Você está na página 1de 11

Vol.8, No.

EARTHQUAKE ENGINEERING AND ENGINEERING VIBRATION

Earthq Eng & Eng Vib (2009) 8: 469-479

December, 2009

DOI: 10.1007/s11803-009-9126-0

Design of controlled elastic and inelastic structures


A. M. Reinhorn1, O. Lavan2 and G.P. Cimellaro3
1. Dept. of Civil, Structural & Environmental Engineering, Univ. at Buffalo- The State University of NewYork, USA
2. Faculty of Civil and Environmental Engineering, TechnionIsrael Institute of Technology, Haifa 32000, Israel
3. Dept. of Structural & Geotechnical Engineering DISTR, Politecnico di Torino, Turin 10129, Italy

Abstract: One of the founders of structural control theory and its application in civil engineering, Professor Emeritus Tsu
T. Soong, envisioned the development of the integral design of structures protected by active control devices. Most of his
disciples and colleagues continuously attempted to develop procedures to achieve such integral control. In his recent papers
published jointly with some of the authors of this paper, Professor Soong developed design procedures for the entire structure
using a design redesign procedure applied to elastic systems. Such a procedure was developed as an extension of other
work by his disciples. This paper summarizes some recent techniques that use traditional active control algorithms to derive
the most suitable (optimal, stable) control force, which could then be implemented with a combination of active, passive
and semi-active devices through a simple match or more sophisticated optimal procedures. Alternative design can address
the behavior of structures using Liapunov stability criteria. This paper shows a unified procedure which can be applied to
both elastic and inelastic structures. Although the implementation does not always preserve the optimal criteria, it is shown
that the solutions are effective and practical for design of supplemental damping, stiffness enhancement or softening, and
strengthening or weakening.

Keywords: active control; integral control; design redesign procedure; inelastic structures;

1 Introduction
In the last 30 years, the possibility of integrated
design of structural/control systems in which both the
structure and its vibration control system are optimized
simultaneously has been extensively researched.
Integrated design of optimal structural/control systems
has been acknowledged as an advanced methodology for
space structures, but not many applications can be found
in civil engineering. Numerous researchers addressed
the (i) topology; (ii) shape; and (iii) size optimization
of structures using some form of control devices (see
references provided by Cimellaro et al., 2009b which
are not repeated here).
The fundamental idea of redesign was proposed by
Smith et al. (1992) and more recently, by Gluck et al.
(1996) in a form close to the one presented in this paper.
The idea of redesign is incorporated into the integrated
design of structural/control systems as a second stage of
a two stage procedure.
(1) First stage: a desired structure is chosen based
Correspondence to: A. M. Reinhorn, Dept. of Civil, Structural &
Environmental Engineering, Univ. at Buffalo- The State Univ.
of New York, 135 Ketter Hall, Buffalo, NY 14260, USA
Tel: 716-645-2839
E-mail: reinhorn@buffalo.edu

Clifford C Furnas Eminent Professor; Senior Lecturer; Assistant


Professor
Received October 11, 2009; Accepted October 31, 2009

viscoelastic braces

on best practice using engineering experience and is


assumed to be fixed. An active often adaptable
controller is designed to obtain a desired performance
requirement, e.g., drift, absolute acceleration, base
shear, etc. of the initial structure. The dynamic response
of the initial structure in this stage is defined as the ideal
response.
(2) Second stage: the structure and the controller are
redesigned by modifying the structural system to deliver
part of the controller actions, and part is preserved to be
delivered by active components to achieve a common
goal prescribed by the performance obtained in the first
step. The structure is therefore redesigned for better
economy and controllability by modifying the structural
system itself, i.e., changing stiffness, damping, and
weights, and by reducing the amount of active control
power needed to achieve the ideal response.
In the following sections, this idea is developed
progressively from simple elastic structures to inelastic
structures.

Design of simple elastic structure with


supplemental viscoelastic braces

The first development of this idea for simple frame


structures and structural systems was suggested by Gluck
et al. (1996), for a frame structure braced by control
devices (active or passive) that control its vibration, for
which the equation of motion may be written as:

470

EARTHQUAKE ENGINEERING AND ENGINEERING VIBRATION

Mx (t ) + Cx (t ) + Kx (t ) = Ef (t ) + Hu(t )

(1)

u1 g11, x g12, x
u
2
=

un g n1, x g n 2, x
g11, x g12, x

g n1, x g n 2, x ...

where matrices M, C, K characterize mass, structural


damping and stiffness related to the deformation x(t) at
various degrees of freedom.
(1) In the first stage, the brace forces are included in
the system as a set of control forces u(t), at locations
indicated by matrix D, designed to reduce the response
due to excitation forces f(t) at locations indicated by E.
The equation of motion can be easily compacted to a
state space formulation (Soong, 1990):
z (t ) = Az (t ) + Bu(t ) + Df (t )

(2)

where z(t) = {x(t), x(t ) }T, and the parameter matrices for
the system, A, for the control location, B, and for force
operation, H, are:
0
A=
-1
M K

z (t ) = Ac z (t ) + Df (t )

with Ac = A + BG

(4)

(5)

The gain matrix G, (the matrix of controlled


coefficients) is obtained by minimizing a performance
index:
J=

tf

( z T (t ) Qz (t ) + uT (t ) Ru(t )) dt

J=

tf

{zT(t) [Q + GT R G ] z(t)} dt,

or
(6)

G = 1 / 2 R 1 B T P
with P solved from
AT P + PA 1 / 2 PBR 1 B T P + 2Q = 0

(7)

The control forces are obtained from Eq. (4) or


explicitly:

(8)

(9)

The coefficients of the passive formulations, kij, cij,


as shown, do not correspond exactly to the gain factors
in Eq. (8) which were derived from the gain coefficients
gij,x, gij , x .
(2) In the second stage, using several approximations,
the stiffness coefficients ki and the damping coefficients ci
are determined here using a least square approximation.
For simplicity of further derivations, Eqs. (8) and (9) can
be transformed using story-drift formulation obtained
from the linear transformation:
x (t) = T d (t)

which is constrained by the equilibrium equation,


Eq. (5). The matrices Q and R are weighting matrices of
factors for the optimization. The selection of matrices Q
and R enable solutions within the structural or resource
limitations as illustrated further in the numerical
example.
The gain matrix G is obtained from the minimization
of the performance index J:

g1n, x x1

x2

g nn, x xn

k2
u1* k1
x1
*
x
u2 k2 k1 + k2 k3
2 +
=
kn

un*
kn kn xn

c2
c1
x1
c c + c c
x
3
2 1 2
2

cn

cn cn xu

Assume for simplicity that the control forces are of


linear form:

in which the gain matrix, G includes the coefficients, Gx,


Gx , for the structural control devices.
The equation of motion reduces to:

g1n, x x1

x2 +

g nn, x xn

The control forces can be supplied by passive


diagonal braces with stiffness and damping properties
such as viscoelastic, fluid, or hysteretic braces. These
forces are dependent on their constant stiffness and
damping coefficient as follows:

1
0
0
; B = -1 ; D = -1
-1
M C
M H
M E
(3)

u(t ) = Gz (t ) = [Gx Gx ] z (t ) = Gx x (t ) + Gx x (t )

Vol.8

where
1 1 1
1 1

T =
1

1
1
1

(10)

Simultaneously, the control forces in the diagonal


braces u(t) in terms of story drifts and drift velocities
are obtained from Eq. (8) with the transformation from
Eq. (10):
(11)
u(t ) = Gd d (t ) + Gd d (t )
in which Gd = T T GxT ; and Gd = T T GxT

No.4

A. M. Reinhorn et al.: Design of controlled elastic and inelastic structures

Using the same transformation, the brace forces of


Eq. (9) can be written as:
u* (t ) = K d d (t ) + C d d (t )

with

T
K d = T T K xT = diag (ki ) ; C d = T C xT = diag (ci )
(12)

where, ki, ci are the supplemental stiffness and damping


properties from each brace at level i.
To determine the individual components of matrices
Kd and Cd in Eq. (12), a least squares approach is
considered. Since the stiffness Kd and damping Cd can
be assumed to be independent, the least squares will be
applied independently for Kd and Cd:
T

kk = g kj , d d j (t ) dt / d k (t ) dt
0

and

ck = g kj , d d j (t ) dt / d k (t ) dt
0

(13)

Other techniques can be used for this redesign


procedure in the second stage as shown by Gluck et
al. (1996). Numerical examples were presented in the
reference mentioned above. The performance of such a
re-designed structure with the additional braces subjected
to several earthquakes was shown to be almost the same
(within 2%) of the actively controlled solution.

Redesign of elastic structures using the


optimal approach

Although alternative methods can be used for


optimal design of structures with passive components
considering Lyapunov stability criteria (Lavan and Levy,
2009), the method below allows design of both active
and passive components.
3.1 Methodology, basic equations and expressions
Cimellaro et al. (2009b) recently expanded the
procedure to allow for adjustment of the initial basic
structure in the second design stage, i.e., the redesign
stage, altering the stiffness K, damping C and weight Mg
matrices by subtraction, not only by additions, producing
weakened and lighter structures which perform better
than the original.
In the first stage, the design follows the same
procedure using the Linear Quadratic Regulator
(LQR) as presented above, although more advanced
control algorithms, such as poles assignments, H2
and H, Soong and Manolis (1987) that can be used
to determine the control forces, u(t). It is important
to note that LQR implies optimality for a white noise
excitation, an assumption that leads to a Ricatti equation

471

and its solution. For any other motion, this is suboptimal (Yang et al., 1990). However, the controllers
designed using LQR were proven efficient in practical
applications for seismic protection (Reinhorn et al.,
1993). Moreover, the active control forces obtained for
each DOF considered in the design procedure can be
easily converted to equivalent passive devices using a
method described in Lavan et al. (2008) and Cimellaro
et al. (2009a).
In the second stage, the structure is redesigned in
order to achieve the same performance, but the control
force is resolved into an active part and a passive part
depending on the constant coefficients which can be
used as structure modifiers. During the redesign process,
mass, stiffness and damping are therefore modified in
order to achieve this goal. At the end of this step, the
building will maintain the same performance, but with
less amount of active control forces. If the control
force is separated u(t)=ua(t)+up(t), where ua(t) is an
active part and up(t) is a passive part, the equation of
the redesigned structure including the change of weight
(mass), stiffness and damping matrices, respectively, by
M, K and C, becomes:

( M + M ) x ( t ) + ( C + C ) x ( t ) + ( K + K ) x ( t )
(14)
= Hua ( t ) + Ef (t )
where similarly to Eq. (4) the active component of the
control force u(t) is:
ua ( t ) = Ga z ( t )

(15)

where Ga is the active part of the controller after


redesign. Therefore, the control law can be written in the
following form:
x ( t )
Hu ( t ) = HG

x ( t )
x ( t )
x ( t )
= HGa
[ K ; C ]
Mx ( t ) (16)
x ( t )
x ( t )
and the closed-loop system after redesign is

( M + M ) x ( t ) + ( C + C ) x ( t ) + ( K + K ) x ( t )
= HGa z ( t ) + Ef (t )
(17)
where ua(t), which is given by the Eq. (14), is
the active part of the controller in Eq. (1) and
Mx ( t ) + Cx ( t ) + Kx ( t ) is the passive part, up(t).
The objective of the redesign is to find the passive
control modifiers (M, K, C) in order to minimize
the control power needed to satisfy Eq. (14) for any
given G. Note that the closed-loop system response
before and after redesign remains unchanged; therefore,
all the designed closed-loop system properties remain
unchanged. Let Bk, Bc and Bm be the stiffness, damping

472

EARTHQUAKE ENGINEERING AND ENGINEERING VIBRATION

and mass connectivity matrices of the structural


system. The changes in the structural parameters can be
expressed in the form:
K = Bk Gk BkT
C = BcGc BcT
M = BmGm BmT
where

(18)

(19)

Substituting the solution of from Eq. (1), into


Eq. (16) yields:
HGz ( t ) = (Gactive + Gpassive ) z ( t )

Gactive = HGa

and

(20)

Gpassive = - I 0 BpGp BpT L

(21)

with
Bk
Bp = 0
0

0
Bc
0

I0 = [ I

and

Gk
Gp = 0
0

0
0 ,
Bm
I

I]

L = 1
M ( HG [ K

0
Gc
0

0
0 ,
Gm

(23)

The necessary and sufficient condition to resolve the


control law into active and passive parts as in Eq. (20),
it is given as follows: According to Smith et al. (1992)
there exists an active controller Ga:
HGa = HG + I 0 BpGp BpT L
if and only if
HH + I 0 BpGp BpT L = I 0 BpGp BpT L

(24)

(26)

An objective function representing the power of the


active part of the control law is given by:
(27)

where RXX is the covariance matrix of the response. A


constraint optimization is formulated
minimize

F ( Gp ) =trace (Ga RXXGaT R )

(28)

where Ga is given by Eq. (25), subjected to the equality


constraints of Eq. (24) and inequality constraint in Eq. (26).
An approach to numerically solve the constrained
optimization problem is to use the Exterior penalty
function method that is part of the Sequential
Unconstrained Minimization Techniques (SUMT)
(Vanderplaats, 2005). The approach consists of creating
an unconstrained objective function of the form:
(29)

where F(Gp) is the original objective function, P(Gp)


is the penalty function and rp is a multiplier which
determines the magnitude of the penalty and is held
constant during a complete unconstraint minimization.
The penalty function P(Gp) is given by the following
expression in this case:
T
P ( Gp ) = trace Z ( GP + S - C )(GP + S - C ) +

trace ( BB + I 0 BpGp BpT LHH + I 0 BpGp BpT L )

( BB

I 0 BpGp BpT LHH + + I 0 BpGp BpT L )

(30)

where Z = diag(zi) is a diagonal matrix where the


scalars zi are chosen such that zi = 1 if the corresponding
inequality constraint Gpi+sici0 is active, and zi = 0 if the
constraint is not active. So the new objective function is
given by the following expression:

(GP , rp ) = trace (Ga RXXGaT R ) +

Such that Ga is given by:


Ga = G + I 0 BpGp BpT L

Gp + S C

(Gp , rp ) = F (Gp ) + rp P (Gp )


(22)

C ])

S = diag[k0i,..c0i,m0i,] is the matrix of the initial


parameters, then these constraints can be presented as:

F ( Gp ) = uaT ( t ) Rua ( t ) dt = trace (Ga RXXGaT R )

Gk = diag ( , ki ,)
Gc = diag ( , ci ,)
Gm = diag ( , mi ,)

where

Vol.8

(25)

where ( )+ denotes the Moore-Penrose inverse of a


matrix. The stiffness and the damping of any element
of the system after redesign cannot be negative while
the weight of any element cannot decrease lower than a
specified bound, therefore imposing specific constraints.
Therefore, if C = diag[ki,..ci,mi,] is a matrix with
diagonal elements containing the specified lower bound
values of the structural elements after redesign and

T
rp trace Z ( GP + S - C )(GP + S - C ) +

rp trace ( BB + I 0 BpGp BpT LHH + I 0 BpGp BpT L )i

( BB

I 0 BpGp BpT LHH + I 0 BpGp BpT L )

(31)

Minimization of Eq. (31) requires that the following


first-order necessary condition is satisfied.
P1 vec diag ( Gp ) = r1

(32)

No.4

A. M. Reinhorn et al.: Design of controlled elastic and inelastic structures

where vec diag(Gp) denotes a vector with diagonal


elements of Gp as its components. We have

(B

P1 =

I B + T RB + I 0 Bp )i( BpT LH + HRXX H + HLT Bp ) +

T T
p 0

rp ( BpT I 0T I 0 Bp )i( BpT LH + HLT Bp ) rp ( BpT I 0T B + I 0 Bp )i

(B

T
p

LH + HLT Bp ) + Z

(33)

and
r1 = vec diag ( BpT I 0T B + T RGHRXX H + HLT Bp )
rp Zvec diag ( S - C )

(34)

Therefore, the following algorithm can be used to


find the optimal solution, where it is assumed that the
matrix P1 is invertible. The general algorithm for the
exterior penalty function approach is shown in Fig. 1.
If a small value of rp is chosen, the resulting function
(GP , rp ) is easily minimized, but may yield large
constraints violations. On the other hand, a large value
of rp will ensure near satisfaction of all constraints but
will create a very poorly conditioned optimization
problem from a numerical standpoint. Therefore, the
algorithm starts with a small value of rp and minimize
(GP , rp ) . Then rp is increased by a factor , say = 3,
and ( GP , rp ) is minimized again, each time beginning
the optimization from the previous solution, until a
satisfactory result is obtained.
3.2 Numerical example - MDOF 9-story shear-type
building
A nine-story structure considered in this example
is 45.73 m (150 ft) by 45.73 m (150 ft) in plan, and

Start
Given Gp0 rp,
Minimize (GP , rp ) as
unconstrained function Eq. (33)
Find Gp to min. (GP , rp )

| (G q ) (G q-1 ) |< (G 0 )

No

rp = rp

Yes
Exit
Fig. 1 Algorithm for the exterior penalty function method

473

37.19 m (122 ft) in elevation. The bays are 9.15 m (30


ft) on center, in both directions, with five bays each in
the North-South (N-S) and East-West (E-W) directions.
The buildings lateral load-resisting system is comprised
of steel perimeter moment-resisting frames (MRFs) with
simple framing on the furthest south E-W frame. The
interior bays of the structure contain simple framing
with composite floors. Typical floor-to-floor heights
(measured from center-of-beam to center-of-beam for
analysis purposes) are 3.96 m (13 ft). The floor-tofloor height of the basement level is 3.65 m (12 ft) and
for the first floor is 5.49 m (18 ft). The floor system is
comprised of 248 MPa (36 ksi) steel wide-flange beams
acting compositely with the floor slab, each frame
resisting one-half of the seismic mass associated with
the entire structure.
The seismic mass at the ground level is 965 t (66.0
kip-sec2/ft), 1010 t (69.0 kips-sec2/ft) for the first level,
989 t (67.7 kip-sec2/ft) for the second through eighth
levels and 1070 t (73.2 kip-sec2/ft) for the ninth level.
The seismic mass of the above ground levels of the entire
structure is 9000 t (616 kip-sec2/ft). More details about
the model can be found in Cimellaro et al. (2009b). The
lateral stiffness of the shear-type model are reported in
column 4 of Table 1, and the first three frequencies of the
shear-type model are 0.45, 1.28 and 1.99 Hz. Rayleigh
proportional damping is considered, including 2% of
damping ratio for the first two modes.
The structure was subjected to the first 30 s of white
noise with amplitude of 0.15 g and with a sampling
frequency of 0.02 s. The drift and acceleration response
during the first stage of the algorithm are shown in Table
3. Columns 2 and 3 show the drift and acceleration
response of the structure with a hypothetical active
control force which is defined here as the Ideal
Response. Initially, the story lateral stiffness is
reduced proportionally to 30% of the initial stiffness
value in order to obtain a first natural period increment
of 83%. Response of the lighter structure is shown in
columns 4 and 5 of Table 3 . Then, a fully active brace
is placed at each story level in order to achieve the same
performance in terms of drift of the uncontrolled initial
structure. Values of the maximum active control force
at each story level are shown in column 8 of Table 3 .
After the structure and controller were designed
independently in first stage, the controller and the
building are redesigned together in the second stage
to achieve the same performance (Ideal Response) by
reducing the amount of active control and changing the
passive components as shown in Table 1 and Table 2.
The initial total energy transferred to the structure from
the controller is equal to 2623.0 Nm and, after redesign,
is equal to 1972.1 Nm, so the percentage of reduction
of the total energy transferred is 24.81%. Results of
the redesign procedure are shown in columns 4, 5 and
6 of Table 3. Comparisons between the fully controlled
structure and the redesigned structure response are
shown in Fig. 2.

474

EARTHQUAKE ENGINEERING AND ENGINEERING VIBRATION

Vol.8

Table 1 Structural parameters after redesign


Original

Story
level
No.

M
kN.s2/m

C
kN.s/m

(1)

(2)

9
8

Redesign
K
10 kN/m

Mopt
kN.s2/m

Copt
kN.s/m

10 kN/m

(3)

(4)

(5)

(6)

(7)

534.1

411.4

100.02

350.3

1352.5

18.21

494.7

1152.8

291.12

342.2

5843.8

62.79

494.7

390.9

71.52

336.0

434.9

15.68

494.7

1077.4

247.63

348.3

2284.2

56.02

494.7

487.5

75.03

361.2

289.9

16.89

494.7

877.3

170.08

370.0

596.6

39.98

494.7

1119.4

224.76

423.7

984.1

54.02

494.7

1301.4

263.02

474.1

1841.0

65.92

503.5

906.5

143.48

441.3

723.7

36.21

4 Redesign of inelastic structures using


optimal approach

Kopt
3

Table 2 Percentage change in structural parameters (negative


indicates removing)
Story
level

Umax

Consider a multi-degree-of-freedom inelastic


building structure subjected to a one-dimensional
external excitation. The general equation of motion of
the inelastic system with active control forces is given by

No.

(%)

(%)

(%)

(kN)

(1)

(2)

(3)

(4)

(5)

-34.4

228.8

-81.8

134.97

-30.8

406.9

-78.4

107.11

Mx (t ) + Cx (t ) + Kx (t )+Ts fs x ( t ) = Hu(t )+Ef (t ) (35)

-32.1

11.3

-78.1

110.81

-29.6

112.0

-77.4

114.80

where x(t) is the displacement vector, M and C are the


mass and inherent damping matrices, respectively, K is
the stiffness matrix of all linear elements, u(t) is the active
control force vector; H is the location matrix for the active
control forces; E is the excitation influence matrix; Ts is
the location matrix of the restoring forces and fs[x(t)] is
a vector of nonlinear restoring forces in the structural

-27.0

-40.5

-77.5

125.26

-25.2

-32.0

-76.5

113.00

-14.4

-12.1

-76.0

113.42

-4.2

41.5

-74.9

108.17

-12.3

-20.2

-74.8

57.02

4.1 Basic equations and expressions

Table 3 Drift and acceleration response for the two stages of the algorithm
Story
level
No.

Drift (%)

(1)

(2)

(3)

(4)

(5)

(6)

(7)

(8)

0.31

3.61

0.80

3.09

0.31

2.03

173.39

0.18

2.97

0.47

2.64

0.15

1.85

159.72

0.94

2.71

2.55

2.46

0.77

1.77

0.27

2.71

0.76

2.86

0.22

0.90

2.64

2.74

2.99

0.86

0.42

3.74

1.39

2.64

0.38

3.50

1.04

0.38

2.95

0.79

2.79

Ideal response
xa (m/s2)

Totally active [LQR]

T*1/T1=1.83#
Drift (%)

Drift (%)

Redesigned structure
Umax (kN)

(9)

(10)

(11)

0.23

1.87

134.97

0.10

1.67

107.11

153.86

0.67

1.89

110.81

1.87

163.49

0.19

1.70

114.80

1.88

165.26

0.73

1.80

125.26

0.43

2.04

151.52

0.34

1.90

113.00

2.67

0.37

2.10

127.90

0.31

1.98

113.42

0.92

2.62

0.38

1.99

97.91

0.31

1.96

108.17

1.87

2.47

0.76

1.98

66.67

0.71

1.93

57.02

xa (m/s2)

Note: #The stiffness is reduced proportionally to 30% of the initial lateral stiffness

Umax(kN)

Drift (%)

xa (m/s2)

xa (m/s2)

No.4

A. M. Reinhorn et al.: Design of controlled elastic and inelastic structures

Ideal response
Active redesign

Story level

Story level

475

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Drift (%)
(a)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Acceleration (m/s2)
(b)

Fig. 2 Response comparison of actively controlled and redesigned system

elements. fs[x(t)] in Eq. (35) can be separated into


two parts, one representing the elastic behavior and the
other representing the nonlinear hysteretic behavior, i.e.,
f s x ( t ) = Kx ( t ) + Hfs x ( t )

(36)

in which K is a linear elastic stiffness matrix for the


linear elastic parts of the nonlinear or hysteretic structural
components; f s =[f1 f2 fl]T is a vector representing the
nonlinear or hysteretic parts of the restoring forces for
nonlinear structural components; and H is a location
matrix for the nonlinear elements. The component i of
the vector of nonlinear forces, which is the nonlinear
force in element i, is modeled using the continuous
evolutionary Sivaselvan-Reinhorn model (Sivaselvan
and Reinhorn, 2000) discretized as follows:
fsi xi ( t + ) , xi ( t + ) = ai ki xi ( t + ) + (1 - i ) f yi ( i + i )
(37)
i = 1, , n
where the normalized nonlinear force, i=(fsi- ikixi) /
(1-i)fyi, is expressed by

i (t ) =

xi ( t )
ni
1 i ( t ) 1i + 2i sgn ( i ( t ) xi ( t ) )

xyi

i = 1, , n

(38)

in which ki is the elastic stiffness; i is the ratio of


post-yielding to pre-yielding stiffness; ni is the power
controlling the transition from the elastic to inelastic
range; 1i and 2i are parameters controlling the shape of

the unloading curve (2i=1 1i, for compatibility with


the plasticity theory); fsi is the portion of the applied
force shared by the hysteretic spring; and fyi is the
yielding force of the hysteretic spring. The procedure
suggested herein can use any inelastic-hysteretic model
without changing the procedure.
Note that the ith DOF is linear elastic if i=1.0. The
term ikixi(t) is the post-yielding hardening linear elastic
stiffness that will appear in the K matrix of Eq. (36).
The design method presented further herein reverses the
conventional procedure by designing the structure after
the controller has been determined. The procedure is
done in two stages as before:
In the first stage, based on the structural parameters
(e.g., strength, damping and weight), a controller is
designed to satisfy the desired performance requirements
(e.g., drift, absolute acceleration, base shear, etc.) of
the structure, which is assumed to be inelastic. An
appropriate nonlinear control law that takes into account
stability issues is adopted, since the structure is intended
to yield, i.e., becoming inelastic. One of the methods
that are able to accommodate the nonlinear behavior of
the system and the uncertainties involved in the structural
analysis is the sliding mode control (see Utkin, 1992 and
Yang et al., 1995). The main advantages of sliding mode
control are the: (i) possibility of stabilizing nonlinear
systems, which cannot be stabilized by using continuous
state feedback laws; (ii) robustness against a large class
of perturbations, or model uncertainties; and (iii) need
for a reduced amount of information in comparison to
classical control techniques. In general, the design of
the sliding mode control can be summarized in two
fundamental steps as shown by Lavan et al. (2008).
The first step is the selection of sliding surfaces such

476

EARTHQUAKE ENGINEERING AND ENGINEERING VIBRATION

that the system exhibits the desired behavior in the


sliding mode. The second step is determining the
control laws that guarantee that the structure reaches the
sliding condition. Yang et al. (1995) presented details
of this procedure. For the first step, a possible law for
the continuous sliding mode control in the case of a
saturated controller is given by
i*Gi i i

ui ( t ) =
*
ui max sgn ( i Gi i )
i = 1, , n

if i*Gi i i ui max
otherwise

(39)

in which 0 i* 1, i 0 , is the sliding margin and


ui max represents the upper bound of the ith control force.
i and Gi are r-vectors obtained using the procedure
described in Yang et al. (1995).
In the second stage, the redesign is accomplished
by modifying the structural parameters subjected
to certain constraints by minimizing the amount of
active control power needed to achieve the desired
performance requirements. An objective function
expressed by the sum of the power of the active control
forces is minimized. The objective function is given by

( )

F1 = u ( t ) Rua ( t ) dt
T
a

(40)

where ua is the active force vector of the controllers after


redesign and is not known apriori, but can be determined
by assuming that the building has the same structural
response x, before and after redesign. Minimizing
Eq. (40) can lead to spikes in the time history of the
active control force ua that cannot be followed by the
controller given in Eq.(39). Therefore, another objective
function that minimizes the maximum of the active
control force was suggested:

( )

F2 = max ( uaT ( t ) Rua ( t ) )

(41)

each objective function, subjected to all constraints,


independently. F1w( ) and F2w( ) are the values of the
objective functions associated with the initial conditions
of the structural parameters .
In order to determine the objective function in Eq. (42)
it is necessary to determine the active force vector
ua. Similar to the force derived in Eq. (14), the active
component can be written as:

ua ( t ) = H 1 Hu ( t ) + Mx ( t ) + Cx ( t ) + Ts f s [ x ( t )]

(43)

If H is not invertible, ua(t) can be approximated by


performing a pseudo-inverse of H or by employing a least
square procedure. Substituting Eq. (38) into Eq. (37),
the hysteretic force can be then written as follows

f si ( t ) = Bi ( t ) + Di ( t ) i

( )
( )

( )
( )

( )

( )
( )

( )
( )

F2 F2 *
+ w
F2 F2 *

where W1 is the weighting factor; F1( ) and F2( ) are,


respectively, the objective function given in Eqs. (40)
and (41); F1*( ) and F2*( ) are the target objective
functions; F1w( ) and F2w( ) are the worst known
value of the two objective functions. F1*( ) and F2*( )
are the most difficult values to define in advance. In this
paper, these values have been determined by optimizing

(44)

Bi ( t ) = i ki xi ( t + ) + (1 i ) ki xi ( t )

1 ( t ) ni + sgn ( ( t ) x ( t ) )
i
i
1i
2i
i

Di ( t ) = (1 i ) f yi i ( t )

i = 1,, n
(45)

Assuming that the history of i is known and it is


the same for the structure before and after redesign,
substituting Eq. (44) in Eq. (43) yields the following
equation:

ua ( t ) = H 1 Hu ( t ) + Mx ( t ) + Cx ( t ) + Ts (1 ) D ( t )

(46)
Finally, the optimization problem can be solved by
substituting Eq.(46) in Eqs.(40) and (41). Define:
M
= 0
0

(42)

i = 1,, n

where i is a multiplier of the yielding restoring force


(1-i)fyi such that the redesign yielding force in the
element i is i multiplied by its original yielding force,
while

A multi-objective optimization problem is then


formulated where the combined objective function is
given by
F1 F1 *
F = W1 w
F1 F1 *

Vol.8

0
C
0

0
0
B

(47)

where
M = diag ( , mi ,)
C = diag ( , ci ,)
B = diag ( , i ,)

i = 1, , n

(48)

are diagonal matrices of the design variables, and i are


in fact weakening or strengthening factors.
Then, if L = diag[,miL,..ciL,iL,] and U = diag[,
U
mi ,..ciU,iU,] are matrices with diagonal elements
containing the specified lower and upper bound values

No.4

A. M. Reinhorn et al.: Design of controlled elastic and inelastic structures

of the structural elements, respectively, and S = diag[,


mio,..cio,io,] is the matrix of the initial parameters,
then these constraints can be presented in matrix format
as
L + S U
(49)

T
( , ) = F ( ) + trace Z L ( + S - L )( + S - L ) +

T
+ trace Z U ( S + U ) ( S + U )

(53)

If a small value of is chosen, the resulting function


( , ) is easily minimized, but may yield large
constraints violations. On the other hand, a large value
of will ensure near satisfaction of all constraints but
will create a very poorly conditioned optimization
problem from a numerical standpoint. Therefore, the
algorithm starts with a small value of and minimize
( , ) . Then is increased by a factor , say = 3,
and ( , ) is minimized again, each time beginning
the optimization from the previous solution, until a
satisfactory result is obtained. Finally, at the end of
the second stage, the building will maintain the same
performance by transforming the control forces in
equivalent changes of the mass, damping matrix and
restoring force, respectively, by M, C and fS.

Therefore, the optimization problem is defined as


follows:
Minimize: F ( ) as objective function

(50)

Subjected to: L + S U as side constraints


The solution of the optimization problem can be
determined numerically using the Exterior Penalty
Function Method (Vanderplaats, 2005) described above
and reformulated here. The approach consists of creating
an unconstrained objective function of the form

( , ) = F ( ) + P ( )

(51)

4.2 Numerical example - MDOF 8-story building

where F() is the original objective function in Eq. (42),


P() is the penalty function and is a multiplier which
determines the magnitude of penalty and is held constant
during a complete unconstraint minimization. The
penalty function P() is given by

An eight-story nonlinear structure is considered


where the parameters follow the example of Yang et al.
(1990), however, the stiffness at each story was scaled to
result in a period of 0.81 s. The mass, damping, stiffness
and yield strength of the shear-type model are reported
in columns 2, 3, 4 and 5 of Table 4. The structure was
subjected to the first 30 s of El Centro earthquake with a
sampling frequency of 0.02 s.
The drift and acceleration responses during first
stage of the algorithm are shown in Table 5. The drift
and the acceleration response of the initial building or,
in other words, the performance level to be achieved
(Target Response), are shown in columns 2 and 3.
The initial story lateral stiffness of the elastic portion
of the Sivaselvan-Reinhorn model is first reduced
proportionally to 50% of the initial stiffness value in

T
P ( ) = trace Z L ( + S - L )( + S - L ) +

T
trace Z U ( S + U ) ( S + U )

477

(52)

where ZL= diag(,zi,) is a diagonal matrix where the


scalars zi are chosen such that zi=1 if the corresponding
lower bound inequality constraint i+si-li0 is active
and zi=0 if the constraint is non active. An analogous
expression of ZU exists for the upper bound.
Therefore, the new objective function is given by

Table 4 Optimal structural parameters after redesign using El Centro earthquake


Story
level
No.

Original

Redesign

C
(kN.s/m)

Fy

Mopt

(kN.s2/m)

(103kN.m)

(kN)

(kN.s2/m)

(2)
345.6

(3)
196

(4)
301.4

(5)

(6)

(7)

(8)

(9)

4521

197.6

768.3

5091.7

123.5

345.6

243

371.8

6321

223.2

486.0

4400.9

207.3

345.6

298

455.4

8653

261.1

745.2

5035.3

223.9

345.6

349

534.6

10692

273.3

1078.3

5905.5

226.2

345.6

386

591.8

12428

289.6

471.2

6495.6

239.7

345.6

410

627.0

13794

308.5

1159.7

6986.7

227.9

345.6

467

704.0

16192

312.6

1950.8

7986.9

218.1

345.6

490

748.0

17952

319.8

2632.1

8806.3

215.2

(1)

Copt
(kN.s/m)

Fy,opt

Umax

(kN)

(kN)

478

EARTHQUAKE ENGINEERING AND ENGINEERING VIBRATION

Vol.8

Table 5 Drift and acceleration response


Story level

Target response

T*1/T1=1.41#

Active response
Umax=250 kN

Redesign approach

No.

Drift (%)

xa (m/s2)

Drift (%)

xa (m/s2)

Drift (%)

xa (m/s2)

Drift (%)

xa (m/s2)

(1)
8

(2)
0.36

(3)
12.28

(4)
0.45

(5)
6.67

(6)
0.43

(7)
7.37

(8)
0.20

(9)
9.86

(10)
123.5

0.65

11.69

1.06

6.76

0.80

7.12

0.41

8.58

207.3

0.81

9.62

0.97

5.63

0.75

4.99

0.65

8.04

223.9

0.71

9.44

0.83

5.91

0.56

5.31

0.70

5.85

226.2

0.61

8.21

0.61

6.10

0.50

5.94

0.62

6.03

239.7

0.71

6.77

0.71

5.83

0.49

5.58

0.81

6.38

227.9

0.58

6.68

0.58

4.63

0.48

4.43

0.68

3.90

218.1

1
0.58
4.35
0.58
3.56
0.48
3.87
#
Note: The stiffness is reduced proportionally to 50% of the initial lateral stiffness

0.67

3.76

215.2

order to obtain a first natural period increment of 41%.


The response of the lightweight structure is shown in
columns 4 and 5 of Table 5. An active brace with the
same maximum saturation control force (Umax = 250 kN)
was introduced at each story level in order to achieve the
same performance in terms of drift of the uncontrolled
initial structure. After the structure and controller are
designed independently in the first stage, the controller
and the building are redesigned together in the second
stage to achieve the same performance (Target
Response) by reducing the objective function in Eq. (42).
The values of the target objective functions selected are
F1*()=1.02107 and F2*()=0.80103, while the value
of the objective functions corresponding to the initial
conditions are F1w()=6.62107 and F2w()=7.72104,
while the value of the weight factor selected is
W1=1.32106. In this case, the total power of the active
control force after redesign is reduced by 24.5%.
The results of the optimal structural parameters (M,
C, Fy and Umax) after redesign are shown in columns 6,
7 8 and 9 of Table 4 while the percentages of variation
with respect to the initial structure are shown in Table
6. The performance in terms of drift and acceleration of
the redesigned structure are shown in columns 8 and 9
of Table 4. Note that the weight is reduced, especially in
the upper stories. Globally, a total weight reduction of
21% is obtained.

5 Concluding remarks
An integrated design procedure for the design of
elastic and inelastic structures equipped with control
systems, either active, or passive, or a combination, was
summarized in this paper. The method follows a two stage
approach. First, an initial structure is assumed including
its topology, stiffness and strength, and a control force is
determined using a rigorous active control algorithm to
ensure its controllability and stability. Then, the control

Umax(kN)

force is resolved in passive and active components. The


passive components are implemented by modifying
the structural system, i.e., weight, stiffness, damping
and strength, while the remaining part is implemented
with an active controller. In order to achieve an optimal
structural configuration, optimization of the passive and
active components is implemented in the second step.
In comparison with a traditional design, the proposed
integrated design procedure is able to achieve an overall
weight saving, a stiffness reduction and a minimal active
control component. Such reductions consequentially
may lead to a reduction of initial investments costs.
Note that the integral design leads to softer, weaker and
highly damped stable structures ensured by the stability
of the algorithms applied in the first stage of design.
This paper is a summary of developments prepared
for multiple journal archival papers reformulated to
obtain a unified and unique approach. More details can
be found in the References. The work is, in particular,
an implementation of the vision of the mentor of the
authors, Professor Emeritus Tsu T. Soong.

Acknowledgement
The authors wish to acknowledge the guidance and
direct contributions of Professor Soong to the integral
design approach and to the inspiration and vision he
always provided.
The authors also acknowledge
the financial support of the Multidisciplinary Center
for Earthquake Engineering Research (MCEER)
headquartered at the University at Buffalo, State
University of New York, which excelled in the
development of seismic protective systems for structures
and equipment.

References
Cimellaro GP, Lavan O and Reinhorn AM (2009a),

No.4

A. M. Reinhorn et al.: Design of controlled elastic and inelastic structures

Design of Passive Systems for Controlled Inelastic


Structures, Earthquake Engineering and Structural
Dyamics., 38(6): 783804.
Cimellaro GP, Soong TT and Reinhorn AM (2009b),
Integrated Design of Controlled Structural Systems,
ASCE/Journal of Structural Engineering, 135(7):
853862.
Gluck N, Reinhorn A and Levy R (1996), Design
of Supplemental Dampers for Control of Structures,
ASCE/Journal of Structural Engineering, 122(12):
13941399.
Lavan O, Cimellaro GP and Reinhorn AM (2008),
A Noniterative Optimization Procedure for Seismic
Weakening and Damping of Inelastic Structures,
ASCE/Journal of Structural Engineering, 134(10):
16381648.
Lavan O and Levy R (2009), Simple Iterative Use of
Lyapunovs Solution for the Linear Optimal Seismic
Design of Passive Devices in Framed Buildings,
Journal of Earthquake Engineering, 13(5): 650666.
Reinhorn AM, Soong TT, Riley MA, Lin RC, Aizawa
S and Higashino M (1993), Full Scale Implementation
of Active Control Part II: Installation and Performance,
ASCE/Journal of Structural Engineering, 119(6):
19351960.
Sivaselvan M and Reinhorn A (2000), Hysteretic

479

Models for Deteriorating Inelastic Structures, ASCE/


Journal of Engineering Mechanics, 126(6): 633640.
Smith MJ, Grigoriadis KM and Skelton R (1992),
Optimal Mix of Passive and Active Control in
Structures, Journal of Guidance, Control and
Dynamics, 15(4): 912919.
Soong TT (1990), Active Structural Control: Theory and
Practice, Longman Scientific, London.
Soong TT and Manolis GD (1987), Active Structures,
ASCE/Journal of Structural Engineering, 113(11):
22902301.
UtkinVI (1992), Sliding Modes in Control
Optimization, Springer: New York.
Vanderplaats GN (2005), Numerical Optimization
Techniques for Engineering Design, Vanderplaats
Research & Developments, Inc., Colorado Spring, CO
80906.
Yang JN, Akbarpour A and Askar G (1990), Effect of
Time Delay on Control of Seismic-Excited Buildings,
ASCE/Journal of Structural Engineering, 116(10):
28012814.
Yang JN, Wu JC and Agrawal AK (1995), Sliding
Mode Control of Nonlinear and Hysteretic Structures,
ASCE/Journal of Engineering Mechanics, 121(12):
13861390.

Você também pode gostar