You are on page 1of 13

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

net/publication/278726673

Performance-Based Topology Optimization for Buildings under Wind and


Seismic Hazards

Conference Paper · April 2015


DOI: 10.1061/9780784479117.192

CITATIONS READS
4 436

4 authors:

Sarah Bobby Seymour M.J. Spence


AIR Worldwide University of Michigan
19 PUBLICATIONS   139 CITATIONS    72 PUBLICATIONS   497 CITATIONS   

SEE PROFILE SEE PROFILE

Enrica Bernardini Ahsan Kareem


University of Michigan University of Notre Dame
28 PUBLICATIONS   195 CITATIONS    449 PUBLICATIONS   8,138 CITATIONS   

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Sarah Bobby on 16 August 2015.

The user has requested enhancement of the downloaded file.


Performance-Based Topology Optimization for Buildings under Wind and
Seismic Hazards

S. Bobby1, S.M.J. Spence2, E. Bernardini3 and A. Kareem4


1
NatHaz Modeling Laboratory, University of Notre Dame, 156 Fitzpatrick Hall,
Notre Dame, IN 46556; PH (574) 631-2540, email: sbobby@nd.edu
2
Department of Civil & Environmental Engineering, University of Michigan, Ann
Arbor, MI 48109; PH (734) 764-8419, email: smjs@umich.edu
3
Formerly at NatHaz Modeling Laboratory, University of Notre Dame, Notre Dame,
IN 46556; PH (574) 323-5293, email: enricabernardini14@gmail.com
4
NatHaz Modeling Laboratory, University of Notre Dame, 156 Fitzpatrick Hall,
Notre Dame, IN 46556; PH (574) 631-6648, email: kareem@nd.edu

ABSTRACT

Topology optimization has recently been explored as a tool for the conceptual design
of efficient structures during the initial building design stage. Although these
techniques were originally developed in a deterministic setting most practical
applications are subject to inherent uncertainties in the design problem (e.g. stochastic
loads, model idealization) and thus there has recently been an effort to incorporate the
effect of these uncertainties into the topology optimization procedure. Within this
context the authors have developed an approach that is able to efficiently account for
the uncertainties in the problem at hand using concepts from Performance-Based
Design. This paper presents a general overview of this method and illustrates its
applicability for the topology design of structures under wind and seismic hazards.

INTRODUCTION

Topology optimization techniques are quickly becoming recognized as a powerful


tool for conceptual building design due to their ability to systematically explore the
structural design space to discover efficient, novel structural layouts (Bobby et al.
2014a, 2014b; Kareem et al. 2013, 2014). These techniques explore variations in both
the boundary and connectivity of the design domain (Bendsøe and Sigmund 2004)
and thus are more versatile for conceptual design problems than either shape
optimization techniques, which explore variations in the boundary of the system but
do not alter its connectivity, or sizing optimization techniques, which may be used to
resize existing members (Christensen and Klarbring 2008). Thus far, the majority of
research considering the conceptual design of buildings using topology optimization
techniques has been formulated in a deterministic setting, e.g. that performed by
Neves et al. (1995), Mijar et al. (1998), Liang et al. (2000), and Stromberg et al.
(2011, 2012), to name a few. However, as a number of uncertainties exist in the
structural design problem, including those in structural parameters and model
idealization as well as the inherently stochastic nature of the environmental loads
acting on the structure (Spence and Kareem 2014), the effect of these must be
considered during the design process in order to ensure the performance of the
structure. This is especially true when using topology optimization techniques to
increase the structural efficiency, as topology optimization methods by nature
eliminate unnecessary material and push the structure to its capacity. It is therefore
possible that neglecting the uncertainties in the problem at hand could result in an
optimal structure that is inadequate from a safety or serviceability standpoint.
The prescriptive building design procedure incorporates the effect of these
uncertainties using safety factors that are chosen to ensure structures meet target
reliability levels for critical performance measures (Ellingwood and Tekie 1999).
However, these do not explicitly ensure the reliability of a particular structure
(Augusti and Ciampoli 2008, Spence and Kareem 2014) and therefore it is possible
that the structural performance could be unsatisfactory. Recently, probabilistic
Performance-Based Design (PBD) methods have been developed that are rigorously
based on principles of structural reliability and are able to ensure target reliability
measures are met for specific structures (Cornell and Krawinkler 2000, Porter 2003,
Ciampoli et al. 2011, Spence and Kareem 2014). Within the field of topology
optimization, Reliability-Based Topology Optimization (RBTO) methods have been
developed to explicitly account for the effect of uncertainties during the layout
optimization process and could be extended for topology optimization within a PBD
setting. RBTO has been developed relatively recently and has focused primarily on
static problems (Maute and Frangopol 2003; Kharmanda et al. 2004; Silva et al. 2010;
Nguyen et al. 2011). Although the method developed by Chun et al. (2012) does
consider stochastic excitation, it is able to ensure requirements on only the
instantaneous failure probability instead of the more desirable first excursion
probability. However, a recently proposed method developed by the authors is able
to determine novel topologies that effectively meet probabilistic constraints written in
terms of first excursion probabilities. This methodology promises to create new
practical design avenues due to its efficiency and has been shown to yield efficient
structural layouts under both wind and seismic hazards (Bobby et al. 2014b, 2014c).
This paper illustrates the key points of this framework and demonstrates its
applicability for the conceptual layout design of structures under wind and seismic
hazards using several case studies.

PERFORMANCE-BASED TOPOLOGY OPTIMIZATION

Problem Statement. The performance-based topology optimization method


presented in this paper explicitly accounts for the stochastic and time-invariant
uncertainties affecting the building design problem by ensuring that probabilistic
performance requirements are met for the optimum structural layout. The actual
problem of interest can be formulated by assuming that the design domain is
discretized using finite elements and written as follows:
n
min V ( ρ ) = ∑ ∫ ρe dΩ
ρ
e =1 Ωe

s.t. λ ( dm j ) − λ0 j ≤ 0, j = 1,..., N (1)


z ( t ) + C ( ρ, U ) z& ( t ) + K ( ρ, U ) z ( t ) =f ( t , U, w ( t ) )
M ( ρ, U ) &&
0 ≤ ρe ≤ 1
where ρ={ρ1,…,ρn}T is the element-wise normalized material density design variable
vector, n is the total number of elements composing the discretized design domain, Ωe
denotes the domain of element e, V is the volume of material in the design domain,
λ(dmj) is the mean annual rate of exceedance of the damage threshold dmj of the jth
probabilistic performance constraint, λ0j is the target mean annual exceedance rate for
the jth performance constraint, N is the total number of performance constraints, U is
a vector containing the uncertain model parameters, z is the displacement vector, M,
C, and K are the mass, damping, and stiffness matrices, respectively, and f is the
time-dependent external loading vector that is dependent on both U and a vector-
valued stationary random process w(t) where t indicates the dependence on time.

Probabilistic Performance Constraints. The probabilistic performance constraints


of Eq. (1) are written in terms of the target mean annual exceedance rate of a damage
measure, which can be estimated using a method developed by researchers at the
Pacific Earthquake Engineering Research (PEER) Center. The PEER method uses
the Total Probability Theorem to develop a convenient formulation for the mean
annual rate of exceedance of a specified performance level using several sub-tasks
that are independent of each other and can be carried out sequentially (Porter 2003).
This method is attractive to use during the topology design process because the sub-
tasks are of interest to end users and experts in different fields of engineering and as
such the sub-tasks are meaningful from a design perspective. Using the PEER
methodology, the mean annual rate of exceedance of a damage measure can be
written in the following way:

λ ( dm ) = ∫ ∫ G ( dm | edp ) ⋅ dG ( edp | im ) ⋅ dλ ( im )
edp im

=η ∫ ∫ G ( dm | edp ) ⋅ f ( edp | im ) ⋅ f ( im ) dim dedp


edp im
(2)

=η ⋅ P ( DM > dm )
where dm denotes a damage measure indicating damage, edp indicates an engineering
demand parameter characterizing the structural response, im indicates an intensity
measure describing the intensity of the hazard at the site of interest, λ (a) is the mean
annual rate of exceedance of event A=a, where capital letters indicate random
variables and lowercase letters indicate their realizations, G(a|b) = P(A > a | B = b)
denotes the complementary cumulative distribution function of random variables A
given B = b, f(a) is the probability density function of a, and η is the mean annual rate
of arrival of significant events. As can be seen in Eq. (2), the various terms are
independent of each other; these can be obtained through separate analyses (see
Figure 1): hazard analysis for λ(im), structural analysis for G(edp|im), and damage
analysis for G(dm|edp).

Figure 1: Schematic illustrating evaluation of performance-based constraints

Uncertainty Models. A number of uncertainties must be considered during the


design process to ensure acceptable structural performance. The PBTO method
described above is general in the sense that it allows the most suitable uncertainty
models to be used during evaluation of the sub-tasks in order to ensure that the
performance of the structure is appropriately described. A number of uncertainty
models have been developed in the literature and utilized by the authors in this PBTO
framework; their use in evaluating the subtasks given in Eq. (2) will be described
here. However, as this framework is versatile these may be substituted if a different
model is preferred during the conceptual design process.
Hazard analysis is primarily concerned with the development of hazard
curves, which describe the probability of events with various intensity measures for a
certain time period (usually annually). The relationship between the intensity
measure and site-specific hazard is succinctly given by Der Kiureghian (2005). With
regard to wind hazards an obvious choice of intensity measure is the mean wind
speed at the top of the building, which can be estimated by extrapolating data from
nearby meteorological stations. One possible uncertainty model for the description of
this intensity measure is an extension of the logarithmic law as suggested by
Minciarelli et al. (2001). Diniz et al. (2004) also gives suggestions of distributions
for the parameters in this model. With regard to seismic hazards, the intensity
measure may be taken as the peak horizontal ground acceleration at the base of the
structure. Site-specific ground motions may be generated using stochastic ground
motion models, such as the point-source model described in Boore (2003). In order
to truly represent the variability observed in actual ground motions the parameters of
the point-source model may be taken as uncertain and chosen from appropriate
distributions, as suggested in Vetter and Taflanidis (2012, 2014). The site-specific
hazard curve may then be developed using information from the ground motion
simulations.
The structural analysis subtask aims to evaluate the probability that a
structural response of interest exceeds a user-defined threshold edp. As the structural
response is dependent not only on the uncertain parameters characterizing structural
properties and accounting for model idealization (suggested distributions for these
parameters are given by Bashor et al. (2005) and Spence and Kareem (2014),
respectively) but also on the external stochastic environmental excitation, the
structural analysis subtask requires the solution of the first excursion of the response
process over the edp threshold during a specific time duration. If the structural
response is evaluated during wind events, aerodynamic loads may be determined by
appropriately scaling measurements obtained via wind tunnel tests (Spence and
Kareem 2014). The typical assumptions of stationary wind excitation and linear
structural behavior may be invoked and therefore the resulting structural response
may be assumed to be stationary. The first excursion probability of the appropriate
response process may be estimated using classic results from random vibration theory
(Pinto et al. 2004). If the environmental excitation is due to earthquake events, the
ground accelerations, which may be generated using the models described in the
previous paragraph, and resulting structural response are non-stationary and thus the
first excursion probability cannot be determined using a closed-form solution.
However, if the time duration of interest is taken as the duration of the earthquake
event, what is actually of interest is the distribution of the largest response during this
event, which may be estimated using simulation techniques.
Damage analysis evaluates the probability that damage has occurred to the
structure given a particular value of edp. An intuitive measure describing the damage
may simply be defined using a ratio of the edp to the capacity associated with the
particular edp (Jalayer et al. 2007). In this case damage would correspond to dm≥1.
Often the actual capacity corresponding to damage is uncertain (e.g. the interstory
drift corresponding to partition damage). An uncertainty model for the capacity has
been suggested by Spence and Kareem (2014).

Solution Strategy. The PBTO problem of interest as formulated in Eq. (1) is


impractical to solve due to the large size of the design variable vector ρ and the
implicit dependence of the probabilistic performance constraints on ρ. In order to
efficiently solve the topology optimization problem a novel simulation-based
decoupling technique has been developed by the authors that simplifies the original
probabilistic optimization problem using a sub-problem defined from information
gathered from a single reliability analysis performed in the current design point. The
sub-problem replaces the original probabilistic constraints with constraints that take a
form similar to those of classic deterministic optimization problems; thus the sub-
problem may be efficiently solved using efficient gradient-based algorithms typically
used to solve deterministic problems. As the sub-problem is formulated from
information gathered from a complete reliability analysis, the relationship defined by
the simplified constraints rigorously considers the uncertainties in the original
problem formulation. The simplified constraints are exact for the current design point;
however as the layout of the structure is updated during the solution of the sub-
problem the constraints become approximate. Thus after the sub-problem has been
solved it must be updated via information gathered from a reliability analysis
performed in the new design point. This procedure is continued until convergence of
both the structural layout and the simplified constraints of the sub-problem. This
novel solution strategy greatly reduces the computational effort needed to solve the
PBTO problem, as the number of costly reliability analyses required for this solution
strategy is significantly curtailed. Figure 2 illustrates this decoupling technique.
Mathematical details are given in Bobby et al. (2014b, 2014c).

Figure 2: Solution strategy

CASE STUDIES

Several case studies will now be given to illustrate the use of the PBTO methodology
in the design of novel, efficient layouts for conceptual building design. Two case
studies will be presented in this section. The objective of both studies will be to
obtain an efficient layout for the bracing of a planar lateral load resisting system
envisioned as part of a 3D building. The Method of Moving Asymptotes was used to
solve the sub-problems of the PBTO method (Svanberg 1987).

Case Study 1. The aim of the first case study is the topology design of the bracing for
a tall building located on the eastern coast of the United States near the city of Miami,
Florida. The objective is to minimize the volume of material used for the bracing
while meeting probabilistic performance constraints describing nonstructural damage.
The dominant hazard in this area is characterized by hurricane events; as such the
models described by Minciarelli et al. (2001) and Diniz et al. (2004) are used to
model the wind hazard. The external loads applied to the structure were determined
by integrating and appropriately scaling pressure measurements obtained from wind
tunnel tests performed at the boundary layer wind tunnel at the Inter-University
Research Centre on Building Aerodynamics and Wind Engineering (CRIACIV)
(Bernardini et al. 2012, Spence et al. 2011) and applying the model described in
Spence and Kareem (2014). The engineering demand parameters of interest were
taken to be the interstory drifts at all floors of the structure and the damage parameter
was taken as that corresponding to the occurrence of non-structural damage. The
nominal capacity for this value was taken as 1/400th of the floor-to-floor height. The
target damage measure was taken to be dmj = 1 for all floors. Two trials are
presented: Trial 1 considers a target mean annual rate of exceedance of λ0j =0.02 and
Trial 2 considers a target mean annual exceedance rate of λ0j =0.002. The initial
design domain is shown by the shaded gray area of Figure 3(a), where the horizontal
and vertical black lines represent frame elements modeling the underlying secondary
system (e.g. floor system) that is used to transfer local loads to the main lateral load
resisting system. Figures 3(b) and 3(c) show the optimal bracings for Trials 1 and 2
respectively. Although a similar high-waisted X-bracing pattern is noticed for both
trials, Trial 2 shows noticeably thicker bracing due to the stricter requirement on the
target mean annual exceedance rate. Figure 4 illustrates trends regarding the
objective and constraint functions. The material volume history given in Figure 4(a)
shows rapid and steady convergence for both trials and illustrates the efficiency of the
method. Figure 4(b) plots the maximum mean annual exceedance rate over all
constraints for each design cycle, and it can be seen that at the final design cycle the
required target mean annual exceedance rate is met for both trials. Finally, Figure
4(c) gives the damage measures with target mean annual exceedance rate for the
initial and final structures for both trials. A damage measure of dm≤1 is shown for
the final structures at all stories, indicating that all constraints were met.

Figure 3: Case Study 1: (a) Initial design domain; Optimal topologies for
(b) Trial 1, (c) Trial 2
Figure 4: Case Study 1 (a) Volume history; (b) Maximum mean annual
exceedance rate over all constraints; (c) Damage measure with target mean
annual exceedance rate

Case Study 2. The second case study considers the topology design of a bracing
system for a three story building under earthquake loading. It is assumed that the
structure is located in California and the earthquake hazard is represented using the
models suggested by Boore (2003) and Vetter and Taflanidis (2012, 2014). As in the
first case study, the engineering demand parameters of interest are taken to be the
interstory drifts at all floors of the structure and the damage parameter was taken as
that corresponding to the occurrence of non-structural damage with a nominal
capacity of 1/400th of the floor-to-floor height. The damage measure was again taken
to be dmj = 1 for all floors. The target mean annual rate of exceedance was taken to
be λ0j =0.01, and earthquake events with magnitudes between 5 and 8 were
considered. The initial design domain is shown by the shaded gray area of Figure
5(a), where the secondary system of the structure is illustrated using horizontal and
vertical black lines representing frame elements. Figure 5(b) shows the optimal
bracing scheme after the problem is solved using the PBTO solution strategy. The
PBTO methodology produced an X-bracing scheme, which is encouraging because an
X-bracing scheme was also determined by Chun et al. (2012) for a structure having a
design domain with the same dimensions. Figure 6 illustrates trends regarding the
objective function and constraints for Case Study 2. The material volume history
given in Figure 6(a) shows rapid and steady convergence, which was achieved after
only 6 design cycles and illustrates the efficiency of the PBTO solution strategy. The
maximum mean annual exceedance rate over all constraints is given in Figure 6(b).
As can be seen, the target mean annual exceedance rate was met for the final design
at design cycle 27. Finally, Figure 6(c) gives the damage measures with target mean
annual exceedance rate for the initial and final structures of Case Study 2. The final
structure displays a damage measure of 1 for the constraints at all stories.
Figure 5: Case Study 2: (a) Initial design domain; (b) Optimal Topology

Figure 6: Case Study 2 (a) Volume history; (b) Maximum mean annual
exceedance rate over all constraints; (c) Damage measure with target mean
annual exceedance rate
CONCLUSIONS

This paper presented an overview of an efficient Performance-Based Topology


Optimization procedure for the conceptual topology (layout) design of buildings
under wind and seismic hazards. The topology optimization problem was formulated
within a setting consistent with that of Performance-Based Design and was solved
using a novel decoupling technique that rigorously considered the uncertain and
stochastic nature of the problem. One of the appeals of this framework is that the
methodology is structure-specific and the design team is able to choose the most
appropriate uncertainty models for a particular design problem at hand. Two case
studies were presented that illustrated the efficiency and applicability of the solution
strategy to structures under wind and seismic hazards. It was found that the PBTO
methodology was able to effectively determine efficient bracing layouts for both case
studies.

ACKNOWLEDGEMENTS

Support for this work was provided by the NSF Grant No. CMMI-1301008. The
authors also thank Krister Svanberg for providing the MMA algorithm, which was
used as the optimization algorithm in this research.

REFERENCES

Augusti, G. and Ciampoli, M. (2008). “Performance-based design in risk assessment


and reduction.” Prob. Eng. Mech. 23(4):496-508.
Bashor, R., Kijewski-Correa, T., and Kareem, A. (2005). “On the wind-induced
response of tall buildings: the effects of uncertainties in dynamic properties
and human comfort thresholds.” Proc. of the 10th Americas Conference on
Wind Engineering.
Bendsøe, M.P. and Sigmund, O. (2004). Topology Optimization: Theory, Methods,
and Applications. Springer: New York.
Bernardini, E., Spence, S.M.J., and Gioffrè, M. (2012). “Effects of the aerodynamic
uncertainties in HFFB loading schemes on the response of tall buildings with
coupled dynamic modes.” Engineering Structures 42:329-341.
Bobby, S., Spence, S.M.J., Bernardini, E., and Kareem, A. (2014a). “A performance-
based methodology for the topology design of tall buildings.” Proc. of
Structures Congress 2014, pp. 2512-2523.
Bobby, S., Spence, S.M.J., Bernardini, E., and Kareem, A. (2014b). “Performance-
based topology optimization for wind-excited tall buildings: A framework.”
Engineering Structures 74:242-255.
Bobby, S., Spence, S.M.J., Bernardini, E., and Kareem, A. (2014c). “Reliability-
based topology optimization for uncertain building systems in seismic zones.”
Proc. of Engineering Optimization 2014. CRC Press, pp.781-786.
Boore, D.M. (2003). “Simulation of ground motion using the stochastic method.”
Pure and Applied Geophysics, 160:635-676.
Christensen, P.W. and Klarbring, A. (2008). An Introduction to Structural
Optimization. Springer Science & Business Media.
Ciampoli, M., Petrini, F., and Augusti, G. (2011). “Performance-based wind
engineering: towards a general procedure.” Structural Safety 33(6):367-378.
Cornell, C.A. and Krawinkler, H. (2000). “Progress and challenges in seismic
performance assessment.” PEER Center News 3:1-4.
Chun, J., Song, J., and Paulino, G.H. (2012). “Topology optimization of structures
under stochastic excitations.” Proc. of the 2012 Joint Conference of the
Engineering Mechanics Institute and the 11th ASCE Joint Specialty
Conference on Probabilistic Mechanics and Structural Reliability.
Der Kiureghian, A. (2005). “Non-ergodicity and PEER’s framework formula.”
Earthquake Engng. Struct. Dyn. 34:1643-1652.
Diniz, S.M.C., Sadek, F., and Simiu, E. (2004). “Wind speed estimation uncertainties:
effects of climatological and micrometeorological parameters.” Prob. Eng.
Mech., 19:361-371.
Ellingwood, B.R. and Tekie, P.B. (1999). “Wind load statistics for probability-based
structural design.” J. Struct. Eng. 125(4):453-463.
Jalayer, F., Franchin, P., and Pinto, P.E. (2007). “A scalar damage measure for
seismic reliability analysis of RC frames.” Earthquake Engng. Struct. Dyn.
36:2059-2079.
Kareem, A., Spence, S.M.J., Bernardini, E., Bobby, S., and Wei, D. (2013). “Using
computational fluid dynamics to optimize tall building design.” CTBUH
Journal, 3:38-43.
Kareem, A., Spence, S.M.J., Bobby, S., and Bernardini, E. (2014). “Optimizing the
form of tall buildings to urban environments.” Proc. of CTBUH 2014
Shanghai Conference.
Kharmanda, G., Olhoff, N., Mohamed, A., and Lemaire, M. (2004). “Reliability-
based topology optimization.” Struct. Multidisc. Optim. 26:295-307.
Liang, Q.Q., Xie, Y.M., and Steven, G.P. (2000). “Optimal topology design of
bracing systems for multistory steel frames.” J. Struct. Eng. 126(7):823-829.
Maute, K. and Frangopol, D.M. (2003). “Reliability-based design of MEMS
mechanisms by topology optimization.” Computers & Structures, 81:813-824.
Mijar, A.R., Swan, C.C., Arora, J.S., and Kosaka, I. (1998). “Continuum topology
optimization for concept design of frame bracing systems.” J. Struct. Eng.
124:541-550.
Minciarelli, F., Gioffrè, M. Grigoriu, M., and Simiu, E. (2001). “Estimates of extreme
wind effects and wind load factors: influence of knowledge uncertainties.”
Prob. Eng. Mech. 16:331-340.
Neves, M.M., Rodrigues, H., and Guedes, J.M. (1995). “Generalized topology design
of structures with a buckling load criterion.” Structural Optimization 10:71-
78.
Nguyen, T.H., Song, J., and Paulino, G.H. (2011). “Single-loop system reliability-
based topology optimization considering statistical dependence between limit-
states.” Struct. Multidisc. Optim., 44:593-611.
Pinto, P.E., Giannini, R., and Franchin, P. (2004). Seismic Reliability Analysis of
Structures. Pavia, Italy: IUSS Press.
Porter, K.A. (2003). “An overview of PEER’s performance-based earthquake
engineering methodology.” Proc. of the 9th International Conference on
Applications of Statistics and Probability in Civil Engineering.
Silva, M., Tortorelli, D.A., Norato, J.A, Ha, C., and Bae, H.R. (2010). “Component
and system reliability-based topology optimization using a single-loop
method.” Struct. Multidisc. Optim. 44:593-611.
Spence, S.M.J., Bernardini, E., and Gioffrè, M. (2011). “Influence of the wind load
correlation on the estimation of the generalized forces for 3D coupled tall
buildings.” J. Wind Eng. Ind. Aerodyn. 99:757-766.
Spence, S.M.J. and Kareem A. (2014). “Performance-based design and optimization
of uncertain wind-excited dynamic building systems.” Eng. Struct. 78:133-
144.
Stromberg, L.L., Beghini, A., Baker, W.F., and Paulino, G.H. (2011). “Application of
layout and topology optimization using pattern gradation for the conceptual
design of buildings.” Struct. Multidisc. Optim. 43(2):165-180.
Stromberg, L.L., Beghini, A., Baker, W.F., and Paulino, G.H. (2012). “Topology
optimization for braced frames: combining continuum and beam/column
elements.” Eng. Struct. 37:106-124.
Svanberg, K. (1987). “The method of moving asymptotes – a new method for
structural optimization.” Int. J. Numer. Meth. Engng. 24:359-373.
Vetter, C. and Taflanidis, A.A. (2012). “Global sensitivity analysis for stochastic
ground motion modeling in seismic-risk assessment.” Soil Dynamics and
Earthquake Engineering 38:128-143.
Vetter, C. and Taflanidis, A.A. (2014). “Comparison of alternative stochastic ground
motion models for seismic risk characterization.” Soil Dynamics and
Earthquake Engineering 58:48-65.

View publication stats

You might also like