You are on page 1of 10

Engineering Structures 32 (2010) 12521261

Contents lists available at ScienceDirect

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

Integrated wind load analysis and stiffness optimization of tall buildings with
3D modes
C.M. Chan a, , M.F. Huang b , K.C.S. Kwok c
a

Department of Civil and Environmental Engineering, The Hong Kong University of Science and Technology, Hong Kong

Institute of Structural Engineering, Zhejiang University, Hangzhou 310058, PR China

School of Engineering, University of Western Sydney, NSW, Australia

article

info

Article history:
Received 17 September 2008
Received in revised form
24 December 2009
Accepted 4 January 2010
Available online 20 January 2010
Keywords:
Tall buildings
Wind loads
Lateraltorsional motions
Three-dimensional modes
Stiffness optimization

abstract
Recent trends towards constructing taller buildings with irregular geometric shapes imply that these
structures are potentially more responsive to wind excitation. The wind-induced motion of modern tall
buildings is generally found to involve with significant coupled lateral and torsional effects, which are
attributed to the asymmetric three-dimensional (3D) mode shapes of these buildings. The 3D coupled
modes also complicate the use of high frequency force balance (HFFB) technique in wind tunnel testing
for predicting the wind-induced loads and effects on tall buildings. This paper firstly presents the analysis
of equivalent static wind loads (ESWLs) on tall buildings with 3D modes provided that the wind tunnel
derived aerodynamic wind load spectra are given. Then an integrated wind load updating analysis and
optimal stiffness design technique is developed for lateral drift design of tall asymmetric buildings
involving coupled lateraltorsional motions. The results of a practical 40-storey building example with
significant swaying and torsional effects are presented. Not only is the technique able to produce the most
cost efficient element stiffness distribution of the structure satisfying multiple serviceability wind drift
design criteria, but a potential benefit of reducing the wind-induced loads can also be achieved by the
stiffness design optimization method.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Recent trends towards constructing taller buildings with irregular geometric shapes imply that these structures are potentially
more responsive to wind excitation. The wind-induced motion of
modern tall buildings is generally found to involve with significant
coupled lateral and torsional effects, which are attributed to the
asymmetric three-dimensional (3D) mode shapes of these buildings. Torsional twisting effects, resulting from an imbalanced distribution of wind loads on a building surface, are further amplified
in tall asymmetric buildings by the presence of significant eccentricities between the center of stiffness of the structural system and
the center of wind forces on the building. Making accurate predictions of wind loads and their effects on such complex asymmetric buildings involving lateral and torsional motion is therefore an
important step in the design process of such structures. For design purposes, marked improvements have been made in many
wind codes and standards to provide reasonable estimation of
wind-induced structural loads on isolated buildings. However, current wind codes, which are developed for general design purpose,

Corresponding author. Tel.: +852 23587173; fax: +852 23581534.


E-mail address: cecmchan@ust.hk (C.M. Chan).

0141-0296/$ see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2010.01.001

still cannot account for wind-induced effects on specific building


shapes significantly different from simple geometric shapes, and
for the interference effects caused by the surrounding objects and
structures in a complex terrain.
In the past few decades, wind tunnels have emerged as an indispensable experimental tool for assessing wind-induced loads and
effects on complex tall buildings. Due to its accuracy and relative
ease in testing, the high frequency force balance (HFFB) technique
has been one of the most widely used methods for determining
aerodynamic wind forces on buildings. Despite its popularity, the
HFFB technique is primarily suitable for buildings with uncoupled
modes. The limitations inherent in the HFFB technique and its applicability to buildings with 3D coupled modes have been firstly
investigated by Yip and Flay and then by other researchers [15].
Indeed, modern buildings with complex geometric shapes and
3D mode shapes often complicate the use of the HFFB technique
which was traditionally developed for uncoupled buildings with
one-dimensional (1D) modes [2,3]. To overcome these limitations, much effort has been made on refining the force balance
data analysis technique to take into account coupling effects on tall
asymmetric buildings with 3D modes [4,5]. Furthermore, Chen and
Kareem [2] presented a framework for the coupled building response analysis and the modeling of the associated equivalent
static wind loads (ESWLs) using HFFB measurements.

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

In a newer wind tunnel testing method using the synchronous


multi-pressure sensing system (SMPSS), the aerodynamic wind
forces can be derived from simultaneous multiple point pressure
measurements on the surface of a building test model. This
technique allows direct computation of mode-generalized forces
for any number of modes of vibration of the building with linear
or nonlinear 3D modes. Using the SMPSS or HFFB technique,
the aerodynamic wind loads can be estimated experimentally
on a rigid scaled model of the prototype. Once the aerodynamic
wind load spectra derived from wind tunnel testing are obtained,
the wind-induced response as well as the ESWLs can then be
computed, given the dynamic properties of the prototype building.
As long as the geometric configuration of the building remains
constant, the same set of aerodynamic load spectra can be used to
estimate the wind-induced responses and ESWLs during the design
optimization process of the structural system of a tall building
without making further wind tunnel tests.
While wind tunnel techniques can provide accurate predictions
of structural loads for a building, current practice normally regards
wind tunnel derived ESWLs as being similar to the code specified
wind loads that are kept constant throughout the structural
design synthesis process. It has not been commonly recognized
by structural engineers that the ESWLs are indeed directly related
to the dynamic properties (such as the stiffness, mass, damping)
of a building. Todays design practice does not normally take
account of the frequency dependent characteristics of ESWLs in
the structural design process. Chan et al. [6] highlighted the
importance of updating the ESWLs on a building during the lateral
stiffness design optimization process of the building. Evidence has
indicated that increasing the natural frequency of wind sensitive
tall buildings by stiffness optimization can generally result in the
benefits of reducing the equivalent static wind loads [6]. However,
the integrated analysis and stiffness design optimization technique
developed by Chan et al. [6] is only applicable to symmetric tall
steel buildings with uncoupled mode shapes. It is necessary that
the integrated technique be further extended to optimize the
lateraltorsional stiffness of general tall buildings with complex
3D modes while accurately accounting for the potential benefit
of reducing the wind-induced forces and moments during the
stiffness design optimization process.
In this paper, an equivalent static wind load analysis method
based on the HFFB or SMPSS technique using measured aerodynamic force data is firstly presented. While the importance of
instantaneously updating the wind-induced load during the structural design synthesis is emphasized, attention is also paid to
the proper evaluation of the dynamic wind loads on asymmetric tall buildings with due consideration of the lateraltorsional
mechanical coupling effects as well as the intermodal coupling of
modal responses. The developed optimization technique applicable to general tall buildings is based on the Optimality Criteria (OC)
approach, which has been widely used and shown particular effective for element sizing design optimization of large building structures [7,8]. The instantaneous change in the ESWLs can be updated
whenever there exist considerable structural modifications during
the structural design process causing a substantial change in the
modal frequencies of the building. Finally, a practical 40-storey residential building with 3D mode shapes involving significant swaying and torsional effects is presented to illustrate the applicability
and effectiveness of the computer-based technique for integrated
wind-induced load analysis and stiffness design optimization of
tall building structures.
2. Determination of equivalent static wind loads
Using the aerodynamic loading measured in the wind tunnel,
the response of a building with 3D coupled mode shapes can be

1253

calculated by modal analysis for any combination of the buildings


mass, stiffness, damping and the oncoming wind speed. As the
aerodynamic wind force is a random process, the jth modal displacement response of the building in the generalized coordinate
system can be computed based on random vibration theory in the
frequency domain [2] as

q2j =

|H (f )|2 SQjj (f )df

(1)

where qj = the standard deviation value of the jth modal displacement. SQjj (f ) is the input mode-generalized wind load spectrum; H
is the mechanical admittance function which can be expressed as



Hj (f ) 2 =

1
2 fj

m2j

4 h

1 f /fj

2 i2

(2)

+ 4j2 f /fj

2

where fj is the jth modal frequency of the building; j is the jth


modal damping ratio; and mj is the jth modal mass.
In the HFFB technique, the jth mode-generalized wind load
spectrum can be given in terms of measured base moment spectra
as
SQjj (f ) =

js jl SMsl (f )

(3)

s=x,y, l=x,y,

where SMsl (f ) denotes the auto or cross power spectral density


between the respective base moments Ms (t ) and Ml (t ); s (s =
x, y, ) represents the mode shape correction that depends on
the manner of mode shapes as well as the local wind pressure
distribution and its correlation with height over a building
surface [2]. In the event that the aerodynamic wind loads are to be
measured by the SMPSS technique, the jth modal force can then
be more accurately evaluated without the need for mode shape
corrections [3].
In general, the mean square modal response in Eq. (1) can be
approximated as the sum of the background component and the
resonant component as follows

q2j = q2jb + q2jr

(4)

in which qjb = the standard deviation value of the background


component of the jth modal displacement; qjr = the standard
deviation value of the resonant component of the jth modal
displacement, which can be simplified into the following algebraic
equation using the white noise assumption as

2
qjr

SQjj (fj )



Hj (f ) 2 df =

1
m2j

2 fj

4

4j

fj SQjj (fj ).

(5)

Similarly, the mean square base moment or torque response


can be written as the sum of the background and the resonant
component as follows [10]

M2 s = M2 bs + M2 rs

(6)

in which the first part of the right-hand side is referred to as the


background component, which can be considered as quasi-static
and can be directly quantified by integrating the area underneath
the corresponding base moment auto spectrum curve of SMss (f ).
The second part of the right-hand side of Eq. (6) represents the
resonant dynamic amplification. For a lightly damped system, such
as a tall building, the resonant component of the mean square
base moment response can be obtained approximately using the
complete quadratic combination (CQC) method as follows

M2 rs =

n X
n
X
j=1 k=1

Mjrs Mkrs rjk =

n X
n
X

js ks qjr qkr rjk

(7)

j=1 k=1

where Mjrs = the standard deviation value of the jth modal

1254

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

component of the resonant base bending moment or torque;


rjk = the intermodal correlation coefficient for the jth and kth
modal response, which can be referred to the recent work of Chen
and Kareem [2], Xie et al. [11] and Huang et al. [12]; js , ks = the
modal participation coefficients of the sth component of the base
moments or base torque, which can be defined as

the two translational and one torsional directions of a building in


association with the default global coordinate system. In theory,
the mean component of both the crosswind and torsional wind
loads should be equal to zero. The alongwind mean force per unit
height can be related to the approaching wind velocity profile and
written as follows

Z
2 H

m (z ) js (z ) zdz , s = x, y
2 fj
Z0 H
js =
2

2 fj
Im (z ) j (z )dz , s =

F (z ) =
(8)

where m(z ) is the mass per unit height; Im (z ) is the rotational


mass moment of inertia about the vertical axis per unit height;
js (z ) (s = x, y, ) = the component of the jth 3D mode shape.
By substituting Eq. (5) into Eq. (7), the standard deviation resonant
component Mrs , of the base moment response Ms can be explicitly
expressed in terms of generalized force spectra as

2
Mrs

n X
n
X
j=1 k=1

js ks rjk
64 3 mj mk fj fk

fj fk

j k

2

SQjj (fj )SQkk (fk ).

(9)

s = M
s +
M

2
bs
rs2
M
+M

(10)

bs = gb Mbs = gb
M

SMss (f )df

(11)

rs = gr Mrs
M
n X
n
X
j=1 k=1

js ks rjk
64

3M M
j k

fj fk

2

!1/2

fj fk

j k

SQjj (fj )SQkk (fk )

(12)

in which the background peak factor, gb , can be approximated by


the gust factor of the oncoming wind velocity, the value of which
is usually taken to be about 3 to 4 [14]. For a Gaussian process, the
resonant peak factor, gr , can be given as [15]
0.577
2 ln v T +
2 ln v T

 z 2
H

BCD

(15)

where = the air density; U H = the wind speed at the top of the
building; = the power law exponent of wind profile; B = the
width of the building; CD = the drag force coefficient of the building. Due to the quasi-static nature of the background component
of the wind loads, the distribution of the background component
ESWLs to the floor levels over the building height can sometimes
be assumed to follow the distribution of the multiplication of the
turbulence intensity profile and the mean alongwind loading profile given in Eq. (15) as
Fbx,by (z ) = R H
Fb (z ) = R H
0

bs and the resonant compowhere the background component M

nent Mrs can be calculated respectively as

= gr

U H2

Using the gust response factor approach, the peak base moment
or torque response can be rewritten as [13]

F (z )I (z )
F (z )I (z ) zdz

Fx (z )I (z ) z
Fx (z )I (z ) zdz

bx,by
M

b
M

(16)

(17)

where Fbs (z ) (s = x, y, ) is the distributed peak background wind


load; I (z ) denotes the turbulence intensity profile.
For a building having 3D coupled mode shapes, the distribution
of the resonant component of the wind loads follows basically the
mass-related modal inertial force distribution as shown in Eq. (18)
Fjrs (z ) = gr Mjrs s (z ) = gr js qr s (z )

(18)

where Fjrs (z ) (s = x, y, ) is the peak resonant wind load


corresponding to the jth mode per unit height; s (z ) (s = x, y, )
is the inertial distribution factor, and is given as

m (z ) jx,jy (z )

, s = x, y

R H
m (z ) jx,jy (z ) zdz
0
s (z ) =
I (z ) j (z )

R m
, s = .

H
I (z ) j (z )dz
0 m

(19)

gr =

(13)

where T is the observation time (usually 3600 s) for the given


wind conditions under consideration; v indicates the mean zerocrossing rate for the base moment component process, and could
be fairly approximated by the first modal frequency value.
Based on the calculated peak base moments or torque, the
wind-induced structural loads (or the so-called ESWLs) on the
building can then be determined by distributing the peak base
moments or torque to the floor levels over the building height
[16,17]. Similar to the base moments, the equivalent static wind
loads expressed in terms of peak load Fs at each floor level, can
also be written into a linear combination of the mean (Fs ), the
Pn
background (Wb Fb ) and the resonant ( j=1 Wjrs Fjrs ) components
as
Fs = Fs + Wb Fbs +

n
X

Wjrs Fjrs ,

(s = x, y, )

(14)

j =1

where Wb = Mb /(M2 b + M2 r )1/2 , Wjrs =

Pn

k=1


Mjrs rjk /(M2 b +

M2 r )1/2 ; Fbs is the sth component peak background wind loads and
Fjrs is the sth component of the jth modal peak resonant wind
loads. Using Eq. (14), the ESWLs can be expressed respectively in

3. Dependence of wind-induced loads on natural frequency of


the structure
For multistorey buildings, the ESWLs corresponding to the
specific incident wind angle are generally expressed in terms of
the alongwind, crosswind and torsional directions; and each directional ESWL consists of the mean, background and resonant
components. Generally speaking, the mean and background components are the major parts of wind loads in the alongwind direction, since the mean wind loads as given by Eq. (15) theoretically
only exist in the alongwind direction. For wind sensitive tall buildings, significant contributions of the resonant component to the
total wind effects on buildings can be found in the crosswind and
torsional directions. The crosswind loading effects is mainly due
to wake excitation by vortex shedding as well as turbulence and
buffeting by wind flow re-attachment on the faces of the building.
Aerodynamically, the presence of turbulent flows of various scales
and the existence of neighboring building obstacles may cause the
uneven distribution of fluctuating wind pressures. The imbalance
in aerodynamic forces and uneven distribution of fluctuating wind
pressures are the major aerodynamic sources of torsional wind
loads on buildings. Even for a symmetric building, instantaneous
fluctuating torsional aerodynamic forces may still exist.

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

Since the mean and background wind loading components are


mainly related to the approaching wind conditions and the aerodynamic shape of the building, they are very much independent
of the building natural frequency [10]. As a result, varying the
buildings natural frequency by modifying the structural stiffness
of the building normally has only a negligible effect on the mean
and background values of the wind-induced structural forces. On
the other hand, as given in Eq. (12), the resonant component of
the wind loads is directly related to the natural frequency, fj , of
the building such that any changes in the natural frequencies of
the structure will subsequently lead to a direct impact on the
wind-induced resonant effects on the building, particularly in the
crosswind and torsional directions. For slender and flexible tall
buildings, the wind-induced responses are mostly dominated by
dynamic amplification effects due to turbulence and buffeting both
in the alongwind and crosswind directions, or sometimes by vortex shedding resonant effects particularly in the crosswind directions [18]. For buildings with elongated plan forms, the resonant
component of torsional wind loading also has a significant contribution to the total wind-induced response of such buildings [19].
Torsional loading effects can be further accentuated in the presence of eccentricities between the center of stiffness and the center
of wind loads. In general, modern tall buildings are wind sensitive
and the ESWLs of these structures are more influenced by the resonant effects in the crosswind and torsional directions.
Once the wind tunnel derived aerodynamic wind load spectra
are available, they can be explicitly expressed in the form of an algebraic function of the modal frequency using regression analysis [20,21]. Such algebraic functions can then be used to directly
update the prediction of wind-induced structural loads on the
building for any instantaneous change in the dynamic properties
of the building during the design optimization synthesis. With the
aid of piece-wise regression analysis, the power spectral density
function of modal wind forces (as shown in Fig. 1) for a typical
tall building can be inversely related to the modal frequency of
the building and is explicitly expressed as an algebraic function in
terms of the modal frequency within the typical range of frequency
for serviceability check, characterizing the descending part of the
power spectra, as follows
SQjj (fj ) j fj

(20)

where j and j are regression constants and normally j > 0 and


j > 0. It is worth noting that some tall buildings with well organized aerodynamic shapes may result in flat modal force spectra
with j 0. In such instances, the modal force spectra and in turn
the dynamic responses of these buildings become less sensitive to
the modification of building stiffness and some other measures,
such as installing mechanical dampers, may become necessary to
mitigate wind-induced vibration of buildings. Substituting Eq. (20)
into Eq. (5), and subsequently into (18) gives the direct dependence
of the resonant components of ESWLs to the modal frequency as
Fjrs (z ) fj

(j +3)/2

gr js
8mj 3/2

j
s ( z ) .
j

1255

However, for dynamically sensitive tall buildings, wind-induced


resonant effects may become critical. In order to make an accurate prediction of the wind-induced structural loads on the building, it is necessary that the ESWLs be always updated whenever
there exists a significant change in the structural properties of the
building. In general, the resonant component of the ESWLs for a tall
building can be reduced by increasing the modal frequencies of the
building through structural optimization by efficiently distributing
structural materials to improve the lateral and tensional stiffness
of the building.
4. Integrated design optimization
4.1. Formulation of the drift design problem
Consider a general tall building consisting of various numbers
of steel frame elements, concrete frame elements and concrete
shear wall elements. The design variables are taken as the crosssection area of steel elements, the breadth and depth dimensions
of the rectangular concrete frame elements, and the thickness
of the concrete shear wall elements. For simplicity, note that all
element sizing design variables are represented by a collective set
of i = 1, 2, . . . , Ni generic variables Xi . The minimum material
cost design of the building structure subject to j = 1, 2, . . . , N g
multiple lateral drift design constraints can be stated as:
Minimize
W (Xi ) =

Ni
X

wi Xi

(22)

i=1

subject to

(j = 1 . . . Ng )

dj dUj
XiL

Xi

XiU

(23)

(i = 1 . . . Ni ).

Eq. (22) defines the design objective of minimizing the material


cost, in which wi represents the corresponding unit material cost
for element i. Eq. (23) defines the set of j = 1, 2, . . . , N g interstorey
drift or top deflection constraints under the equivalent static wind
load conditions where dUj represents the predefined allowable
interstorey drift or overall top deflection limit. In general, the
allowable drift ratio for buildings appears to be within the range
of 1/750 to 1/250, with 1/500 being typical [22]. Eq. (24) defines
the element sizing constraints in which superscript L denotes the
lower size bound and superscript Udenotes the upper size bound
of member i.
To facilitate numerical solution of the design optimization problem, the implicit drift constraints (Eq. (23)) must be formulated explicitly in terms of various design variables. Using the principle of
virtual work, the collective set of lateral drift constraints can be expressed explicitly as
dj (Xi ) = dj (Ais , Bi c , Dic , tiw )

(21)

For wind sensitive tall buildings where the value of the exponent j is normally greater than 0, the resonant ESWLs can be reduced by increasing modal frequency according to Eq. (21).
The resonant components of ESWLs would be combined
together with the mean components as well as background components using Eq. (14) to obtain peak load at each floor level for
design purposes. For normal low-rise buildings, wind-induced
structural loads are dominated by static mean components and
quasi-static background components. The wind-induced resonant
effects on low-rise buildings are small and negligible such that
wind loads can be considered as constant static design loads.

(24)

Ns

X
i s =1
Nw

eis j
Ais

+ e0is j +

Nc

X
ic =1

e0iw j

iw =1

tiw

(j = 1, . . . , Ng )

e1iw j
ti3w

e0ic j
Bic Dic

dUj

e1ic j
Bic Dic 3

e2ic j
Bic 3 Dic

(25)

in which Ais is the cross-section area of is = 1, 2, . . . , N s steel


frame elements, Bic , Dic are the breadth and depth dimensions of
ic = 1, 2, . . . , N c rectangular concrete frame elements, and tiw is
the thickness of iw = 1, 2, . . . , N w concrete shear wall elements,
respectively; eij and e0ij are the virtual strain energy coefficient and
its correction factor of steel members respectively; e0ij , e1ij and e2ij

1256

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

(a) 0 wind.

(b) 90 wind.

Fig. 1. Wind-induced modal force spectra for the building.

are the virtual strain energy coefficients of concrete members [7].


Once the finite element analysis is carried out for a given structural
design under the ESWLs and virtual loading conditions, the internal
element forces and moments are obtained and the elements
virtual strain energy coefficients can then be readily calculated.

variables, one can derive a set of Ng simultaneous linear equations


to solve for the Ng unknown j [7]:
Ng
X
j =1

Upon establishing the explicit formulation of the drift design


constraints, the next task is to apply a suitable numerical technique for solving the optimal design problem. A rigorously derived
Optimality Criteria (OC) method [79], which has been shown to
be computationally efficient for large-scale structures, is employed
in this paper. In this OC approach, a set of optimality criteria corresponding to multiple design constraints for the optimal design
is first derived and a recursive algorithm is then applied to indirectly solve the optimal solution by satisfying the derived optimality criteria. To seek for a numerical solution using the OC method,
the constrained optimal design problem must be transformed into
an unconstrained Lagrangian function which involves both the objective function (Eq. (22)) and the set of explicit drift constraints
(Eq. (25)) associated with corresponding Lagrangian multipliers. By
differentiating the Lagrangian function with respect to each sizing
design variable and setting the derivatives to zero, the necessary
stationary optimality conditions can be obtained and then utilized
in the following recursive relation to resize the active sizing variables Ais , Bic , Dic , tiw , which are represented by the generic variables Xi for simplicity:

(
Xi

= Xi 1

Ng
X

j=1

(i = 1, 2, . . . , N )

vj

Ni X dk
X
i X
i

i =1

dj
Xi

W
Xi


Ni 
X

dk
Xi
dUk dvk
Xi v
i=1

(k = 1, 2, . . . , Ng ).

4.2. Optimality Criteria method

v+1

! )
dj / Xi
j
+1
W / Xi
v

(26)

where j denotes the associated Lagrangian multiplier for the jth


drift constraint, v represents the current iteration number; and
is a relaxation parameter. During the resizing iterations, any member reaching its size bounds is deemed as an inactive member, and
its size is set at its corresponding size limit. The two partial derivatives with respect to each element sizing design variable involved
in Eq. (26) can be analytically determined, since the structure material cost function W (Eq. (22)) and the drift constraint dj (Eq. (25))
are explicit expressions given the element sizing design variables.
Before Eq. (26) can be used to resize Xi , the Lagrangian
multipliers j must first be determined. Considering the sensitivity
of the kth drift constraint in response to changes in the design

(27)
v

Having the current design variables Xi the corresponding vj


values are readily determined by solving Eq. (27). Having the
current values of vj , the new set of design variables Xiv+1 can, in
turn, be obtained from Eq. (26). Therefore, the recursive application
of the simultaneous equations of Eq. (27) to find the vj and the
resizing formula of Eq. (26) to find the design variables constitute
the OC algorithm. By successively applying the OC algorithm until
convergence occurs, the optimal design solution is then found.
More details of the OC algorithm can be referred to the work of
Chan [7,8].
4.3. Procedure of integrated design optimization
The proposed integration of the wind-induced dynamic load
analysis and optimal structural design procedure can be outlined
step by step as follows:
1. Given the geometric shape of a tall building, determine the
aerodynamic wind load spectra by wind tunnel tests.
2. Develop the finite element model for the building and carry out
an eigenvalue analysis to obtain the natural frequencies and
mode shapes of the building.
3. Based on the current set of dynamic properties of the building,
determine the floor-by-floor ESWLs Eq. (14) for the building
using the wind tunnel derived aerodynamic wind load spectra.
4. Apply the derived ESWLs to the building and carry out a static
structural analysis to estimate the peak drift responses of the
building.
5. Establish the explicit expression of the drift constraints Eq. (25)
and formulate explicitly the optimal drift design problem.
6. For the current set of element sizing design variable Xi , solve the
system of simultaneous linear equations Eq. (27) for the set of
Lagrange multiplier j .
7. For the current value of vj , find the new set of design variables
Xiv+1 from Eq. (26).
8. Repeat steps 6 and 7 and check the convergence of the recursive
1
process: if all Xiv+1 = Xiv and vj = v
, then proceed to step 9.
j

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

1257

9. Check the convergence of the material cost objective function:


if the cost of the structure for three consecutive reanalysisand-redesign cycles is within certain prescribed convergence
criteria, for example within 0.1% difference in the current total
material cost, then terminate the design process with the minimum material cost design for the structure; otherwise, return
to step 2, update the finite element model using the current set
of design variables and repeat the eigenvalue analysis and design optimization process.
Fig. 2 shows a schematic flow chart for the proposed integrated
wind load analysis and stiffness optimization technique for the
lateral drift design of a tall building structure. The overall
procedure can be programmed and integrated with commercially
available finite element method (FEM) software such as SAP2000.
Eigenvalue analysis and structural analysis can therefore be carried
out by the FEM software within the integrated design optimization
procedure.
5. Illustrative example: 40-storey shear wall residential
building
5.1. The 40-storey building and the wind tunnel test
An optimization study of a 40-storey residential building commissioned by the Hong Kong Housing Authority was carried out at
the Hong Kong University of Science and technology. A 3D computer model of the building and a structural layout plan are given
in Figs. 3 and 4, respectively. With an elongated width of 73m, a
narrow depth of 12 m and a total height of 122 m, the building
has a critical aspect ratio (height/depth) of over 10.4. In view of its
elongated and slender configuration, the building is anticipated to
be wind sensitive and to exhibit significant swaying and twisting
responses. The building is of reinforced concrete construction with
coupled shear walls. As shown in Figs. 3 and 4, multiple structural
shear walls are coupled by lintel beams whenever possible to provide the total lateral and torsional resistance of the building. The
effectiveness of the structural resisting system depends on many
factors, such as the configuration of the structural form, the variable thickness of the shear walls and the variable dimensions of the
lintel beams.
A wind tunnel test was carried out at the CLP Power Wind/Wave
Tunnel Facility (WWTF) of the Hong Kong University of Science
and Technology. One 1:2000 scale topographical model incorporating relevant parts of the Hong Kong territory was firstly used
to determine the representative approaching wind profiles for the
building. Wind loads on the building were then measured by the
HFFB technique using a 1:400 scale rigid model subjected to the
specific wind profiles obtained from the topographical study. Wind
tunnel measurements were taken for totally 36 incident wind angles at 10 intervals for the full 360 azimuth. Two critical incident wind directions corresponding to two perpendicular incident
wind angles were identified by the wind tunnel test. One was the
0-degree wind perpendicular to the wide face acting in the short
direction (i.e. along the Y -axis) of the building; another one was
the 90-degree wind perpendicular to the narrow face acting in the
long direction (i.e. along the X -axis). It was found that while the
global maximum overturning moment about X -axis occurs at incident wind angle of 0 , the global maximum overturning moment
about Y -axis and torsional moment about the vertical Z -axis occur
at incident wind angle of 90 during the wind tunnel study. The
power spectral density functions of modal forces for the building at
two critical incident wind angles are shown in Fig. 1. Based on the
topographical study and the Hong Kong wind code [23], the design
wind speeds with a 50-year return period were adopted as 54.0
m/s and 47.7 m/s at the gradient height of 300 m for the 0 and the

Fig. 2. Flow chart of integrated analysis and design optimization process.

90 wind, respectively. Provided the wind load spectra, the ESWLs


on the building corresponding to these two most critical incident
wind directions (0 and 90 wind) have been considered and analyzed during the design optimization process.
Table 1 presents a breakdown of the results of the maximum
base shear forces and base torque for the initial building. It is noted
that in Table 1 the maximum base shear force (Fx ) and torsional
moment (Mzz ) are found under the 90 wind, and the maximum
base shear force Fy is calculated under the 0 wind. Both maximum base shear forces (Fx and Fy ) occur in the alongwind direction corresponding to the two respective incident wind angles. As
shown in Table 1 the mean and background components together
have summed up to be slightly over 70% of the total alongwind
loads for both base shears FX and FY respectively, while the resonant component contributes the remaining 30% of each base shear.
Since the base shear forces acting on the building are dominated
by the quasi-static mean and background components, which are
weakly dependent on the dynamic property of the building, only
slightly reduction in their values can be achieved by the means of
stiffness design optimization. However, in the torsional direction
about the vertical axis of the building, the dynamic resonant loading is found to be the dominating component that contributes to

1258

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

Zone 3
(22/F-Roof)

Zone 2
(12/F-21/F)

Zone 1
(G/F-11/F)

Fig. 3. The 3D view of the residential 40-storey building.

Fig. 4. Typical floor layout plan with variable shear wall elements of the residential 40-storey building.

Table 1
Breakdown of maximum wind loads for the building before optimization.
Wind loads

Mean component

Background component

Resonant component

Sum

Base shear Fx (kN)


Base shear Fy (kN)
Base torque Mzz (kN m)

1075 (33%)
9179 (40%)
30,680 (15%)

1221 (37%)
7220 (31%)
26,660 (13%)

962 (30%)
6644 (29%)
143,990 (72%)

3258 (100%)
23,040 (100%)
201,340 (100%)

Table 2
Wind-induced base shears and base moments for the building before and after optimization.
Wind loads

Hong Kong wind code

Wind tunnel results before optimization

Wind tunnel results after optimization

Wind load reduction (%)

Base shear Fx (kN)


Base shear Fy (kN)
Base moment Mxx (kN m)
Base moment Myy (kN m)
Base torque Mzz (kN m)

6,570
24,170
1,530,500
441,070
0

3,258
23,040
1,723,700
241,130
201,340

3,223
22,690
1,702,800
239,390
183,740

1.1
1.5
1.2
0.7
8.7

72% of the total base torque. As the resonant component of wind


loads is inversely proportional to the modal frequency of the building according to Eq. (21), any considerable modification in the dynamic stiffness of the building may lead to a significant change in
the torsional wind loading on the building.

In the preliminary design phase of the building before the wind


tunnel test, wind loads on the building were calculated according
to the Hong Kong Wind Code [23]. The alongwind loads in terms
of base shear and base moment derived from the Hong Kong Wind
Code are given in Table 2. The code-specified base shear Fy is found

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

1259

Fig. 5. 3D mode shapes of the 40-storey residential building.

to be slightly larger than the value predicted by wind tunnel test,


while the base shear Fx is substantially overestimated by the wind
code. Unlike the Hong Kong Wind Code in which there is no specific
guideline for predicting the wind-induced torsional load, the wind
tunnel study gives a significant base torque of 201,340 kN m for the
initial building, indicating an eccentricity of 8.7 m equal to 11.8%
of the width of the building.
The initial member sizes were established on the basis of a
preliminary strength design. Once the finite element model having
2479 frame elements and 8523 shell panels was set up for the
building, an eigenvalue analysis was then carried out to determine
the dynamic properties of the building (i.e., the natural frequencies
and mode shapes). The 3D mode shapes of the building are given
in Fig. 5. The three fundamental coupled vibration modes have the
first natural frequencies of 0.307 Hz (mainly torsional vibration),
the second of 0.323 Hz (swaying primarily in the short direction of
the building) and the third of 0.464 Hz (swaying primarily in the
long direction). The first vibration mode of the building is indeed a
torsional mode, indicating significant dynamic torsional effects on
the building. In the dynamic analysis, the values of the damping
ratio for first three vibration modes were taken as 1.5%, 1.5%
and 2%.
In this stiffness optimization study, the major design variables
are the thicknesses of variable shear walls. All variable shear wall
elements for the design optimization task are highlighted as Wall
Group 16 as shown in Fig. 3. Due to some practical planning and
constructability design considerations, the shear walls that are not
highlighted in Fig. 3 are kept unchanged during the optimization
process. For this 40-storey building, all shear walls are maintained
to have uniform thickness, except Wall Group 3 in which three
variations of wall thickness are allowed along the height of the
building. Three vertical zones of wall thickness variations are
allowed with zone one from ground floor to the 11th floor, zone
two from the 12th floor to the 21st floor, and zone three from
the 22nd floor to the main roof of the building as illustrated in
Fig. 2. Grade 45 concrete is used for the first 20 stories of the
building and Grade 35 concrete for the upper 20 stories. In the
design optimization process, the top deflection limit of H/500,
where H = 114.1 m is the height of the main roof from the ground,
was imposed at the four top corners of the building.
5.2. Results and discussion
Fig. 6 presents the design history of the total normalized cost
of the structure. Although relatively stringent restrictions due to

Fig. 6. History of the normalized structure cost for the 40-storey residential
building.

planning and constructability requirements have been imposed,


the stiffness design optimization technique has achieved a 9.9% decrease of the total initial material cost of the building. A careful
scrutiny on the optimized structure indicates that such a cost saving has been attained by more effectively stiffening the torsional
resistance of the building through deepening the lintel beams at
the two end walls (see Fig. 3) and thickening the top parapet, which
provides a ring beam wrapping along the edge of the main roof
level of the building (see Fig. 2). As a result of the torsional stiffness enhancement, the thickness of five out of the six variable shear
walls, except wall group 5, can be reduced or kept constant as
shown in Table 3. Consequently, the reduction in the wall thickness has also resulted in a net increase of 193 m2 usable floor area.
As shown in Fig. 6, a somewhat zigzag design history can be
observed at the first few design cycles. At the beginning stage of
the design history, as the structural stiffness of the building is improved by the OC optimization technique, there exists an increase
in the frequencies of the structure and a corresponding reduction
of the ESWLs, leading to a subsequent reduction in the lateral drift
response. As a result of the reduced drift response, the structure
tends to be weakened in the immediate following design cycle,
thus causing a drop in the frequencies and subsequently an increase of the ESWLs, which, in turn, leads to an increased structural cost of the building. The end result gives a fluctuating zigzag
design process, which converges to the final optimum design after
12 successive design cycles, producing a net 9.9% cost saving.

1260

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261

Table 3
Original and optimized thickness of variable shear walls.
Wall group

Original thickness (mm)


Optimized thickness (mm)

400
400

525
300

525
350/250/200a

300
200

300
425

525
300

The three values of optimized thickness correspond to Zones 1 to 3, respectively.

(a) Before optimization.

(b) After optimization.

Fig. 7. Lateral deflection profiles at the center and the corner of the building before and after optimization.

The results of the wind-induced base forces and moments for


the optimized structure are compared to that of the initial structure
as shown in Table 2. It is evident that only some minor reductions
are found in the base shear forces (Fx and Fy ) and base overturning
moments (Mxx and Myy ), since their values are more dominated
by the mean and background components of the wind loads.
However, a more significant reduction in the base torque (Mzz ) has
been achieved for the optimized structure because of the fact that
the value of the base torque is more dominated by the dynamic
resonant component of the wind load. An increase in the torsional
stiffness, by structural optimization, leading to an increase in the
first modal frequency (torsional) of the building from 0.307 Hz to
0.318 Hz, has consequently resulted in the benefit of 8.7% reduction
in the base torque.
With a significant wind-induced torsion applied on the residential building, the building is found to sway and twist with substantially larger deflection at the most distant corner of the building. As
shown in Fig. 7(a), the maximum top deflection of the initial structure at the most critical corner position is found to be 0.24 m, giving
a slight 5% violation in the top drift limit. It is worth noting that a
much smaller deflection of 0.10 m is found at the center of the top
floor of the structure. An increase of 140% in the lateral deflection
at the top corner of the initial building is due to torsional twisting
induced by wind. After optimization, the maximum top drift ratio
of the optimized structure is kept within the allowable drift ratio
of 1/500 as shown in Fig. 7(b). It is evident that the stiffness optimization technique is capable of achieving a more cost efficient
design by redistributing the structural material to maximum the
lateral and torsional stiffness of the building while satisfying the
specified drift constraints.
6. Conclusions
This paper presents an integration of an aerodynamic wind load
analysis and a lateraltorsional stiffness optimization technique
for wind-induced drift design of tall buildings with 3D modes.

Encouraging results have been found in the serviceability drift design optimization of a 40-storey practical residential building. The
computer-based integrated design optimization method is able
to produce the most cost efficient structural stiffness distribution
of the building satisfying multiple lateral drift design constraints
incorporating with torsional effects under multiple wind loading
conditions. The integrated design optimization technique is also
capable of achieving an additional benefit of wind load reduction
by instantaneously updating wind-induced structural loads during
the design synthesis process. The results of the 40-storey building indicate that the wind-induced torsional loads on the building
with the asymmetric elongated plan form have been substantially
reduced by the stiffness optimization method.
Acknowledgements
The work described in this paper was partially supported by the
Research Grants Council of the Hong Kong Special Administrative
Region, China (Project Nos. CA04/05.EG01 and 611006) and
the National Natural Science Foundation of China (Project No.
90815023), and was based upon research conducted by M.F. Huang
under the supervision of C.M. Chan for the degree of Doctor
of Philosophy in Civil Engineering at the Hong Kong University
of Science and Technology. Special thanks are due to the Hong
Kong Housing Authority for taking the lead to support the use of
the structural optimization technique and providing consent to
present the results of the example building.
References
[1] Yip DYN, Flay RGJ. A new force balance data analysis method for wind response
predictions of tall buildings. J Wind Eng Ind Aerodyn 1995;54/55:45771.
[2] Chen X, Kareem A. Coupled dynamic analysis and equivalent static wind loads
on buildings with three-dimensional modes. J Struct Eng ASCE 2005;131(7):
107182.
[3] Huang G, Chen X. Wind load effects and equivalent static wind loads of tall
buildings based on synchronous pressure measurements. Eng Struct 2007;29:
264153.

C.M. Chan et al. / Engineering Structures 32 (2010) 12521261


[4] Holmes JD, Rofail A, Aurelius L. High-frequency base balance methodologies for tall buildings with torsional and coupled resonant modes. In: Proceedings of the 11th international conference on wind engineering. 2003,
p. 238188.
[5] Flay GJ, Buttgereit VO, Bailey KI, Obasaju E, Brendling WJ. A comparison of force
and pressure measurements on a tall building with coupled mode shapes. In:
Proceedings of the 11th international conference on wind engineering. 2003,
p. 237380.
[6] Chan CM, Chui JKL, Huang MF. Integrated aerodynamic load determination
and stiffness optimization of tall buildings. Struct Des Tall Spec Build 2009;
18:5980.
[7] Chan CM. Optimal stiffness design to limit static and dynamic wind responses
of tall steel buildings. Eng J AISC 1998;35(3):94105.
[8] Chan CM. Optimal lateral stiffness design of tall buildings of mixed steel and
concrete construction. J Struct Des Tall Build 2001;10(3):15577.
[9] Grierson DE, Chan CM. An optimality criteria method for tall steel buildings.
Advances in Engineering Software 1993;16(2):11925.
[10] Davenport AG. How can we simplify and generalize wind loading? J Wind Eng
Ind Aerodyn 1995;54/55:65769.
[11] Xie J, Kumar S, Gamble S. Wind loading study for tall buildings with similar
dynamic properties in orthogonal directions. In: Proceedings of the 11th
International Conference on Wind Engineering. 2003, p. 239096.
[12] Huang MF, Chan CM, Kwok KCS, Hitchcock PA. Cross correlation of modal
responses of tall buildings in wind-induced lateraltorsional motion. J Eng
Mech ASCE 2009;135(8):80212.

1261

[13] Kareem A, Zhou Y. Gust loading factor-past, present and future. J Wind Eng Ind
Aerodyn 2003;91:130128.
[14] Zhou Y, Gu M, Xiang HF. Along-wind static equivalent wind loads and response
of tall buildings. I: Unfavorable distributions of static equivalent wind loads.
J Wind Eng Ind Aerodyn 1999;79(12):13550.
[15] Davenport AG. Note on the distribution of the largest value of a random
function with application to gust loading. Proc Inst Civ Eng 1964;28:
18796.
[16] Holmes JD. Effective static load distributions in wind engineering. J Wind Eng
Ind Aerodyn 2002;90:91109.
[17] Chen X, Kareem A. Equivalent static wind loads on tall buildings: New model.
J Struct Eng ASCE 2004;130(10):142535.
[18] Kareem A. Lateral-torsional motion of tall buildings to wind loads. J Struct Eng
ASCE 1985;111(11):247996.
[19] Boggs DW, Hosoya N, Cochran L. Sources of torsional wind loading on
tall buildings: Lessons from the wind tunnel. In: Proceedings of the 2000
Structures Congress & Exposition. 2000.
[20] Islam S, Ellingwood B, Corotis RB. Wind-induced response of structurally
asymmetric high-rise buildings. J Struct Eng ASCE 1992;118(1):20722.
[21] Chan CM, Chui JKL. Wind-induced response and serviceability design
optimization of tall steel buildings. Eng Struct 2006;28:50313.
[22] Ad Hoc Committee on Serviceability Research. Structural serviceability: A
critical appraisal and research needs. J Struct Eng ASCE 1986;112:264664.
[23] Hong Kong Building Department (HKBD). Code of Practice on Wind Effects in
Hong Kong Buildings Department, Hong Kong; 2004.

You might also like