the static (left) and dynamic (right) load 100 %
models. Wind incidence direction equal to 60. 7.0 9.4
80

60
95.2 94.7 91.9 89.3
40 86.0

20

0
V1 V2 V10 V25 V50
Speed (V) Angle ()
(b) Discrete dynamic method
Fig. 12. Random parameters sensitivity for the low performance level and wind
incidence direction taken as aleatory.

Looking closely at Fig. 11, the findings about the wind direction-
ality become apparent, since only the ultimate limit state is found
Fig. 10. Random parameter sensitivity for the static (left) and dynamic (right) load to be sensitive to changes in a. Fig. 12 exhibits the sensitivity
models. Wind incidence direction equal to 90. indexes based solely on ULS, highlighting the growing importance
of the a angle, as the maximum wind speed increases.

the steel yield strength F y , the dead load factor D and the exponent 7. Concluding remarks
p resulted practically irrelevant, allowing the replacement of these
random variables by their means. The Young modulus had a rela- In this study, the PBWE methodology was applied to the prob-
tively low importance, exhibiting sensitivity indexes in the range abilistic assessment of the performance of steel towers. Two differ-
of 1.55.7%. ent models to determine the wind action on latticed towers were
R.K. Tessari et al. / Engineering Structures 143 (2017) 549557 557

considered: a static and a dynamic model. Through a case study [20] Ciampoli M, Petrini F, Augusti G. Performance-based wind engineering:
towards a general procedure. Struct Saf 2011;33(6):36778.
involving a telecommunications tower, it was found that both
[21] Ciampoli M, Petrini F. Performance-based aeolian risk assessment and
models lead to very similar safety levels. The assumption that wind reduction for tall buildings. Probab Eng Mech 2012;28:7584.
always blows from the most unfavorable direction proved to be [22] Petrini F, Gkoumas K, Bontempi F, Damage and loss evaluation in the
excessively conservative. Structural parameters SP such as steel performance-based wind engineering. In: Safety, Reliability, Risk and Life-
Cycle Performance of Structures and Infrastructures Proceedings of the 11th
yield strength, dead load factor and exponent p, frequently mod- International Conference on Structural Safety and Reliability, ICOSSAR 2013,
eled as stochastic, showed little contribution to failure probabili- 2013. pp. 17911797.
ties, given the dominance of wind field parameters. [23] Griffis L, Patel V, Muthukumar S, Baldava S. A framework for performance-
based wind engineering. In: Advances in Hurricane Engineering: Learning
The case study presented herein allowed us to illustrate the from Our Past - Proceedings of the 2012 ATC and SEI Conference on Advances
potential of the PBWE probabilistic approach for the design of steel in Hurricane Engineering, 2013. pp. 12051216.
towers. By comparing the reliability index associated with differ- [24] Ciampoli M, Petrini F. Performance-based design of structures under aeolian
hazard. In: Proceedings of the 11th International Conference on Applications
ent hazard intensities, it was shown how specific design strategies of Statistics and Probability in Civil Engineering, 2011. pp. 899906.
can be drawn, in order to optimize structural performance. By [25] Petrini F, Ciampoli M. Performance-based wind design of tall buildings. Struct
associating different performance levels to different mean return Infrastruct Eng 2012;8(10):95466.
[26] Petrini F, Ciampoli M, Augusti G. The role of uncertainties in aeolian risk
periods, and by requiring similar failure probabilities, one can assessment. Computational Methods in Stochastic Dynamics, Vol. 22 of
avoid dominance of one performance measure over the others, in Computational Methods in Applied Sciences. Springer; 2011. p. 187208.
terms of consequences of failure. http://dx.doi.org/10.1007/978-90-481-9987-7.
[27] Beck AT, Kougioumtzoglou IA, dos Santos KRM. Optimal performance-based
design of non-linear stochastic dynamical rc structures subject to stationary
Acknowledgments wind excitation. Eng Struct 2014;78:14553. http://dx.doi.org/10.1016/j.
engstruct.2014.07.047. URL http://www.sciencedirect.com/science/article/pii/
The authors kindly acknowledge financial support of this S0141029614004684.
[28] Li G, Hu H. Risk design optimization using many-objective evolutionary
research project by Brazilian agencies CNPq (National Council for algorithm with application to performance-based wind engineering of tall
Research and Development) and CAPES (National Council for buildings. Struct Saf 2014;48:114.
Post-Graduate Education). [29] Lungu D, Gelder PV, Trandafir R. Comparative study of eurocode 1, iso and asce
procedures for calculating wind loads; 1996.
[30] Zhou Y, Kijewski T, Kareem A. Along-wind load effects on tall buildings:
References Comparative study of major international codes and standards. J Struct Eng
ASCE 2002;128(6):78896. http://dx.doi.org/10.1061/(asce)0733-9445(2002)
[1] Augusti G, Ciampoli M. Performance-based design in risk assessment and 128:6(788). URL <GotoISI>://WOS:000175745800012.
reduction. Probab Eng Mech 2008;23(4):496508. http://dx.doi.org/10.1016/j. [31] Holmes JD. Developments in codification of wind loads in the asia pacific. In:
probengmech.2008.01.007. URL <GotoISI>://WOS:000259894400017. 7th Asia-Pacific Conference on Wind Engineering, APCWE-VII; 2009.
[2] Ellingwood BR. Proability-based codified design: past accomplishments and [32] Holmes J, Tamura Y, Krishna P, Comparison of wind loads calculated by fifteen
future challenges. Struct Saf 1994;13(3):15976. http://dx.doi.org/10.1016/ different codes and standards, for low, medium and high-rise buildings. In:
0167-4730(94)90024-8. URL <GotoISI>://WOS:A1994NE27700003. 11th Americas Conference on Wind Engineering; 2009.
[3] Hendawi S, Frangopol DM. System reliability and redundancy in structural [33] Merce RN, Guimares MJR, Doz GN, Brito JLV. Anlise de torres metlicas
design and evaluation. Struct Saf 1994;16(12):4771. http://dx.doi.org/ submetidas ao do vento: um estudo comparativo. Revista Sul-Americana
10.1016/0167-4730(94)00027-n. URL <GotoISI>://WOS:A1994PQ24900006. de Engenharia Estrutural 2007;4(1):6181.
[4] Ellingwood BR, Tekie PB. Wind load statistics for probability-based structural [34] Davenport A. The response of slender structures to wind. Wind Climate in
design. J Struct Eng ASCE 1999;125(4):45363. http://dx.doi.org/10.1061/ Cities, Vol. 277 of NATO Advanced Study Institute Series. Springer; 1995. p.
(asce)0733-9445(1999)125:4(453). URL <GotoISI>://WOS:000079280500011. 20939. http://dx.doi.org/10.1007/978-94-017-3686-2.
[5] Zhang H, Ellingwood BR, Rasmussen KJR. System reliabilities in steel structural [35] Associao Brasileira de Normas Tcnicas, Rio de Janeiro, NBR 6123. Foras
frame design by inelastic analysis. Eng Struct 2014;81:3418. devidas ao vento em edificaes; 1988.
[6] Blue Book, Sacramento, CA, 1999. [36] Telecomunicaes Brasileiras S/A, Prtica 240410-600. Procedimentos de
[7] Ghobarah A. Performance-based design in earthquake engineering: projeto para torres metlicas auto-suportadas, estaiadas e postes metlicos;
state of development. Eng Struct 2001;23(8):87884. URL <GotoISI>:// 1997.
WOS:000169055000001. [37] Telecomunicaes Brasileiras S/A, Prtica 240400-702. Especificaes gerais
[8] Bertero RD, Bertero VV. Performance-based seismic engineering: the need for a para adoo de parmetros bsicos e apresentao de memorial de clculo
reliable conceptual comprehensive approach. Earthquake Eng Struct Dyn para torres e postes metlicos; 1997.
2002;31(3):62752. http://dx.doi.org/10.1002/eqe.146. [38] Holmes JD, Banks R, Roberts G. Drag and aerodynamic interference on
[9] Lee T-H, Mosalam KM, Probabilistic seismic evaluation of reinforced concrete microwave dish antennas and their supporting towers. J Wind Eng Ind
structural components and systems, PEER Report 2006/04. Aerodyn 1993;50:2639.
[10] Aktan AE, Ellingwood BR, Kehoe B. Performance-based engineering of [39] Carril-Jr CF, Isyumov N, Brasil R. Experimental study of the wind forces on
constructed systems 1. J Struct Eng 2007;133(3):31123. rectangular latticed communication towers with antennas. J Wind Eng Ind
[11] Federal Emergency Management Agency, Washington, D.C., FEMA-273. NEHRP Aerodyn 2003;91(8):100722. http://dx.doi.org/10.1016/s0167-6105(03)
guidelines for the seismic rehabilitation of buildings; 1997. 00049-7. URL <GotoISI>://WOS:000185717100004.
[12] USGS, Significant earthquakes archives. earthquake hazards program. [40] Tessari RK. Performance-based design of steel towers subjected to wind
apresenta informaes acerca dos terremotos mais significativos ocorridos action. So Carlos, SP: University of So Paulo; 2016. Master thesis.
at o presente momento. disponvel em: http://earthquake.usgs.gov/ [41] American Society of Civil Engineers, Reston, VA, ASCE 10:97. Design of Latticed
earthquakes/eqarchives/significant/; 2014. Steel Transmission Structures; 1997.
[13] Porter KA. An overview of peers performance-based earthquake engineering [42] W. R. Inc., Mathematica 9.0 (2012). URL http://www.wolfram.com/.
methodology. Appl Stat Probab Civil Eng 2003;1 and 2:97380. URL [43] Hess PE, Bruchman D, Assakkaf IA, Ayyub BM. Uncertainties in material and
<GotoISI>://WOS:000189453300130. geometric strength and load variables. Naval Eng J 2002;114(2):13965.
[14] Gunay S, Mosalam KM. Peer performance-based earthquake engineering met- http://dx.doi.org/10.1111/j.1559-3584.2002.tb00128.x. URL <GotoISI>://
hodology, revisited. J Earthquake Eng 2013;17(6):82958. http://dx.doi.org/ WOS:000179964600009.
10.1080/13632469.2013.787377. URL <GotoISI>://WOS:000320571400003. [44] Ellingwood B, Galambus TV, MacGregor JG, Cornell CA. Development of a
[15] Robertson I, Riggs HR, Development of performance based tsunami Probability Based Load Criteria for American National Standard
engineering; 2011. URL https://nees.org/resources/1861. A58. Cambridge, MA: National Bureau of Standards; 1980.
[16] Barbato M, Petrini F, Ciampoli M. A preliminary proposal for a probabilistic [45] Joint Committee on Structural Safety, Denmark, JCSS Probabilistic model code;
performance-based hurricane engineering framework. Structures Congress 2001. URL http://www.jcss.ethz.ch/.
2011 Proceedings of the 2011 Structures Congress. p. 161829. [46] Beck AT, de Santana Gomes WJ. A comparison of deterministic, reliability-
[17] Barbato M, Petrini F, Unnikrishnan VU, Ciampoli M. Performance-based based and risk-based structural optimization under uncertainty. Probab Eng
hurricane engineering (pbhe) framework. Struct Saf 2013;45:2435. Mech 2012;28:1829.
[18] Paulotto C, Ciampoli M, Augusti G. Some proposals for a first step towards a [47] Beck AT, Gomes WJ, Lopez RH, Miguel LF. A comparison between robust and
performance based wind engineering. In: Proceedings of the IFED international risk-based optimization under uncertainty. Struct Multidiscip Optim 2015;52
forum in engineering decision making; First Forum, December, 59, 2004. (3):47992.
[19] Petrini F. A probabilistic approach to performance-based wind engineering
(pbwe). Dottorato di ricerca: Universit degli Studi di Roma La Sapienza,
Rome, Italy; 2009.

You might also like