vibration control of civil engineering structures via

28
Vibration control of Civil Engineering structures via Linear Programming P. Rentzos, G.D. Halikias and K.S. Virdi Control Engineering Research Centre School of Engineering and Mathematical Sciences City University, London EC1V 0HB, UK Email: [email protected], [email protected], [email protected] Abstract The paper presents a novel active-control design approach which minimizes the peak response of regulated signals rather than, e.g., r.m.s or energy levels optimized by traditional control techniques. This objective is more relevant for active control of civil- engineering structures, as failure occurs after a maximum displacement is exceeded in a structural member, while control constraints typically arise from hard saturation limits on the actuator signal and its rate. The design method is formulated in discrete-time and involves the parametrization of all finite settling-time stabilizing controllers. This leads to a linear programming optimization framework, in which the peak response of the structure is directly minimized, subject to linear constraints on the actuator’s peak level signal and its rate. The design method is illustrated via a simulation study based on a simple model corresponding to a benchmark design problem. The simulation results compare favourably to those obtained via LQG active control. Finally, some practical implementation issues related to the method are discussed. 1

Upload: others

Post on 24-Oct-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Vibration control of Civil Engineering structures via

Vibration control of Civil Engineering structures

via Linear Programming

P. Rentzos, G.D. Halikias and K.S. Virdi

Control Engineering Research Centre

School of Engineering and Mathematical Sciences

City University, London EC1V 0HB, UK

Email: [email protected], [email protected], [email protected]

Abstract

The paper presents a novel active-control design approach which minimizes the peak

response of regulated signals rather than, e.g., r.m.s or energy levels optimized by

traditional control techniques. This objective is more relevant for active control of civil-

engineering structures, as failure occurs after a maximum displacement is exceeded in a

structural member, while control constraints typically arise from hard saturation limits

on the actuator signal and its rate. The design method is formulated in discrete-time

and involves the parametrization of all finite settling-time stabilizing controllers. This

leads to a linear programming optimization framework, in which the peak response of

the structure is directly minimized, subject to linear constraints on the actuator’s peak

level signal and its rate. The design method is illustrated via a simulation study based

on a simple model corresponding to a benchmark design problem. The simulation

results compare favourably to those obtained via LQG active control. Finally, some

practical implementation issues related to the method are discussed.

1

Page 2: Vibration control of Civil Engineering structures via

1. Introduction

Structural control aims at protecting structures from severe natural hazards such as

earthquakes and large wind loads. The simplest structural control scheme involves

a Tuned-Mass-Damper (TMD), which consists of a pendulum vibrating at the same

frequency as the natural frequency of the structure, opposing its movement to mitigate

the response. Control schemes such as passive, active, semi-active or hybrid have been

proposed and implemented with various degrees of success [15].

Over the last few years a wide range of design methodologies have been proposed in the

area of structural control, including non-linear/sliding-mode control, pole-placement

and observer-based methods, adaptive control, fuzzy/neural-based methods, reliability-

based control and optimal control [9]. Optimal control appears to be the design method

increasingly favoured by most researches, mainly due to important recent theoretical

advances in this field and to the design flexibility that this method offers. The two

most important optimal control paradigms, around which most other optimal-control

methods cluster, are LQR/LQG optimal control and H∞ optimization methods. In

addition, new design optimal algorithms have been proposed to account for different

objectives, assessed via analytical, simulation and experimental results.

Optimal control design methods are typically formulated as optimization problems

involving the minimization of a norm, such as the H2 or H∞ norm, of the closed-loop

transfer function between an input disturbance signal and the regulated outputs. For

example, the H2 norm measures the expected power of the regulated signal (mean-

square value). Normally, the input disturbance signal in this case is assumed to be a

random white-noise process. Weighting factors or filters can be employed to emphasize

specific frequency ranges of the input or output spectrum.

Consider the diagram shown in Figure 1, where n(t) represents a zero-mean white-noise

vector signal, i.e.

E[n(t)n′(t)] = I (1)

Suppose that T (s) is a stable transfer-function with T (∞) = 0 and let e(t) represent

2

Page 3: Vibration control of Civil Engineering structures via

T(s)n(t) e(t)

Figure 1: Transfer function with white noise input

the response of T (s) when n(t) is applied to its input. Then the H2-norm of T (s) is

defined as

‖T (s)‖2 = E

(limt→∞

1

2t

∫ t

−t

‖e(t)‖2dt

)1/2

(2)

where ‖e(t)‖ denotes the Euclidean-norm of e(t), i.e. ‖e‖2 = e′e. Typically, T (s) is

an implicit function of the structure and the active controller. The H2-optimal control

problem is to choose the controller which stabilizes T (s) (internally) and minimizes

(2). The regulated (vector) signal e(t) typically includes the control effort as one of its

components.

The H2-problem is intimately related to the deterministic Linear Quadratic Regulator

(LQR) problem, which involves the minimization of:

J [u] =

∫ ∞

0

(xT Qx(t) + uT Ru + 2xT Nu)dt (3)

where x(t) denotes the system’s state-vector, u(t) is the control signal, Q = Q′ ≥ 0,

R = R′ > 0 and N are appropriate weighting matrices penalizing the state-vector

and control signal. It is well known [12] that the solution of the H2 problem can

be decomposed in two separate sub-problems. The first subproblem is to optimally

estimate the state-vector (in the mean-square sense), whose solution is provided by

Kalman-filtering theory. The second sub-problem is to find the control signal which

minimizes the deterministic cost of (3), subject to constraints in terms of the system’s

dynamics x = Ax + Bu. The solution is to let the control signal u(t) be a linear

function of the state:

u(t) = −Kcx(t) (4)

where Kc is the optimal state-feedback matrix, defined via the solution of an algebraic

3

Page 4: Vibration control of Civil Engineering structures via

Riccati equation. Then, the so-called separation principle (or certainty equivalence

principle) guarantees that the overall optimal solution of the H2 problem is still

obtained when the optimal state-feedback Kc is applied to the state estimates (obtained

from the Kalman filter), rather than the states themselves.

The H∞ optimal control problem assumes bounded-energy disturbance signals and

minimizes the maximum input-output energy transfer, given by the infinity-norm of

their transfer function, i.e.

min ‖T (s)‖∞ = minK∈S

maxω∈R

σ(T (jω)) (5)

where ‖ · ‖∞ is the infinity-norm, S denotes the set of all stabilizing controllers and

σ(·) is the largest singular value of a matrix. H∞-optimal control is essentially a

frequency-domain design methodology and is especially powerful in dealing with model

uncertainties. Thus, it can result in more robust designs than H2-optimal control,

although it can be conservative if the disturbances are naturally modelled as (filtered)

white-noise signals.

In this paper a novel approach is presented for minimizing the peak value of the

regulated signal, subject to peak magnitude and rate constraints on the control

signal. The method is developed in discrete time, using a finite-settling time (dead-

beat) parametrization, leading to a linear-programming optimization framework.

The method is particularly relevant to active vibration control of civil engineering

structures: Structural members fail after a maximum displacement is exceeded, and

thus direct optimization of peak output levels is more significant than, say, rms or

output energy levels. In addition, control constraints for systems of this type normally

arise in the form of hard saturation limits on actuator signals and their rates. Again,

using the proposed method such constraints can be directly addressed. In contrast,

in the LQG or H∞ design framework the designer can only penalize control signal

power or energy (possibly frequency weighted). Minimization of peak responses in

the context of active vibration control has been investigated by various researchers,

e.g. [10], using an adaptive bang-bang control methodology. The proposed method is

4

Page 5: Vibration control of Civil Engineering structures via

more straightforward as it relies on a fixed parameter-controller which does not require

on-line tuning. Since the problem is solved via Linear Programming, the resulting

controller will be denoted as LPOC (Linear Programming-Optimal Controller).

2. Structural model

The design algorithm is described in a step-by-step procedure. A simple benchmark

design problem from the area of active vibration control is presented alongside the

algorithm for illustration purposes.

The structure model chosen for the example employs active tendon control, since this is

reported in the literature to achieve the best results (disregarding cost considerations)

[14]. A model structure described in [13] was proposed as a benchmark problem and

has been investigated by a number of researchers. The model represents a simple and

regular 3-storey structure. A schematic of the structure is shown in Figure 2 below and

its parameters summarized in Table 1. The parameters of the linear actuator (force

constant kf , back-emf constant ke and armature resistance R) are defined in Table 2.

For simplicity only the ground and first floors are initially considered. The tendons

are connected between the ground and first floor and produce a pair of equal and

opposite forces. The structure is a scaled-down version of a real building with small

masses and dimensions, suitable for experimental work. A high value is assumed for

the base stiffness to account for the interaction between the base of the building and

the surrounding ground. The main objective of the controller is to minimize first-floor

acceleration when subjected to a force at the base.

The structure is idealized as a mass-spring-damper system shown in figure 2. In this

diagram, us is the actuator force and ν is the external-disturbance acceleration signal

(representing an earthquake) assumed to act at its base. The main design objective is

to minimize the peak value of the first regulated signal, chosen to represent first-floor

acceleration. This is equivalent to minimizing the l∞ norm of the impulse response of

5

Page 6: Vibration control of Civil Engineering structures via

Table 1: Structural parameters

Floor mi (Kg) ci (Ns/m) ki (N/m)

Base (i = 0) 5 100 16000

First (i = 1) 1.72 0.078 2600

Table 2: Actuator parameters

kf (N/A) ke (Vs/m) R (Ω)

2.0 2.0 1.5

the system, corresponding to the transfer function between the external disturbance

and first-floor acceleration. Constraints on the amplitude and rate on the actuation

signal will be subsequently imposed.

3. The design algorithm

The overall design procedure consists of the following steps:

(i) Definition of generalized plant: The two regulated signals are chosen as first-

floor acceleration and the control signal u representing actuator’s input voltage. It is

required that the controller stabilizes the system and

minK∈S

maxt≥0

|x1| (6)

subject to:

|u(t)| ≤ umax for all t ≥ 0 (7)

In addition, to avoid highly discontinuous or high-rate signals we may impose

constraints on the derivative of the control, i.e.,

|u(t)| ≤ umax for all t ≥ 0 (8)

6

Page 7: Vibration control of Civil Engineering structures via

c

k

c

k1

1

xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

0

0

xxxx

xxxx

x

x

0

1

us

xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

xxxxxxx

xxxxx

xxxxx

xxxxxxxx

xxxxxxxxxxx

xxxxxxxxxxxx

xxxxxx

m0

m1

ν

Figure 2: One-storey structure

P

K

z

z

uy

1

2

ν

Figure 3: Generalized plant

Choosing as state-variables the displacements and velocities x0, x1, x0 and x1, a state-

space description of the model is given as x = Ax + B1ν + B2u where ν denotes

the disturbance input and u is the input actuator voltage. The state-space matrices

defining the model are given as:

A =

0 0 1 0

0 0 0 1

−k0+k1

m0

k1

m0

c1m0

+kf ke

Rm0

c1m0

+kf ke

Rm0

k1

m1− k1

m1

c1m1

+kf ke

m1R− c1

m1+

kf ke

m1R

(9)

7

Page 8: Vibration control of Civil Engineering structures via

and

B1 =

0

0

1

0

, B2 =

0

0

− kf

m0R

kf

m1R

(10)

Choosing as the only measurement the first-floor acceleration signal, defines the output

equation of the system as y = Cx + Du, where

C =[−k0+k1

m0

k1

m0− c1+c0

m0+

kf ke

Rm0

c1m0

+kf ke

m0R

](11)

and

D =kf

m1R(12)

Choosing the vector of regulated signals as z = (x1 u)′, the generalized plant (see figure

3) has a state-space description:

x = Ax + B1ν + B2u (13)

z1

z2

y

=

C

0

C

x +

0

0

0

ν +

D

1

D

u (14)

Note that there is no direct feed-through term from the disturbance ν to z or y.

Signals z1 and z2 define the two regulated outputs, in this case first-floor acceleration

and control input effort u, respectively (i.e. z1 = x1 and z2 = u). Variable y represents

the measured output, in this case also first-floor acceleration (i.e. y = x1).

(ii) System discretization: The solution to the optimization problem will be

obtained in discrete-time. Thus we first need to discretize the system using an

appropriate sampling interval. The zero-order hold discretization can be employed

using the standard procedure for transforming between continuous and discrete-time

state-space models [5]. The sampling period was chosen as Ts = 0.01s. The

corresponding Nyquist frequency fN = fs/2 = 50 Hz is significantly higher than the

frequencies of all system modes. We will still use the same notation for the discrete-

time state-space realization of the generalized plant, by appropriately redefining the

state-space matrices.

8

Page 9: Vibration control of Civil Engineering structures via

(iii) Youla parametrization of all stabilizing controllers: All stabilizing

controllers and the corresponding closed-loop transfer functions between disturbance

and regulated signals can be defined in terms of two matrices F and H (stabilizing state-

feedback and output injection matrices, respectively). Matrices F and H can be any

two matrices such that A+B2F and A+HC are asymptotically stable (all eigenvalues

inside unit circle). The parametrisation proceeds by first expressing the discrete plant

G(z) as the ratio of two stable, relatively prime transfer functions. Note that the

procedure is identical in the continuous and discrete domains, with the exception that

“stability” needs to be defined appropriately in each domain. In addition, note that in

defining the parametrization we have complete freedom in the choice of state-feedback

and output injection matrices, as long as A + B2F and A + HC are asymptotically

stable; here F and H will be chosen so that all eigenvalues of A + B2F and A + HC

are placed at the origin; this is always possible under appropriate controllability and

observability assumptions, which are satisfied in this case. The algorithm is described

next:

Pole placement at the origin: The problem of pole placement is stated as follows:

Find F such that all eigenvalues of A + B2F are placed at the origin. This can

be achieved with a standard method via Ackermann’s formula [2], given that the

pair (A,B2) is controllable, by transforming it via a state-space transformation T to

controllable canonical form (Ac, Bc) and selecting Fc such that det(sI−Ac−BcFc) = sn.

The state feedback matrix F is easily obtained from the inverse transformation T−1.

This can be achieved by the following steps:

• Define characteristic polynomial of matrix A,

p(s) = det(sI − A) = sn + αn−1sn−1 + αn−2s

n−2 + ... + a0 (15)

9

Page 10: Vibration control of Civil Engineering structures via

• Define Ac, Bc in canonical controllable form:

Ac =

−αn−1 −αn−2 · · · −α1 −α0

1 0 · · · 0 0

0. . .

......

. . . 0

0 0 1 0

, Bc =

1

0...

0

(16)

Note that the first row of Ac contains the coefficients of the characteristic

polynomial in descending order with negative signs.

• Define matrices Γ, Γc, T via

Γ = [B AB A2B . . . An−1B] (17)

Γc = [Bc AcBc A2cBc . . . An−1B] (18)

and

T = Γ−1c Γ (19)

• Obtain the state-feedback matrix F as

F = [αn−1 αn−2 . . . α0]T (20)

The problem of selecting H so that A + HC has all eigenvalues at the origin is dual to

the state feedback problem described above.

The set of all stabilizing controllers can now be parametrized in bilinear (linear-

fractional) form, while the set of corresponding closed-loop systems is given in linear

(more precisely affine) form, i.e.

T (z−1) = T1(z−1)− T2(z

−1)Q(z−1)T3(z−1) (21)

where Q(z−1) is a free stable parameter. Concrete state-space realizations of Ti(z−1)

can be found in [4].

10

Page 11: Vibration control of Civil Engineering structures via

(iv) Formulation of optimisation problem in terms of linear constraints: First

partition the closed-loop equations as:

y(z−1)

u(z−1)

=

t11(z

−1)

t21(z−1)

t12(z

−1)

t22(z−1)

q(z−1)t3(z

−1)

ν(z−1) (22)

where y(z−1) and u(z−1) are the regulated output responses to a discrete pulse

ν(z−1) = 1 and tji (z−1) ∈ H∞. Note that we have also set q(z−1) = Q(z−1) to emphasize

the fact that in this case the free parameter is scalar. Equation (22) can be alternatively

written as: y(z−1)

u(z−1)

=

t11(z

−1)

t21(z−1)

t12t3(z

−1)

t22t3(z−1)

q(z−1) (23)

Hence the transfer function between ν(z−1) → y(z−1) can be written as:

T (z−1) =y(z−1)

ν(z−1)=

b(z−1) + c(z−1)q(z−1)

a(z−1)(24)

where we have defined b(z−1) = t11(z−1) and c(z−1) = −t12(z

−1)t3(z−1). Note that under

the assumptions made earlier (all eigenvalues of A + B2F and A + HC2 placed at the

origin), we have that a(z−1) = 1. The degree of both b(z−1) and c(z−1) is r, where r

denotes the number of state variables (in this example r = 4). Parametrize q(z−1) as

a finite-impulse-response filter of degree p, i.e.

q(z−1) = q0 + q1z−1 + q2z

−2 + . . . + qpz−p (25)

Also write:

b(z−1) = b0 + b1z−1 + b2z

−2 + . . . + brz−r (26)

c(z−1) = c0 + c1z−1 + c2z

−2 + . . . + crz−r (27)

y(z−1) = y0 + y1z−1 + y2z

−2 + . . . + yNz−N (28)

Then:

y(z−1) = b0 + . . . + brz−r + (c0 + . . . + crz

−r)(q0 + . . . + qpz−p) (29)

11

Page 12: Vibration control of Civil Engineering structures via

so that deg[y(z−1)] = N = r + p. The equations can be written in matrix form as:

y0

y1

...

yr

yr+1

...

yN

=

b0

b1

...

br

0...

0

+

c0 0 · · · 0

c1 c0...

.... . .

...

cr cr−1 · · · c0

0 cr · · · c1

.... . .

...

0 · · · 0 cr

q0

q1

...

qp

(30)

Note that the response is forced to be dead-beat, i.e. yr+p is the last non-zero sample

of the regulated output. This is due to the restriction on q(z−1) which is taken to be

an FIR filter and may lead to a conservative solution unless r is taken to be large.

Ideally r should be selected to make NTs, a reasonable transient before the structure

is fully stabilized. It is expected (and can be established formally) that in the limit

N → ∞ the deviation from optimality can be made arbitrarily small. The equations

can be written compactly in matrix form as y = b + Cq where vector q contains the

coefficients of the polynomial q(z−1) and where C is a Toeplitz matrix.

(v) Formulation into a linear programming problem: Since all constraints are

linear, the minimization of the peak response of the regulated signal can be formulated

as a linear programming problem of the form:

min c′x subject to Ax ≤ b (31)

Let δ be the maximum absolute value of the regulated signal (first-floor acceleration)

that we wish to minimize. Then:

−δ ≤ yk ≤ δ for all 0 ≤ k ≤ N (32)

Now yk = ck′x + bk, where c′k denotes the k-th row of the C-matrix, and bk = bk for

0 ≤ k ≤ r and bk = 0 for k > r. Thus, separating the two equations we can write:

−ck′q− δ ≤ bk and ck

′q− δ ≤ −bk (33)

12

Page 13: Vibration control of Civil Engineering structures via

for all 0 ≤ k ≤ N , which can be written in matrix form as:

−1 C

−1 −C

δ

q

b

−b

(34)

where 1 represents a column vector of ones. Setting x = (δ q)′, the problem is now in

the standard linear programming form:

min δ =[

1 0 · · · 0]x (35)

subject to (34). The solution to the problem will result in the optimal peak-value of

the regulated signal and the coefficients of the optimal q(z−1) , from which the optimal

controller can be recovered via the Youla parametrization in bilinear form.

(vi) Introducing constraints to the problem: In the above formulation, the

peak value of the regulated output is minimized for an impulsive loading without any

constraints on the size or rate of the control input. This is unrealistic and may result

in highly discontinuous control signals that would be difficult to implement or could

cause stability problems, especially in the presence of model uncertainty, due to the

potentially excessive bandwidth of the closed-loop system. The first constraint limits

the magnitude of the control signal and corresponds to the actuator’s saturation limits.

Hence we require that:

|uk| ≤ umax for all k ≥ 0 (36)

Now using Youla parametrization and the fact that the control effort has been chosen

as the second regulated output, u(z−1) may be written in the form:

u(z−1) = β(z−1) + γ(z−1)q(z−1) (37)

where we have defined β(z−1) := t21(z−1) and γ(z−1) := −t22(z

−1)t3(z−1). Note again

that β(z−1), γ(z−1) and q(z−1) are all polynomials in z−1 (the first two due to special

type of parametrization, and the third due to truncation). Similarly to the last section

13

Page 14: Vibration control of Civil Engineering structures via

the polynomial equation can be written in matrix form:

u0

u1

...

ur

ur+1

...

uN

=

β0

β1

...

βr

0...

0

+

γ0 0 · · · 0

γ1 γ0...

.... . .

...

γr γr−1 · · · γ0

0 γr · · · γ1

.... . .

...

0 · · · 0 γr

q0

q1

...

qp

(38)

where βi and γi are the coefficients of β(z−1) and γ(z−1), respectively. Writing the

equation in compact form u = β + Γq as before, and its k-th row as uk = βk + γ′kq,

the constraints |uk| ≤ umax for all k, may be expressed by a pair of linear inequalities:

−γ′kq ≤ umax + βk and − γ′

kq ≤ umax − βk (39)

for all k. Equivalently, this can be written in matrix form as: 0 −Γ

0 Γ

δ

q

umax1 + β

umax1− β

(40)

In order to make the response “smoother” an additional constraint needs to be included

limiting the rate of actuator signal, u (slew-rate constraint). Now,

∆uk = uk+1 − uk

= βk+1 + γ′k+1q− βk − γ′

kq

= (βk+1 − βk) + (γ′k+1 − γ′

k)q

and we require

|∆uk| ≤ (∆u)max for all k ≥ 0 (41)

This may be written as a pair of linear inequalities:

(γ′k+1 − γ′

k)q ≤ (∆u)max − (βk+1 − βk) (42)

and

−(γ′k+1 − γ′

k)q ≤ (∆u)max + (βk+1 − βk) (43)

14

Page 15: Vibration control of Civil Engineering structures via

for all k, or, in matrix form as: 0 −(Γ− Γ)

0 Γ− Γ

δ

q

(∆u)max1− (β − β)

(∆u)max1− (β + β)

(44)

where Γ and β denote the matrix Γ and vector β with the first row eliminated, while Γ

and β denote the matrix Γ and vector β with the last row eliminated. The inequalities

can now be augmented to the previous set of linear inequalities (40), and solved in a

linear programme to impose additional rate constraints on the control signal.

Using the constraints of the previous section, we can write:

−1 C

−1 −C

0 −Γ

0 Γ

0 −(Γ− Γ)

0 Γ− Γ

δ

q

b

−b

umax1 + β

umax1− β

(∆u)max1− (β − β)

(∆u)max1− (β − β)

(45)

which is the overall set of inequalities of the LP optimization problem.

4. Application results and Discussion

The LP design method was first applied to the structure without any control constraints

with a filter length of r = 200 samples, corresponding to a deadbeat response of

approximately 2 seconds. The two regulated signals (1st floor acceleration and actuator

voltage) are shown in Figures 4 and 5 respectively.

In order to assess the effectiveness of the LPOC controller, its responses are compared

with those obtained via Linear Quadratic Regulator (LQR) design. The design involves

a quadratic cost-function consisting of two penalty terms, acceleration and control

effort. Both weighting factors were set to 1, penalizing equally the two terms. The

design was carried out both in continuous and discrete-time (with a sampling rate of

100 Hz), producing almost identical results.

15

Page 16: Vibration control of Civil Engineering structures via

0 0.5 1 1.5 2 2.5−0.25

−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

0.2

0.251st floor acceleration

time sec

x 1 m/s

2

Figure 4: First-floor acceleration (unconstrained LP design)

0 0.5 1 1.5 2 2.5−30

−20

−10

0

10

20

30u

time sec

volts

Figure 5: Actuator voltage (unconstrained LP design)

16

Page 17: Vibration control of Civil Engineering structures via

0 0.1 0.2 0.3 0.4 0.5 0.6

−10

−5

0

5

10

1st floor acceleration

time sec

x 1 m/s

2

LPLQR

Figure 6: First floor acceleration (Constrained LP and LQR design)

The objective of LQR is to minimise the performance index defined in equation (3)

subject to plant dynamic constraints x = Ax + Bu. The terms included in the cost

function are:

• First floor acceleration x1(t) and

• Control input u(t)

resulting in an optimization index of the form:

J [u] =

∫ ∞

0

(x21 + ρu2)dt (46)

Here x1 represents acceleration of mass m1, u is the control input effort (actuator

voltage), and ρ is a penalty coefficient initially taken as 1. The performance index can

be formulated in the standard form of equation (3) by defining Q = CC ′, R = D2 + ρ

and N = C ′D.

Subsequently, the LP design was again carried out, this time with control constraints

on the peak control signal and its rate. The peak-magnitude control constraint was set

at 15 Volts, slightly less than the peak control signal obtained from the LQR simulation

(around 16 Volts) and the maximum rate constraint was set at 40 Volts/s. The two

17

Page 18: Vibration control of Civil Engineering structures via

0 0.1 0.2 0.3 0.4 0.5 0.6

−15

−10

−5

0

5

10

15

u

time sec

volts

LPLQR

Figure 7: Actuator voltage (Constrained LP and LQR design)

8 10 12 14 16 18 20 22 24 26 28 30

−200

−100

0

100

200

Unc

ontr

olle

d

8 10 12 14 16 18 20 22 24 26 28 30−100

−50

0

50

100

Acc

eler

atio

nLQ

R

8 10 12 14 16 18 20 22 24 26 28 30−100

−50

0

50

100

time (s)

LPO

C

LPOC

LQR

Uncontrolled

Figure 8: First floor acceleration (Constrained LP and LQR design for Earthquake

input)

18

Page 19: Vibration control of Civil Engineering structures via

8 10 12 14 16 18 20 22 24 26 28 30−150

−100

−50

0

50

100

150LPOC and LQR voltage (volts)

Vol

tage

(V

olts

)

8 10 12 14 16 18 20 22 24 26 28 30−150

−100

−50

0

50

100

150

time (s)

LQR

LPOC

Figure 9: Actuator voltage (Constrained LP and LQR design for Earthquake input)

Table 3: Comparison between LP and LQR methods

LQR LP Constrained LP

|x|max(m/s2) 13 0.21 5.5

|u|max(volt) 16 24 15

regulated signals resulting from the two designs (LQR and constrained LP) are shown

in figures 6 and 7. The main results of all simulations are also summarized in Table 3.

The unconstrained LP method yields excellent results in terms of optimizing the peak

signal level. The maximum acceleration is about 60 times smaller than the peak

acceleration resulting from the original LQR design, the peak voltage control level

increasing by a factor of 1.5. However, the resulting acceleration profile (Figure 12)

clearly indicates that the response is unrealistic for practical implementation. The

acceleration reaches its peak positive value of 0.21 m/s2 extremely fast and swings to

to its minimum negative value 0.21 m/s2 almost 10 ms later, requiring a huge slew-rate

from the actuator. Subsequently, the acceleration fluctuates between the two extreme

values for a few cycles of progressively increasing frequency before decaying to zero

19

Page 20: Vibration control of Civil Engineering structures via

after about 1.5 seconds (0.5 seconds earlier than the set deadbeat horizon) exhibiting

highly-damped oscillations. This behavior can be explained as follows: The maximum

acceleration is reached very fast (first peak) because the disturbance is an impulse. To

counter the acceleration increasing excessively, the controller produces a large negative

force, followed by a large positive force a few milliseconds later, to limit acceleration

increase in the opposite direction. After the maximum acceleration is reached the

controller’s primary goal is to keep it at the same level and thus gradually reduces

the applied forces. Finally, the acceleration reaches zero as the system settles to its

equilibrium. Thus the method works theoretically in the sense that it succeeds to

minimize peak acceleration, as indicated by the flat regions of the acceleration signal

at positive and negative peaks of the same magnitude. However, the response is clearly

unrealistic and thus the controller cannot be implemented in practice. The high rate

of the control signal (especially in the early part of the response) means that even if

this control profile could be generated by the actuator, the resulting closed-loop system

would have an unrealistically large bandwidth, and hence the system would have poor

stability margins and would be highly susceptible to model uncertainties.

By setting an acceptable limit in the rate of change of the control signal (|u|max = 40

Volts/s which is about ten times less than the fast rates of the early response observed

in Figure 5) the response of the system to the impulsive loading becomes acceptable.

The maximum acceleration for the constrained LP design is almost 2.5 times less than

the peak value obtained by LQR, while the controller peak signal (15 volts) is slightly

less than the peak value obtained from the LQR design (16 Volts). This improvement is

made despite the fact that the LQR controller is based on state-feedback (all four states

assumed measurable), whereas the LP controller uses output feedback only (first-floor

acceleration being the only measurement).

Next the impulsive load was replaced by a real earthquake signal consisting of the east-

west acceleration component of the Loma Prieta earthquake. The simulation results

with the LPOC and LQR controller are shown in figures 8 and 9. Note that both

rms and peak responses of the acceleration signal are reduced using LPOC control for

20

Page 21: Vibration control of Civil Engineering structures via

comparable levels of the control signal. The parameters chosen here are: Sampling

time Ts = 0.015 s, umax = 0.2 Volts, ∆umax = 0.3 Volts/s and N = 200 samples.

The LPOC technique was applied to the design of the three-storey building using the

same actuator arrangement described previously, the objective being to minimize the

1st floor accelerations which is assumed to be the only measurement. The state space

model used is given as: x = Ax + B1ν + B2u, the states being the displacement and

velocity variables x0, x1, x2, x3, x0, x1, x2 and x3. The state-space matrices A, B1 and

B2 in this case are:

A =

0 0 0 0 1 0 0 0

0 0 0 0 0 1 0 0

0 0 0 0 0 0 1 0

0 0 0 0 0 0 0 1

−k0+k1

m0

k1

m00 0 c1

m0+

kf ke

Rm0

c1m0

+kf ke

Rm00 0

k1

m1−k1+k2

m1

k2

m10 c1

m1+

kf ke

m1R− c1+c2

m1− kf ke

m1Rc2m1

0

0 k2

m2−k2+k3

m2

k3

m20 c2

m2

c2+c3m2

c3m2

0

0 0 k3

m3− k3

m30 0 c3

m3− c3

m3

(47)

and

B1 =

0

0

0

0

1

0

0

0

, B2 =

0

0

0

0

− kf

m0R

kf

m1R

0

0

(48)

Choosing as the only measurement the first-floor acceleration signal, defines the output

equation of the system as y = Cx + Du, where

C =[

k1

m1−k1+k2

m1

k2

m10 c1

m1+

kf ke

m1R− c1+c2

m1− kf ke

m1Rc2m1

0]

(49)

21

Page 22: Vibration control of Civil Engineering structures via

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

−15

−10

−5

0

5

10

15

1st floor acceleration

time sec

x 1 m/s

2

LPOCLQR

Figure 10: LQR and LPOC Acceleration of 3-storey building

and

D =kf

m1R(50)

The following values were used for the simulation: Ts = 0.015 s, umax = 0.2 Volts,

∆umax = 0.2 Volts/s and N = 200 samples. The responses compare favourably with

those obtained by LQR and are shown in Figure 10 and 11. Note that the LPOC is

capable of achieving significant reduction in peak acceleration levels using a significantly

reduced (peak) level voltage.

5. Further design considerations

The optimization problem described in previous sections minimizes the peak response

of one regulated signal (first floor acceleration) subject to peak and rate constraints of

another regulated signal (control input). Although the external disturbance is assumed

to be a unit impulse, the method can be easily modified to take into account any

disturbance of finite-duration. Assuming, for example that the external disturbance

signal has a z-transform,

ν(z−1) = ν0 + ν1z−1 + . . . + νlz

−l (51)

22

Page 23: Vibration control of Civil Engineering structures via

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−20

−15

−10

−5

0

5

10

15u

time (s)

volts

LPOCLQR

Figure 11: LQR and LPOC voltage of 3-storey building

equation (24) can be still written as:

y(z−1) = b(z−1) + c(z−1)q(z−1) (52)

by redefining b(z−1) ← b(z−1)ν(z−1) and c(z−1) ← c(z−1)ν(z−1) (and similarly for

u(z−1)).

Once the optimal “free parameter” q(z−1) is obtained in the form of an FIR filter via

the LP programme solution, the corresponding optimal controller can be recovered in

bilinear (lower linear fractional) form as:

Kopt = Fl(J, qopt) = J11 + J12qopt(I − J22qopt)−1J21 (53)

where

J =

J11 J12

J21 J22

=

A + B2F + HC + HDF −H B2 + HD

F 0 I

−(C + DF ) I −D

(54)

(see [4]). The corresponding closed loop transfer functions between the external

disturbance (ν(z−1)) and the regulated outputs (y(z−1) and u(z−1)) are obtained in

affine form via equation (23). The frequency response of the closed-loop system between

23

Page 24: Vibration control of Civil Engineering structures via

0 50 100 150 200 250 300−60

−40

−20

0

20

40

60

Angular frequency (rads/s)

Mag

nitu

de (

dB’s

)

Uncontrolled, LQR and LPOC frequency Bode plot

UncontrolledLQRLPOC

Figure 12: Bode plots of LPOC

ν(z−1) and y(z−1) in figure 12 shows that the gain is significantly reduced over all

frequencies. This suggests that the controller will be effective for arbitrary disturbance

inputs, not just an impulse.

A disadvantage of the method is that it results in high order controllers (of degree N =

r + p). This can result in a heavy computational load and implementation difficulties

and hence order-reduction techniques should be applied for practical purposes. Suitable

approximation techniques include balanced truncation or Hankel-norm approximation

methods [18], [7]. In this case, since the bulk of the controller complexity is due to the

high degree of the FIR filter q(z−1), model reduction techniques developed specifically

for systems of this type may be more appropriate. A Hankel-norm approximation

method of FIR systems by low-order infinite-impulse-response (IIR) systems was

developed in [8] based on a model-reduction technique applicable to general discrete-

time descriptor systems [1]. A nice aspect of this approach is that the reduced order

controller is guaranteed to be stabilizing (due to Youla parametrization), irrespective

of the approximation order. However, the performance properties of the design may

deteriorate, especially for approximations of a low degree.

24

Page 25: Vibration control of Civil Engineering structures via

6. Conclusions

The paper presents a LP-based algorithm aiming to minimize the peak value of a

regulated signal, an objective which is especially relevant for the design of active

vibration control of civil engineering structures. Linear constraints are introduced

to limit the magnitude of the control signal and its rate, resulting in “smooth”

responses and low-bandwidth control schemes which can be implemented in practice.

The design algorithm was developed in parallel to a simple example involving a

scaled-down scalar benchmark model of a one-storey building, although extensions

to the multivariable case and multiple regulated signals are straightforward. It was

demonstrated via simulations that the design method is capable to reduce significantly

the peak acceleration response of the model compared to LQR designs, even after the

introduction of constraints on the control-signal. Other advantages of the method

include the ability to formulate realistic constraints involving the magnitude and

rate of regulated signals (rather than rms or energy content) and to provide indirect

control of the overall damping by specifying the settling-time horizon. Although the

disturbance signal was assumed to be an impulse, more general disturbance models

can be accommodated.

A number of issues related to the design require further investigation. These include

a full robustness analysis and the possibility of incorporating the method within a

larger multi-objective optimization framework (e.g. using multiple regulated signals

and a mixture of linear and quadratic constraints, which can be tackled via quadratic

programming). Another important issue is related to controller complexity. The design

method tends to produce high-order controllers, in the form of a bilinear transformation

of a high-order FIR filter. This can be approximated by a low-order IIR filter resulting

in an overall low-order controller (which is still stabilizing) using a recently derived

Hankel-norm model-reduction algorithm for discrete-time descriptor systems [1], [8].

The approximation order should be chosen so that small magnitude and phase errors

are introduced in the controller’s frequency response, especially in the cross-over region

which determines the gain and phase margins of the feedback loop. Alternatively,

25

Page 26: Vibration control of Civil Engineering structures via

direct closed-loop approximation techniques with quantifiable measures of performance

deterioration can be considered [6].

Alternative control design methods aiming to minimize the peak response of the

regulated signal have recently been reported both in the area of active vibration

control [10] and also in the general control literature [16], [17], [3]. Reference [10]

is based on an adaptive bang-bang methodology, which clearly offers advantages in

the case of uncertainty about the disturbance-signal, but is also difficult to apply

in practice. A systematic general approach is l1 optimal control, which attempts to

minimize the peak amplification gain between disturbance input and regulated output

[3], [11]. Interestingly, the method also results in a Linear Programming optimization

framework. However, as l1 is an induced norm, all bounded disturbance signals are

taken into account in the formulation of the optimization problem. As a result, the

design may be conservative, unless the method can be restricted to specific models of

disturbance signals that are likely to arise in practice, i.e. signal classes whose spectral

content is similar to typical seismic acceleration signals. Assessing the potential merits

of this method relative to the proposed approach needs further investigation.

References

[1] M.M. Al-Hussari, I.M. Jaimoukha and D.J.N. Limebeer, A descriptor approach for

the solution of one-block distance problems, Proc. IFAC World Congress, Sydney,

Australia, 1993.

[2] P.J. Antsaklis and A.N. Michel, Linear Systems, McGraw-Hill, 1998.

[3] M.A. Dahleh and J.J. Diaz Bobillo, Control of uncertain systems: A linear

programming approach, Prentice-Hall, Englewood-Cliffs, New Jersey, 1995.

[4] B.A. Francis, A course in H∞ optimal control theory, Springer Verlag, Lecture

Notes in Control and Information Sciences, New York, 1987.

26

Page 27: Vibration control of Civil Engineering structures via

[5] G.F. Franklin, J.D. Powell and M.L. Workman, Digital Control of Dynamical

Systems, Addison-Wesley, Reading, Massachusetts, 1990.

[6] P.J. Goddard and K. Glover, Performance preserving controller approximations,

IEEE Trans. Auto. Control, Vol. 43, No 7, July 1998.

[7] M. Green and D.J.N. Limebeer, Linear Robust Control, Prentice Hall, Englewood

Cliffs, NJ, 1995.

[8] G.D. Halikias, I.M. Jaimoukha and D.A. Wilson, A numerical solution to the matrix

H2/H∞ optimal control problem, Int. Journal of Robust and Nonlinear Control, Vol.

7, No 7, pp. 711-726, July 1987.

[9] G.W. Housner, L.A. Bergman, T.K. Caughey, A.G. Chassiakos, R.O. Claus, S.F.

Masri, R.E, Skelton, T.T. Soong, B.F. Spencer and J.T.P.Yao, Structural control:

Past, present and future, Journal of Engineering Mechanics, Sept. 1997

[10] C. W. Lim, T. Y. Chung, S. J. Moon, Adaptive bang-bang control for the vibration

control of structures under earthquakes, Journal of earthquake engineering and

structural dynamics, pp.1977-1994, July 2003.

[11] J.S. MacDonald and J.P. Pearson, l1 Optimal Control of multivariable systems

with output norm constraints, Automatica, Vol. 27, pp. 317-329, 1991.

[12] J.M. Maciejowski, Multivariable Feedback Design, Addison Wesley Publishing

Company, 1989.

[13] H. Nishimura and Akihito Kojima. Seismic Isolation Control, IEEE control

Systems, 1991.

[14] T.T. Soong, Active structural control: Theory and Practice, Longman Scientific

and Technical, 1990.

[15] B.F. Spencer, Jr. and M.K. Sain Controlling Buildings: A new Frontier in feedback

IEEE Control Systems Magazine, 17 (6), 19-35, 1997.

27

Page 28: Vibration control of Civil Engineering structures via

[16] M. Sznaier, T. Amishama and T. Inanc, H2 control with domain constraints:

Theory and applications, IEEE Transactions on Automatic Control, Vol. 48, No

3, March 2003.

[17] M. Sznaier, A mixed l∞/H∞ optimization approach to robust controller design,

SIAM Control and Optimization, Vol. 33, No 4, pp. 1086-1101, July 1995.

[18] K. Zhou, J.C. Doyle and K. Glover, Robust and Optimal Control, Prentice Hall,

Englewood Cliffs, NJ, 1996.

28